Você está na página 1de 7

Journal of Alloys and Compounds 619 (2015) 513519

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Effects of deep cryogenic treatment on the microstructure and


mechanical properties of commercial pure zirconium
Chao Yuan, Yunpeng Wang, Deli Sang, Yijun Li, Lei Jing, Ruidong Fu , Xiangyi Zhang
State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao, Hebei 066004, PR China
College of Materials Science and Engineering, Yanshan University, Qinhuangdao, Hebei 066004, PR China

a r t i c l e i n f o a b s t r a c t

Article history: The effects of deep cryogenic treatment (DCT) on the microstructure and mechanical properties of com-
Received 25 April 2014 mercial pure zirconium were investigated. Experimental results indicated that DCT induced a change in
Received in revised form 3 August 2014 grain orientation and improved internal stress, which in turn increased dislocation density that led to
Accepted 26 August 2014
improved hardness. Hardness in basal planes was found to be signicantly larger than that in prism
Available online 16 September 2014
planes. Moreover, strength was enhanced in DCT-treated zirconium and the ductility was comparable
to that of as-annealed zirconium. This phenomenon was due to the increase in dislocation density and
Keywords:
the good ductility resulting from the motion of pre-existing dislocations and specic dislocation cong-
Metals and alloys
Deep cryogenic treatment
urations. DCT led to the transformation of tensile fracture mode from mixed-rupture characteristics of
Microstructure quasi-cleavage and dimples to quasi-cleavage, thereby increasing compatible deformation capabilities.
Mechanical properties The possible mechanisms underlying microstructural modication, tensile strength, and hardness
Zirconium improvement were discussed.
2014 Elsevier B.V. All rights reserved.

1. Introduction persed carbides, and removal of residual stresses [6,7,12]. Previous


studies have shown that the hardness and abrasion resistance of
Cryogenic treatment is a very old process that is widely used for DCT-treated samples evidently improve because of transformation
high-precision components. Shallow cryogenic treatment is set at of abundant retained austenite to martensite, secondary carbide
low temperatures (about 80 C), whereas deep cryogenic treat- precipitation [13], and the precipitation of nanosized g-carbides
ment (DCT) is set at near-liquid-nitrogen temperatures (about in primary martensite [14]. In addition, DCT combined with tem-
196 C). DCT improves certain properties beyond the enhance- pering treatment enables the enhancement of fatigue properties
ment obtained by normal cold treatment [13]. DCT is different for the precipitation of ne carbides and the reduction in compres-
from those methods of severe plastic deformation (SPD) processing sive residual stress [15].
at cryogenic conditions, such as cryogenic rolling, pinning and Compared with studies on ferrous metals, research on the effect
milling. Instead of nanostructured and ultrane-grains in SPD-pro- of DCT on nonferrous metals such as Mg, Al, and Ti alloys are lim-
cessed metals, the variation of crystal defect and phase transforma- ited. Asl et al. [2] found that after DCT, tiny laminar b phase parti-
tion may play more important role in increasing of the properties cles almost dissolve, and the coarse divorced eutectic b phase
of DCT-treated metals. Over the past few decades, the effects of extends into the neighboring matrix. As a result, the mechanical
various cryogenic treatments on the performance of steel have properties of AZ91 Mg alloy signicantly improve. Jiang et al.
received considerable interest. DCT reportedly increases the nor- [16] found that DCT induces the renement of grains to approxi-
mal temperature strength and hardness of steels [4,5], provides mately 0.13.0 lm, thereby improving the strength and elongation
dimensional stability or microstructural stability [6], and improves of 3102 Al alloy foil.
wear [79] and fatigue resistance [10,11]. The improvement in As aforementioned, evident changes in microstructure such as
mechanical properties can be ascribed to the complete transforma- phase transformation or grain renement have been veried for
tion of retained austenite into martensite, precipitation of ne dis- DCT-treated alloys. However, changes in microstructure and
mechanical properties that can be introduced by DCT if no phase
Corresponding author at: State Key Laboratory of Metastable Materials Science transformation occurs are worth investigating. In this study,
and Technology, Yanshan University, Qinhuangdao, Hebei 066004, PR China. Tel.: coarse-grained zirconium (Zr) with a hexagonal close packed
+86 335 858 7046; fax: +86 335 807 4545. (hcp) structure and higher microstructure stability at cryogenic
E-mail address: rdfu@ysu.edu.cn (R. Fu).

