Sie sind auf Seite 1von 8

Materials Science & Engineering A 582 (2013) 389–396

Contents lists available at SciVerse ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

Effect of tungsten content on microstructure and mechanical


properties of swaged tungsten heavy alloys
U. Ravi Kiran a,b,c,n, S. Venkat a, B. Rishikesh b, V.K. Iyer c, M. Sankaranarayana a,b,c,
T.K. Nandy a,b,c
a
Defence Metallurgical Research Laboratory, Kanchanbagh, Hyderabad 500058, India
b
Defence Research and Development Laboratory, Kanchanbagh, Hyderabad 500058, India
c
Heavy Alloy Penetrator Plant, Tiruchirapalli 620025, India

art ic l e i nf o a b s t r a c t

Article history: This paper describes the effect of tungsten content on microstructure and mechanical properties of swaged Co-
Received 6 March 2013 containing tungsten heavy alloys with varying tungsten (90W–7Ni–2Fe–1Co, 93W–4.9Ni–1.4Fe–0.7Co and
Received in revised form 95W–3.5Ni–1Fe–0.5Co). With increasing tungsten while tensile strength goes through a maximum, both
10 June 2013
percent elongation and impact energy decrease. Microstructure and fractographic analyses have been carried
Accepted 11 June 2013
Available online 20 June 2013
out in order to explain the trends in mechanical properties. Predominant transgranular fracture of tungsten
grains combined with ductile dimple failure of the matrix in 90% W alloy is responsible for superior properties
Keywords: of this alloy in comparison to the alloy with 95% W. The highest tensile strength attained in 93% W alloy is
Tungsten heavy alloy attributed to predominant cleavage failure of W-grains. The results clearly indicate that the matrix volume
Swaging
fraction, contiguity and matrix mean path greatly influence the mechanical properties of tungsten heavy alloys.
Tensile properties
& 2013 Elsevier B.V. All rights reserved.
Impact strength
Fractography

1. Introduction alloying additions. Katavic [10] explained that the mechanical


properties of W–Ni–Fe–Co are far superior to those of W–Ni–Fe
Tungsten heavy alloys (WHA) are two-phase composites com- alloys. In a recent study, we demonstrated that a Co-containing W
prising nearly rounded tungsten grains dispersed in a low melting alloy exhibits significantly improved mechanical properties by
point ductile matrix containing Fe, Ni, Cu and Co [1–3]. Tungsten heat treatment plus swaging deformation [11].
heavy alloys are currently the key materials for kinetic energy Increasing tungsten results in increased density of the alloy
penetrator applications due to their high density, strength and leading to higher depth of penetration. However, the alloys
good ductility [4]. While the high density of tungsten heavy alloy containing higher tungsten exhibit inferior mechanical proper-
helps in realization of higher depth of penetration, their excellent ties [12]. Since cobalt addition leads to improvement in proper-
mechanical properties such as high strength, ductility and impact ties, it was decided to study the effect of tungsten on
property ensure the survival of penetrator from the rigors of microstructure and mechanical properties of cobalt containing
extremely demanding conditions experienced during their launch tungsten heavy alloys and also correlate mechanical properties
and terminal ballistics [5]. The as-sintered WHAs are generally with underlying microstructure. Although W alloys with
vacuum-annealed followed by oil quenching to remove trapped
improved mechanical properties were successfully developed
hydrogen, ensure homogenization of the matrix and also reduce
[9–12], limited results are available on the effect of tungsten
the interfacial segregation of impurities [6–8]. Mechanical proper-
content on tensile and Charpy impact properties in the thermo-
ties of the penetrator blanks can be further improved by thermo-
mechanically processed condition [13]. In the present investiga-
mechanical processing that involves repeated heat treatment and
swaging. Islam et al. [9] conducted experiments on heat treated tion, WHAs that comprise about 90–95 wt% W and balance
W–Ni–Fe alloys and showed that the binding strength between W matrix alloy containing Ni, Fe and Co have been investigated in
and the matrix phase has a major influence on the ductility of detail with respect to microstructure, tensile properties and
WHAs. This can be addressed by either heat treatment or suitable impact toughness. Three different compositions with tungsten
content of 90, 93 and 95 wt% have been investigated in this
study. In all alloys, the ratio of Ni, Fe and Co that preferentially
n
Corresponding author at: Defence Metallurgical Research Laboratory,
partition to matrix was kept constant. The effect of relative
Kanchanbagh, Hyderabad 500058, India. Tel.: +91 9490956731;
fax: +91 4024344535. volume fraction of tungsten and the matrix phase on mechanical
E-mail address: uravikiran@rediffmail.com (U. Ravi Kiran). properties of heavy alloys have also been discussed.