http://dx.doi.org/10.1016/j.jallcom.2014.08.201
0925-8388/ 2014 Elsevier B.V. All rights reserved.
514 C. Yuan et al. / Journal of Alloys and Compounds 619 (2015) 513519

temperature was chosen as a model metal because of the limited


deformation capability that results in easily saturated dislocation
density [17]. Some unusual changes in mechanical properties were
found in DCT-treated Zr, and the possible mechanisms in terms of
microstructural modication were discussed.

2. Experimental

The material used in this study was commercial pure Zr plate (chemical
composition summarized in Table 1). The plate was annealed in a vacuum at
1123 K for 4 h to obtain a homogeneous polycrystalline structure. The average grain
size of the as-annealed sample was approximately 3050 lm. The samples cut from
the as-annealed plate were placed in a liquid nitrogen environment at 77 K for 24 h.
Afterwards, the samples were taken out and cooled to room temperature.
X-ray diffraction (XRD) patterns of the as-annealed and DCT-treated samples
were measured using a Rigaku D/max 2500 X-ray diffractometer (18 kW) with Cu
Ka radiation (wavelength k = 1.54 ) in continuous-scanning mode over 2h = 30
80 at a step of 0.02. Differential scanning calorimetry (DSC) was conducted on a
Diamond DSC (PerkinElmer Inc., UK). For dynamic scans, samples were subjected
to temperatures ranging from 20 C to 160 C at a cooling rate of 20 C/min.
Microstructure features before and after DCT were observed by optical microscopy
(OM), electron backscatter diffraction (EBSD), and transmission electron micros-
copy (TEM). EBSD analysis was performed using a HITACHI S-4800 scanning elec-
tron microscopy (SEM) system, whereas TEM observations were carried out with
a JEOL-2010 TEM system at a voltage of 200 kV.
A FM-ARS9000 Vickers microhardness tester was used to measure the hardness
of samples. Indentations were carried out with a given load of 200 g and a dwell
time of 10 s. Hardness distributions over a square area of 4.5  4.5 mm2, with a
space between adjacent indentations of 0.5 mm were obtained. The tensile samples
were machined with a gauge length of 30 mm and a cross-section of 10  1.5 mm2.
Four repeated tensile tests were performed on a MTS test system at a strain rate of
1  104 s1 at room temperature under extensometer-measured strain control.
Fracture surfaces after the tensile test were observed by SEM, and microstructures
near the fracture surfaces were observed by TEM.

3. Results and discussion

3.1. XRD and DSC analyses

Fig. 1(a) shows the XRD patterns of as-annealed and DCT- Fig. 1. XRD patterns of as-annealed and DCT-treated samples (a), and the cryogenic
DSC curve of Zr (b).
treated samples. No phase transformation was observed in DCT-
treated Zr compared with as-annealed Zr. However, the intensities
of the diffraction peaks of DCT-treated Zr were slightly enhanced, fraction of HAGBs. No phase transformation and precipitation
which can be attributed to the variation in lattice constants caused occurred during DCT, consistent with the XRD and DSC results.
by DCT. Changes in lattice constants were also observed in DCT- Fig. 2(c) and (d) shows the IPF maps of Zr before and after DCT,
treated Mg [18,19]. To further conrm the microstructure stability respectively. The maximum intensity indicated the number of ran-
of Zr at cryogenic temperature, a cryogenic DSC curve was obtained dom orientations. A strong preference was observed toward the
and is shown in Fig. 1(b). The smooth feature of the DSC curve indi- (0 0 0 1) orientation, which was corroborated by the OIM maps in
cated that no transformation occurred during DCT. Fig. 2(a) and (b). Furthermore, the intensity of prism planes was
lower and the grain orientations were much closer to the (0 0 0 1)
3.2. EBSD analysis basal plane orientation after DCT. This nding can be attributed
to the ordered arrangement of atoms resulting from the small
To gain deep insight into microstructural changes, EBSD analy- changes in lattice constants [18,19]. The misorientation angle dis-
sis was performed. In this analysis, orientation imaging microscopy tributions in Fig. 2(e) and (f) shows that the crystallite boundaries
(OIM) maps, inverse pole gure (IPF) maps, and boundary misori- were mainly high angle in nature. The HAGB fractions of Zr before
entation angle distributions were obtained before and after DCT, and after DCT were about 81.2% and 85.3% of the total grain bound-
as shown in Fig. 2. Fig. 2(a) and (b) shows the OIM maps of Zr ary length, respectively. Thus, DCT improved the formation of
before and after DCT, with high-angle (HAGBs; grain boundary HAGBs. The increase in HAGB fractions may be related to the
misorientations P 15) and low-angle (LAGBs; grain boundary extrinsic dislocations formed during DCT (Fig. 3). The extrinsic dis-
misorientations < 15) grain boundaries depicted by black and locations were probably easier to react with the dislocations in
white lines, respectively. The crystallographic directions corre- LAGBs [20,21]. Therefore, the accumulation and rearrangement of
sponding to various colors can be inferred from the IPF triangle dislocations resulted in the transformation of LAGBs to HAGBs,
shown at the bottom right corner of Fig. 2(a) and (b). The micro- leading to increased HAGB and decreased LAGB fractions.
structure was characterized by equiaxed grains and a higher

3.3. TEM observations


Table 1
Chemical composition (in wt%) of the investigated pure Zr. The TEM images of the samples before and after DCT are pre-
Fe + Cr C N H O Hf Zr sented in Fig. 3. The dislocation density of the DCT-treated sample
was apparently higher than that of the as-annealed sample. When
0.2 0.05 0.01 0.005 0.16 4.5 Balance
the temperature dropped from room temperature to cryogenic
C. Yuan et al. / Journal of Alloys and Compounds 619 (2015) 513519 515

Fig. 2. Representative OIM maps of (a) as-annealed and (b) DCT-treated samples. IPF maps of (c) as-annealed and (d) DCT-treated samples. Distribution of boundary
misorientation angle of (e) as-annealed and (f) DCT-treated samples.

temperature, the volume of the material contracted. This volume substantially varied from approximately 140190 Hv, with an average
contraction released great compression deformation energy that value of 159.5 Hv. By contrast, the hardness of DCT-treated Zr gen-
served as the driving force for the formation and movement of dis- erally changed from approximately 160220 Hv, with an average
locations [18]. The improvement in dislocation density played a value of 197.5 Hv. Evidently, the average hardness signicantly
signicant role in enhancing mechanical properties and deforma- increased after DCT. The difference between the two average hard-
tion behavior, as discussed in the following sections. ness values reached 23.8%. Although the spatial distribution of
hardness was not uniform from one area to another, the increase
3.4. Hardness measurements in hardness of the DCT-treated sample was a holistic rather than
localized phenomenon, as can also be conrmed by the hardness
To better understand the variation in hardness of Zr, contour evolution characterized by the color maps in Fig. 4. The inhomoge-
maps of the spatial distribution of the Vickers hardness (Hv) over neity of hardness distribution may be closely related to the grain
a 4.5  4.5 mm2 square area before and after DCT are shown in orientation. An in situ comparison between the OIM map and
Fig. 4(a) and (b), respectively. The hardness of as-annealed Zr OM image of hardness indentation of DCT-treated Zr is presented
516 C. Yuan et al. / Journal of Alloys and Compounds 619 (2015) 513519

Fig. 3. TEM images of (a) as-annealed and (b) DCT-treated samples.

among different grain orientations reached up to 30.2%. The depen-


dency of hardness values on orientation has also been observed in
previous reports [2224].
For hcp Zr, the axial ratio c/a was 1.593, which was less than
that of the ideal value (i.e., <1.633). Consequently, the lattice resis-
tance for prism planes was lower than that for basal or pyramidal
planes; thus, slip preferred to occur on prism planes [25]. In other
words, plastic deformation more easily occurred on prism planes
than on basal planes. Consequently, the hardness in basal planes
was much larger than that in prism planes. Therefore, the overall
improvement of hardness may be closely related to both changes
in grain orientation and dislocation density. After DCT, the grain
orientation tended to be more consistent, resulting in ordered
strengthening [18]. Meanwhile, the increase in dislocation density
increased the force to overcome the resistance of dislocation
motion. Thus, the improvement in plastic deformation resistance
resulted in increased hardness.