0921-5093/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2013.06.041
390 U. Ravi Kiran et al. / Materials Science & Engineering A 582 (2013) 389–396

2. Experimental [16] and evaluation was carried out in a 300 J  2 J pendulum type
impact tester at a velocity 5.24 m/s at room temperature. At least
The nominal composition of the alloys prepared in the present five tests were performed for each alloy and the results of the
study is listed in Table 1. Commercially available elemental impact tests are average of 5 data points. Fractography of the failed
tungsten, nickel, iron and cobalt powders, with their purity and tensile and impact test pieces were carried out using Scanning
average particle size listed in Table 2, were used as starting electron microscope (Make: FEI, Quanta-400, Netherlands). Bulk
materials for synthesizing liquid phase sintered WHAs. The pow- hardness of the alloys was measured using Vickers hardness tester
ders were mixed in a ball mill, and reduced at 700 1C in H2
atmosphere for 2 h. The reduced powders were cold iso-statically
pressed into cylindrical rods of 45 mm diameter and 300 mm
length. Pre-sintering of the rods was carried out at 900 1C for 2 h,
in hydrogen sintering furnace and sintering was conducted at
1460 1C for 2 h in hydrogen atmosphere. The tungsten heavy alloy
blanks were subjected to warm die swaging (500 1C) from 38 mm
to 32 mm diameter followed by heat treatment at 1150 1C/2 h/oil
quench. Subsequently another swaging was carried out involving a
reduction in diameter from 32 mm to 28 mm thereby imparting a
total deformation (first plus second swaging) of 41% with an
intermediate heat treatment.
Samples were prepared for microstructural evaluation by stan-
dard metallographic procedures involving cutting, mounting, grind- 50µm
ing and polishing. The size of the tungsten particles, the volume
fraction of the matrix phase, contiguity and matrix mean path of the
sintered WHAs were measured from optical micrographs taken
from several locations of a specimen. The volume fraction was
measured using an image analyzing software. Contiguity defined as
the relative fraction of W–W interfacial area was determined by
placing grid lines over the binary image and counting the number
of tungsten–tungsten and tungsten–matrix intercepts. It was calcu-
lated using the equation CSS ¼ 2NSS/(2NSS+NSL), where NSS and NSL
are the number of W–W grain boundary and W–matrix interface
boundary intercepted, respectively. The matrix mean path width is
defined as the total matrix area divided by two times the difference
between the total length of W grain perimeters and the total length
of W–W grain contacts [14]. 50µm
The X-ray mappings of elements like W, Ni, Fe and Co were
done using an electron probe micro analyzer (EPMA) (Make:
Sx-100, Cameca, France). The matrix composition of all the alloys
was also determined by EPMA. All swaged samples were subjected
to tensile testing. Three specimens were used for each measure-
ment. The tensile specimens were prepared as per ASTM standard
E8M-04 [15] and tested using a Masanto Tensometer. Charpy
impact specimens were prepared as per ASTM standards (E23)

Table 1
Chemical composition of the investigated tungsten heavy alloys.

Constituent WNFC1 WNFC2 WNFC3

W 90 93 95 50µm
Ni 7 4.9 3.5
Fe 2.0 1.4 1
Co 1.0 0.7 0.5 Fig. 1. Optical micrographs of (a) WNFC1, (b) WNFC2, and (c) WNFC3 tungsten
heavy alloys.

Table 2
Characteristics of the elemental powders used in the present study.