3.5. Tensile properties

The tensile engineering and true stressstrain curves of


as-annealed and DCT-treated Zr are presented in Fig. 6. For
as-annealed Zr, the yield strength (0.2% offset, r0.2) and ultimate
tensile strength (rUTS) were 288 and 393 MPa, respectively. The
elongation to fracture reached 28.7%. Correspondingly, r0.2 and
rUTS of DCT-treated Zr increased by about 60% and 40%, respec-
tively, compared with as-annealed Zr. Meanwhile, the elongation
to fracture reached 26.5%, which was similar to that of as-annealed
Zr. This nding indicated that DCT was a better route to improving
the strength and keeping a good ductility of Zr.
As depicted in Fig. 6(b), DCT-treated Zr also exhibited high
strain hardening capabilities relative to as-annealed Zr. The strain
hardening exponent (n) is dened by [26]
r K en 1
where r is the true stress, e is the true strain, and K is a constant.
The n values for as-annealed and DCT-treated Zr were 0.09 and
0.11, respectively. The enhancement in strain hardening capabilities
Fig. 4. Spatial distributions of the hardness values of Zr: (a) before and (b) after of DCT-treated Zr was further conrmed by an improved strain
DCT. hardening rate (H) dened by [26]
 
1 @r
in Fig. 5. The letters in Fig. 5(a) correspond to those in Fig. 5(b), and H 2
grain M indicates the reference grain. The hardness values of grains
r @e
AD in Fig. 5(b) were 211.65, 259.54, 199.35, and 233.57 Hv, Variations in H with e before and after DCT are displayed in
respectively. The hardness values varied with the orientation Fig. 7. The variation trends were consistent with continuous strain
between basal and prism planes. Planes close to the (0 0 0 1) basal hardening to signicant strains, but the strain hardening rate of the
plane orientation were signicantly harder than planes close to DCT-treated sample was larger than that of the as-annealed
the prism 1 0 1  0 and 2 1 1
 0 planes. The hardness difference sample.
C. Yuan et al. / Journal of Alloys and Compounds 619 (2015) 513519 517

M (a) (b)
A
D
B C
B C
D

A M

Fig. 5. (a) OIM map and (b) corresponding OM image of hardness indentation of the DCT-treated Zr.

Fig. 7. Normalized strain hardening rate (H) against the true strain of as-annealed
and DCT-treated samples.

compatible than that of as-annealed Zr. Many dislocations were


also observed within the grains, which were divided into smaller
blocks (Fig. 8(b)). During DCT, a large number of pre-existing dislo-
cations were introduced into the grains (Fig. 3(b)), and these grains
were movable during tensile deformation [2732]. At the initial
stage of deformation, more dislocations and dislocation tangles
formed. The increase in dislocation density and resistance of
dislocation motion resulted in improved strength. With increased
strain, many irregular dislocation tangles changed into parallel dis-
location lamellas along with the tensile direction. With further
increased strain, dislocation accumulation and rearrangement
occurred to form dislocation cells, dense dislocation walls, and
nanoscale subgrains and grains. As a result, the coarse grains were
divided into smaller blocks. These blocks made the plastic
Fig. 6. Tensile (a) engineering and (b) true stressstrain curves of as-annealed and
DCT-treated samples.
deformation of DCT-treated Zr more uniform and compatible,
thereby resulting in the ductility being comparable to that of
as-annealed Zr.
DCT-treated Zr exhibited enhancement of strength and main-
tained the good ductility compared with as-annealed Zr. The vari- 3.6. Fracture behavior
ation in mechanical properties can be attributed to the difference
in dislocation density and dislocation congurations produced dur- The fracture surfaces of the as-annealed and DCT-treated Zr
ing deformation [17,27,28]. To reveal the mechanism underlying sample are shown in Fig. 9. As shown in Fig. 9(a), the fracture sur-
the increase in strength and maintain the good ductility in DCT- face of edge regions of the as-annealed sample exhibited mixed
treated Zr, TEM observations of the fractured tensile samples were characteristics with dimples and a small amount of quasi-cleavage
performed (Fig. 8). As shown in Fig. 8(a), for as-annealed Zr after facets, whereas the fracture surface in the middle region as shown
tensile deformation, the grains were apparently distorted and elon- in Fig. 9(b) mainly contained many quasi-cleavage facets. The
gated, and masses of dislocations were found in the grains. How- dimples in the edge of fracture surface of as-annealed sample
ever, the deformation of DCT-treated Zr was more uniform and were related to the noticeable necking caused by severe strain
518 C. Yuan et al. / Journal of Alloys and Compounds 619 (2015) 513519

Fig. 8. TEM images of (a) as-annealed and (b) DCT-treated samples after tensile tests.

(a) (b)

500 m 500 m

(c) (d)

500 m 500 m

Fig. 9. SEM images showing the fracture surfaces of as-annealed and DCT-treated Zr tensile samples. (a) and (b) are the edge and middle region of as-annealed Zr sample; (c)
and (d) are the edge and middle region of DCT-treated Zr sample.

localization during tensile test. For comparison, the fracture characteristics in the edge and middle region exhibited by as-
surfaces of DCT-treated Zr in the same regions are presented in annealed Zr. By contrast, the motion of pre-existing dislocations
Fig. 9(c) and (d), respectively. The fracture characteristics of two and specic dislocation congurations effectively enhanced the
regions were mainly consisted of quasi-cleavage facets. No obvious compatible deformation capability, resulting in the same fracture
necking was observed in the edge and middle regions of the characteristics in both regions of the DCT-treated sample.
fracture surfaces.
The above ndings indicated that DCT restricted strain localiza- 4. Conclusion
tion occurring in the edge regions of tensile samples. For as-
annealed Zr, plastic instability (i.e., necking) occurred in the edge Based on the above-described investigations, the following con-
region of tensile samples before ultimate fracture. Moreover, clusions were drawn.
Fig. 2(e) and (f) shows that the fraction of LAGBs before DCT was
slightly larger than that after DCT. LAGBs are reportedly associated (1) The grain orientations were much closer to the (0 0 0 1) basal
with the early onset of plastic instability during tensile testing plane orientation after DCT. DCT increased HAGB fractions
[3335]. The plastic instability accounted for the different fracture and dislocation density.
C. Yuan et al. / Journal of Alloys and Compounds 619 (2015) 513519 519