Powder W Ni Fe Co

Purity (wt%) 99.9 99.6 99.6 99.6


Average particle size (mm) 11 6 5 4
Major impurities (wt%)
C o 0.01 o 0.5 o 0.05 o 0.02
O o 0.1 o 0.05 o 0.1 o 0.05
P – o 0.004 o 0.05 –
Apparent density (g/cm3) 4.81 1.03 3.23 2.23
Tap density (g/cm3) 5.84 1.36 3.53 2.81
Flow rate (g/s) Non-flowable Non-flowable Non-flowable Non-flowable
U. Ravi Kiran et al. / Materials Science & Engineering A 582 (2013) 389–396 391

(Vickers Instruments, Model: 1965, UK). All tests were conducted WNFC3 alloy which may be due to shape accommodation since
at a load of 30 kg and indentation time of 30 s was maintained. the volume fraction of W phase is higher. Another noticeable
The Vickers hardness value for each sample was an average of five feature in WNFC2 and WNFC3 alloys is the presence of flat
readings taken at random locations throughout the sample of W–matrix interfaces which can be again attributed to shape
10  10 mm2 area. Micro-hardness of the tungsten grains and the accommodation effect as the particles coarsen during liquid phase
matrix phase in as swaged condition were measured separately by sintering. Usually the grains below a critical size tend to shrink and
micro-hardness tester (Make: MAT-SUZAWA, Japan). dissolve while grains above the critical size grow at the expense of
dissolving finer grains. Thus, grain growth is proportional to the
total net flux, Jt, of tungsten atoms between two particles. Grain
3. Results and discussion growth (Jt) can be expressed as follows [17]:
Bw C w
3.1. Microstructure Jt ¼ − ΔGr ð1Þ
r*
Optical microstructures of the heat treated alloys WNFC1 where Bw and Cw are the mobility and concentration of W atoms
(90W–7Ni–2Fe–1Co), WNFC2 (93W–4.9Ni–1.4Fe–0.7Co) and respectively in the liquid matrix phase; rn the effective distance of
WNFC3 (95W–3.5Ni–1Fe–0.5Co) are shown in Fig. 1(a–c). It can diffusion; and ΔGr the molar free energy difference between large
be seen that the W particles in Fig. 1a are smaller and more and small W particles. Coarser grains in WNFC3 alloy is due to
roundish than those in Fig. 1b and c. Also the particles are grain growth that is facilitated by relatively shorter diffusion
interconnected, more so in Fig. 1c. The amount of matrix phase distance of W atoms in the matrix whose volume fraction is low.
is the highest in 90% tungsten alloy (WNFC1) and it decreases with To determine the tungsten concentration in the matrix phase of
increasing tungsten (Table 3). Irregular tungsten grains are seen in the alloys, electron microprobe analysis of the phases (Fig. 2) was
carried out. Results of the analysis given in Table 4 show that the
Table 3
Microstructural parameters of tungsten alloys. Table 4
Chemical analysis (wt%) of matrix phase in tungsten alloys.
Alloy Grain size (mm) Volume fraction Contiguity Matrix mean
(tungsten) (matrix) path (mm) Alloy W Ni Fe Co

WNFC1 36 0.34 0.42 7.5 WNFC1 32.7 46.1 13.6 7.48


WNFC2 47 0.22 0.55 5.5 WNFC2 28.8 49.5 13.4 8.56
WNFC3 59 0.16 0.73 4.2 WNFC3 24.7 45.7 27.3 2.2

W-BSE Ni Fe Co W

W-BSE Ni Fe Co W

W-BSE Ni Fe Co W

Fig. 2. Elemental mapping of WNFM tungsten heavy alloy: (a) WNFC1, (b) WNFC2, and (c) WNFC3.
392 U. Ravi Kiran et al. / Materials Science & Engineering A 582 (2013) 389–396

tungsten solubility in the matrix phase decreases with increase function of tungsten content. Grain size and contiguity increase
in W content of the alloys. The volume fraction (Fig. 3) of the with increase in tungsten content. Increase in contiguity of
tungsten grains measured from the image analysis reveals that as W/W grains with increase in volume fraction of W is well known
the W content increases from 90% (WNFC1) to 95% (WNFC3), the [18–20]. Swaging to 41% deformation leads to increase in the
volume fraction of the tungsten grains increases from 70% to about aspect ratio of the tungsten grains (Fig. 5) in all the alloys, but the
86%. Fig. 4 shows the tungsten grain size and contiguity as a microstructural parameters such as contiguity and volume fraction
remain similar.

3.2. Mechanical properties

The yield strength, ultimate tensile strength, ductility, impact


and hardness of the three alloys are listed in Table 5. The bulk
hardness of the alloys is found to increase marginally with
increasing tungsten content. It is observed that the microhardness
of the tungsten grains is same in all the three alloys, while the
micro-hardness of matrix phase in WNFC1 alloy (428 HV0.1) is
slightly higher than those of other alloys (423 HV0.1 and 416 HV0.1).
This marginally lower microhardness in high tungsten alloys may
be ascribed to softening of the matrix phase due to decrease in W
content. The difference in hardness, however, is much less pro-
nounced as compared to the difference in tensile and elongation
values of the alloys. Stress–strain curves are shown in Fig. 6. It is
clear as the tungsten content increases, percent elongation
decreases while tensile and yield strength vary only marginally.
Fig. 7 shows the effect of tungsten content on yield strength, UTS
Fig. 3. Volume fraction and matrix mean path of tungsten alloys.
and percent elongation. While yield strength increases, UTS goes
through a maximum at 93% W. Yield strength dependence can be
explained in terms of the following equation that is bassed on
classical Hall–Petch equation [21]:
 
1−V M 1=2
sy ¼ so þ kGb ð2Þ
DV M

where sy is the yield strength, so is the intrinsic strength, k is a


constant, G is the shear modulus, b is the Burgers vector, D is the
tungsten grain size and VM is the matrix volume fraction. In this
equation, the matrix volume fraction (VM) is expected to play a major
role. As W content increases, VM decreases (Table 3) thereby
increasing the yield strength. Thus, increase in yield strength can
be ascribed to decreasing mean free path in the matrix phase that is
attributed to decreasing volume fraction of the matrix phase with
increasing tungsten. UTS depends upon three factors: (1) volume
fraction of tungsten (that in turn decides the mean free path),
(2) work haderning during plastic defomation, (3) percent elonga-
tion. Increase in any of these parameters with others remaining
constant will lead to increase in the UTS value. Assuming that work
hardening does not vary much in this composition range, UTS
dependence can be understood in terms of W volume fraction
Fig. 4. Grain size and contiguity variation in tungsten alloys. and percent elongation. With inreasing tungsten, while the volume

30µm 30µm 30µm

Fig. 5. Scanning electron micrograph of swaged (a) WNFC1, (b) WNFC2, and (c) WNFC3 tungsten alloys.
U. Ravi Kiran et al. / Materials Science & Engineering A 582 (2013) 389–396 393

Table 5
Mechanical properties of swaged tungsten heavy alloys.

Alloy Yield strength (MPa) Tensile strength (MPa) Yield ratio (%) Elongation (Pct) Impact energy (J) Bulk hardness (VHN) Micro-hardness

W Matrix

WNFC1 1360 7 4 14007 5 97.1 107 1 65 528 75 530 7 3 428 72


WNFC2 14107 5 14357 3 98.2 77 1 45 532 74 529 7 4 423 71
WNFC3 14187 5 1420 7 3 99.8 47 1 14 538 73 529 7 2 416 73

Fig. 8. Impact and percent elongation as a function of contiguity.


Fig. 6. Stress–strain curves of swaged WNFC1, WNFC2 and WNFC3 tungsten alloys.

reason for this behavior is elaborated further in the following


paragraph.
The impact energy values have dropped from 65 J to 14 J with
increase in tungsten content. It appears there is one to one
correlation between percent elongation and impact. With increas-
ing tungsten, the contiguity which is a direct indication of W–W
contact increases. Since W–W contact are the weakest constituent
in the microstructure [23], the failure initiates at this sites. Fig. 8.
shows impact and percent elongation as a function of contiguity.
In both cases, the trend is similar pointing towards a common
mechanism of failure.
SEM tensile fractographs of the WHAs are shown in Fig. 9.
Fracture surface in alloys is characterized by mainly by four
distinctive failure features such as (a) tungsten cleavage,
(b) matrix rupture, (c) tungsten–tungsten intergranular failure
and (d) tungsten–matrix interfacial separation. We attempt to
provide of a quantitative analysis of the fracture features by
determining relative fraction of different fracture modes. For
example relative contribution of W cleavage in any fractograph
Fig. 7. Effect of tungsten content on tensile properties of tungsten alloys.
shown in Fig. 9 is given by the following equation:
% W cleavage ¼ No: of cleaved W grains=ðNo: of cleaved Wgrains

fraction of W phase increases, percent elongation decreases. þNo: of W  matrix failure