(2) Hardness was closely related to grain orientations and dislo- [8] R. Thornton, T. Slatter, A.H. Jones, R. Lewis, Wear 271 (2011) 23862395.
[9] T. Slatter, R. Lewis, A.H. Jones, Wear 271 (2011) 14811489.
cation density. The hardness in basal planes was much
[10] P.J. Singh, S.L. Mannan, T. Jayakumar, D.R.G. Achar, Eng. Fail. Anal. 12 (2005)
higher than that in prism planes. After DCT, dislocation den- 263271.
sity increased the resistance of dislocation motion, thereby [11] S. Zhirafar, A. Rezaeian, M. Pugha, J. Mater. Process. Technol. 186 (2007) 298
improving hardness. 303.
[12] V. Leskovek, B. Podgornik, Mater. Sci. Eng., A 531 (2012) 119129.
(3) Strength levels were high and good ductility could still be [13] H.H. Liu, J. Wang, B.L. Shen, H.S. Yang, S.J. Gao, S.J. Huang, Mater. Des. 28 (2007)
achieved after DCT. The enhancement in strength was due 10591064.
to the increase in dislocation density, and the maintenance [14] P.F. Stratton, Mater. Sci. Eng., A 449 (2007) 809812.
[15] A. Bensely, S. Venkatesh, D. Mohan Lal, G. Nagarajan, A. Rajadurai, Krzysztof
of good ductility resulted from the motion of pre-existing Junik, Mater. Sci. Eng., A 479 (2008) 229235.
dislocations and specic dislocation congurations. [16] X.Q. Jiang, N. Li, H. He, X.J. Zhang, C.C. Li, H. Yang, Mater. Sci. Forum 546549
(4) The fracture surfaces exhibited mixed features with quasi- (2007) 845848.
[17] D.F. Guo, M. Li, Y.D. Shi, T.Y. Ma, Z.B. Zhang, T.S. Wang, X.Y. Zhang, Mater. Sci.
cleavage facets and dimples for as-annealed Zr and quasi- Eng., A 581 (2013) 133139.
cleavage facets for DCT-treated Zr. The grains were divided [18] J.W. Liu, G.F. Li, D. Chen, Z.H. Chen, Chin. J. Aeronaut. 25 (2012) 931936.
into smaller blocks because of the interaction of dislocations [19] X.F. Zhang, G.H. Wu, W.C. Liu, W.J. Ding, Trans. Nonferrous Met. Soc. China 22
(2012) 28832890.
during the tensile process of DCT-treated Zr. Consequently, [20] P. Xue, B.L. Xiao, Z.Y. Ma, Mater. Sci. Eng., A 532 (2012) 106110.
the compatible deformation capability of materials was [21] Y.H. Zhao, J.F. Bingert, Y.T. Zhu, X.Z. Liao, R.Z. Valiev, Z. Horita, T.G. Langdon,
enhanced and good ductility could still be achieved. Y.Z. Zhou, E.J. Lavernia, Appl. Phys. Lett. 92 (2008) 081903.
[22] D.N. French, D.A. Thomas, Trans. AIME 233 (1965) 950952.
[23] T. Takahashi, E.J. Freise, Philos. Mag. 12 (1964) 18.
[24] B. Roebuck, P. Klose, K.P. Mingard, Acta Mater. 60 (2012) 61316143.
Acknowledgment [25] L. Zhang, Y. Han, J. Lu, Nanotechnology 19 (2008) 165706.
[26] G.E. Dieter, Mech. Metall., third ed., McGraw-Hill, London, 1988.
[27] D.F. Guo, M. Li, Y.D. Shi, Z.B. Zhang, T.Y. Ma, H.T. Zhang, X.Y. Zhang, Mater. Sci.
Financial support from the National Basic Research Program of Eng., A 558 (2012) 611615.
China (Grant No. 2010CB731606) is greatly acknowledged. [28] D.F. Guo, M. Li, Y.D. Shi, Z.B. Zhang, H.T. Zhang, X.M. Liu, X.Y. Zhang, Mater.
Lett. 66 (2012) 305307.
[29] X.X. Huang, Scr. Mater. 60 (2009) 10781082.
References [30] L. Lu, X. Chen, X. Huang, K. Lu, Science 323 (2009) 607610.
[31] X.X. Huang, N. Hansen, N. Tsuji, Science 312 (2006) 249251.
[1] R.F. Barron, Cryogenics 22 (1982) 409414. [32] Z.W. Shan, Raja K. Mishra, S.A. Syed Asif, Oden L. Warren, Andrew M. Minor,
[2] K.M. Asl, A. Tari, F. Khomamizadeh, Mater. Sci. Eng., A 523 (2009) 2731. Nat. Mater. 7 (2008) 115119.
[3] M. Koneshlou, K.M. Asl, F. Khomamizadeh, Cryogenics 51 (2011) 5561. [33] R. Kapoor, N. Kumar, R.S. Mishra, C.S. Huskamp, K.K. Sankaran, Mater. Sci. Eng.,
[4] A. Molinari, M. Pellizzari, S. Gialanella, G. Straffelini, K.H. Stiasny, J. Mater. A 527 (2010) 52465254.
Process. Technol. 118 (2001) 350355. [34] P.L. Sun, C.Y. Yu, P.W. Kao, C.P. Chang, Scr. Mater. 52 (2005) 265269.
[5] P. Baldissera, Mater. Des. 31 (2010) 47254730. [35] N. Kumar, R.S. Mishra, C.S. Huskamp, K.K. Sankaran, Mater. Sci. Eng., A 528
[6] P. Baldissera, C. Delprete, Open Mech. Eng. J. 2 (2008) 111. (2011) 58835887.
[7] B. Podgornik, F. Majdic, V. Leskovsek, J. Vizintin, Wear 288 (2012) 8893.

Você também pode gostar