These two opposing effects leads to attainment of a UTS maximum þNo: of W−W failure þ No: of matrix failureÞ
at intermediate W-content. An empirical correlation between
ð4Þ
the elongation to failure and microstructures of tungsten alloys have
been proposed to explain the trend in ductility. The equation is as Counting is done by using a linear intercept method (similar to
follows [22]: the determination of contiguity). Table 6 gives the quantitative
analysis of the fracture modes. Corresponding trend as a function
ε ¼ k1 þ k2 V M ð1−Cw Þ ð3Þ of tungsten is shown in Fig. 10. The high strength in WNFC2
tungsten alloys are mainly associated with the higher fraction of
where ε is the elongation to failure, k1 and k2 representing a tungsten cleavage fracture surface, whereas reduced strength in
collection of material constants and Cw is the tungsten–tungsten WNFC3 alloys appears to be due to tungsten–matrix and tungsten–
contiguity. It is evident from Eq. (3) that failure strain decreases as tungsten interface failure. As mentioned earlier, the failure of the
W–W contiguity increases with increasing W content (Table 3). The W–W and W–matrix boundaries is generally associated with low
394 U. Ravi Kiran et al. / Materials Science & Engineering A 582 (2013) 389–396

50µm

W-W Fig. 10. Relative fraction of fracture surfaces in tungsten alloys tested under
tension.
W

50µm

W
W-M
50µm

W-W

50µm

Fig. 9. Scanning electron micrographs of tensile fracture surfaces of alloys


(a) WNFC1, (b) WNFC2, and (c) WNFC3 tungsten heavy alloys. W–tungsten
cleavage, M–Matrix rupture, W–W–tungsten–tungsten grain boundary separation,
and W–M–tungsten–matrix interface separation.

50µm
Table 6
Percent of fracture surfaces (tensile specimens) in the swaged tungsten heavy
alloys.

Alloy Tungsten Matrix Tungsten– Tungsten–


cleavage (W) rupture (M) Tungsten (W–W) matrix (W–M)

WNFC1 29 64 5 2
WNFC2 36 52 8 4
WNFC3 24 40 20 16

strain and therefore lower tensile strength [24]. Now the relative
proportion of W–W grain boundaries in the fracture surface
depends only on the amount of tungsten. Thus, while higher
ductility in WNFC1 is mainly because of relatively larger propor- 50µm
tion of tungsten cleavage and matrix rupture, more frequent
de-cohesion of W–W and W–M interfaces in WNFC3 is responsible Fig. 11. SEM micrographs of Charpy impact fracture surfaces in (a) WNFC1,
for lower ductility. Thus, the decrease in strength and ductility in (b) WNFC2, and (c) WNFC3 tungsten heavy alloys.
U. Ravi Kiran et al. / Materials Science & Engineering A 582 (2013) 389–396 395

alloys with W greater than 93% tungsten points towards a limita-


tion imposed by microstructure with increasing tungsten content.
Figs. 11 and 12 show the SEM micrographs of impact fractured
surfaces of WNFC1, WNFC2 and WNFC3 tungsten heavy alloys
tested at room temperature. The lower impact energy obtained for
WNFC3 samples is found to be also associated with a cleavage of
tungsten grains with numerous fine secondary cracks due to low
volume fraction of matrix (Fig. 11c). In addition to W–W contiguity,
tortuosity of cracks may be also able explain the trend in impact
value that decreases with increasing tungsten. As the volume
fraction of the matrix phase decreases, the regions of high energy
10µm failure decreases and probability that a crack gets deflected once it
interacts with such features comes down. Thus while tortuosity of
crack is expected to be more in WNFC1 and WNFC2 because of the
higher volume fraction of the matrix phase, linear crack propaga-
tion is expected in WNFC3 that contains coarse tungsten grains
with thin matrix film (Fig. 12). As a result, the observed energy in
WNFC2 and WNFC3 alloys is less than that of WNFC1 alloy. Fig. 13
shows impact toughness and ductility plots as a function of matrix
mean free path. It is clearly seen that higher matrix mean path
containing alloy shows higher impact and elongation to failure. In
summary, it is concluded the matrix volume fraction, contiguity
and matrix mean path have a profound effect on tensile and
impact properties of WHAs.
10µm

4. Conclusions

The microstructure and mechanical properties of swaged 90, 93


and 95 wt% tungsten heavy alloys were investigated. 90% heavy
alloys showed higher elongation and impact as compared to alloys
containing higher W due to relatively higher matrix volume
fraction, larger matrix mean path and low contiguity. On the other
hand 93% alloys exhibited the highest tensile strength (about
1435 MPa) which has been attributed to higher fraction of tung-
sten cleavage. 95% alloys showed the lowest elongation and impact
properties primarily because increased contiguity and reduced
volume fraction of the matrix phase. Thus, mechanical properties
10µm of cobalt containing tungsten heavy alloys were found to be
strongly dependent on microstructural parameters such as matrix
volume fraction, contiguity and matrix mean path, which are
Fig. 12. SEM micrographs of matrix fracture surfaces in (a) WNFC1, (b) WNFC2, and
(c) WNFC3 tungsten heavy alloys. governed by overall tungsten content.

Acknowledgments

Authors convey their sincere gratitude to Dr. Amol Gokhale,


Outstanding Scientist & Director, DMRL for encouragement and
kind permission to publish this work. Authors gratefully acknowl-
edge the financial support provided by DRDO. They would also like
to thank the staff members of Powder Metallurgy Group, DMRL for
their help in experimentation. Thanks are also due to Shri Bijoy
Sarma, Division Head, Powder Metallurgy Division (retired) for his
helpful suggestions and guidance to carry out this investigation.

References

[1] A. Upadhyaya, Mater. Chem. Phys. 67 (2001) 101–110.


[2] H.J. Ryu, S.H. Hong, W.H. Baek, Mater. Sci. Eng. A 291 (2000) 91–96.
[3] B. Huang, J. Fan, S. Liang, X. Qu, J. Mater. Process. Technol. 137 (2003) 177–182.
[4] J.R. Li, J.L. Yu, Z.G. Wei, Int. J. Impact Eng. 28 (2003) 303–314.
[5] S. Pappu, C. Kennedy, L.E. Murr, L.S. Magness, D. Kapoor, Mater. Sci. Eng. A 262
(1999) 115–128.
[6] F.E. Sczerzenie, H.C. Rogers, Hydrogen in MetalsASM, New York, 1974.
[7] L.G. Bazenova, A.D. Vasiljev, R.V. Minakova, V.J. Trefilov, Powder Metall. Int. 1
(1980) 45–51.
[8] B. Katavic, Sci. Sinter. 32 (2000) 153–166.
Fig. 13. Impact and elongation as function of matrix mean path. [9] S. Islam, F. Humail, M. Tufail, Int. J. Refract. Hard Mater. 25 (2007) 380–385.
396 U. Ravi Kiran et al. / Materials Science & Engineering A 582 (2013) 389–396

[10] B. Katavic, M. Nikacevic, Proceedings of the 2nd International Conference, [17] K. Eun-Pyo, H. Moon-Hee, H.I. Moon, Metall. Mater. Trans. A 30A (1999) 627.
Serbia and Montenegro, Belgrade, 2005, pp. 135–140. [18] J.L. Johnson, R.M. German, Metall. Mater. Trans. A 27B (1996).
[11] U.Ravi Kiran, M. Sankaranarayana, T.K. Nandy, Int. J. Refract. Hard Mater. 33 [19] W. Yunxin, R.M. German, R. Bollina, Mater. Sci. Eng. A 344 (2003) 158–167.
(2012) 113–121. [20] S. Farooq (Ph.D. thesis), Rensselaer Polytechnic Institute, 1998.
[12] R.M. German, L.L. Bourguignon, B.H. Rabin, J. Met. (1985). [21] M.F Ashby, Philos. Mag. 18 (1970) 399.
[13] K.S. Churn, J.W. Noh, H.S. Song, E.P. Kim, W.H. Baek, Proceedings of the [22] K.S. Churn, R.M. German, Metall. Trans. A 15A (1984) 331–338.
International Conference, M.P.I.F, 1992, p. 397. [23] R.M. German, J.E. Hanafee, in: R.M. German, K.W. Lay (Eds.), Processing of
[14] T. Weerasooriya, P. Moy, ARL-TR-1694 (1998). Metal and Ceramic Powders, TMS-AIME, Warrendale, 1982, pp. 267–282.
[15] ASTM E8-04, Test Methods for Tension Testing of Metallic Materials, 2004. [24] L. Ekbom, in: H.H. Hausner, H.W. Antes, G.D. Smith (Eds.), Modern Develop-
[16] ASTM E23-0a. Standard Test Methods for Un-Notched Bar Impact Testing of ments in Powder Metallurgy, vol. 14, Princeton, NJ1981, pp. 177–188.
Powder Structural Parts, 2002.

Das könnte Ihnen auch gefallen