Sei sulla pagina 1di 5

Materials Transactions, Vol. 51, No. 6 (2010) pp.

1067 to 1071
#2010 The Japan Institute of Metals

Evaluation of Dislocation Density in a Mg-Al-Mn-Ca Alloy


Determined by X-ray Diffractometry and Transmission Electron Microscopy
Takashi Shintani*1 , Yoshinori Murata*2 , Yoshihiro Terada and Masahiko Morinaga
Department of Materials, Physics and Energy Engineering, Graduate School of Engineering,
Nagoya University, Nagoya 464-8603, Japan

Metallic materials suffering deformation store elastic strain. Evaluation of this strain energy is important for understanding the mechanical
and physical properties of the materials. Although direct evaluation of the stored energy is difficult, it can be evaluated by determining the defect
energy of dislocations induced by the deformation. Thus, a practicable method of evaluating the strain energy is to measure the dislocation
density in metallic materials. The average and representative dislocation density can be estimated by X-ray diffraction (XRD) analysis. We have
estimated the dislocation density of a magnesium alloy with hexagonal crystals by a modified Warren–Averbach analysis based on a modified
Williamson–Hall plot using XRD profiles. The dislocation density value obtained by this method agrees with those reported previously. We
found that the modified Warren–Averbach method is still a powerful method for evaluating the dislocation density in hexagonal crystals.
[doi:10.2320/matertrans.M2010021]

(Received January 21, 2010; Accepted March 8, 2010; Published April 21, 2010)
Keywords: X-ray diffraction, transmission electron microscopy, magnesium alloys, dislocations

1. Introduction Table 1 Chemical composition of AX 52 die-cast alloy (mass%).

Al Ca Mn Mg
In the past decade, owing to its extremely light weight, 4.98 1.72 0.29 bal.
magnesium alloys have been used in the automotive
industry for improving fuel efficiency through vehicle mass
reduction. However, low temperature capability of the In recent years, XRD profile analysis has been developed
alloys has restricted their practical use. It is therefore to reveal microstructural details, such as dislocation density
desirable to develop magnesium alloys with high temperature and crystallite size. The measurement of microstructural
capability.1) details in hexagonal crystals is more difficult than in cubic
AM (Mg-Al-Mn) series alloys are commonly used in crystals because of the complexity of their crystal structures.
automobiles because of a good combination of their mecha- The multiple whole-profile (MWP) fitting procedure2) is the
nical properties, corrosion resistance and die-castability. principal analysis method for measuring microstructural
Calcium is a cost-effective, lighter alternative to rare-earth details in hexagonal crystals, such as Mg,3,4) Ti5,6) and Zr.7)
elements for improving the resistance to high temperatures However, the principle of the procedure is complicated, and
of AM series alloys.1) The creep strength of AM50 die-cast also the results depend strongly on the input data with
alloys has been improved by the additions of calcium. arbitrary properties. The purpose of this study is to apply
AM50 + xCa (x ¼ 0:47, 0.95 and 1.72 mass%) die-cast a simpler method proposed originally by Ungár et al.8) to
alloys are superior in high temperature capability to AM50, analyse microstructural details in Mg-based alloys. The
as reported by Itoh et al.1) Also, the value of the dislocation results are compared with those obtained by TEM and their
density of AX52 (AM50 + 1.72Ca) die-cast alloy has been validity is discussed.
measured with transmission electron microscopy (TEM)
techniques. The chemical composition of AX52 is shown in 2. Method of Analysing X-ray Diffraction Profiles
Table 1.1)
The dislocation density can be measured by direct 2.1 Calculation of b2 C
methods, such as TEM techniques, and by indirect methods, In hexagonal crystals, the average contrast factor of
such as X-ray diffraction (XRD) or neutron techniques. These dislocation, Chk:l corresponding to (hk.l) reflections can be
experimental techniques are complementary. Direct tech- written as9)
niques can reveal microstructural information over extremely
Chk:l ¼ Chk:0 ð1 þ q1 x þ q2 x2 Þ; ð1Þ
small areas of samples, whereas the latter two techniques can
reveal average and representative data. However, the sample where x ¼ ð2=3Þðl=gaÞ2 , and q1 and q2 are parameters that
preparation procedures for TEM studies are more difficult depend on the elastic properties of the material. Chk:0 is the
and complicated than those for XRD or neutron studies. average contrast factor corresponding to (hk.0) reflections.
Furthermore, it is known that the small thickness of TEM a and l are the lattice constant in the basal plane and the last
samples results in low dislocation density. index of the (hk.l) reflection, respectively. g is the diffraction
vector, which is a function of the fourth-order invariant of
*1Graduate Student, Nagoya University the (hk.l) indices and the lattice constants. Hereafter, Chk:l
*2Corresponding author, E-mail: murata@numse.nagoya-u.ac.jp is represented as C for simplicity.
1068 T. Shintani, Y. Murata, Y. Terada and M. Morinaga

Table 2 The most common slip systems in hexagonal crystals. for the same Burgers vector types. Substituting eq. (1) in
(a) Edge dislocations eq. (4), the following three equations are obtained:
Major slip
Sub-slip systems Burgers vector Slip plane
Burgers vector
1X 3
ðiÞ
systems types qðmÞ
1 ¼ hi Chk:0 b2i qðiÞ
1 ;
P i¼1
Basal BE h21 1 0i {0001} a
Prismatic PrE h2 110i f011 0g a 1X 3
ðiÞ
qðmÞ
2 ¼ hi Chk:0 b2i qðiÞ
2 ; ð5Þ
PrE2 h0001i f011 0g c P i¼1
PrE3 h2 113i f011 0g cþa X
3
Pyramidal PyE h1 21 0i f101 1g a hi ¼ 1;
i¼1
PyE2 h2 113i f21 1 2g cþa
PyE3 h2 113i f112 1g cþa where
PyE4 h2 113i f101 1g cþa X
3
ðiÞ ðmÞ
(b) Screw dislocations P¼ hi Chk:0 b2i ¼ b2 Chk:0 ; ð6Þ
i¼1
Burgers vector
Slip systems Burgers vector 0  hi  1: ð7Þ
types
S1 h21 1 0i a
q1 ðmÞ and q2 ðmÞ are the measured q1 and q2 , respectively,
S2 h2 113i cþa which are determined from the modified Williamson–Hall
S3 h0001i c plot mentioned in section 2.2. They can be obtained from the
ðiÞ
XRD profiles. Chk:0 , q1 ðiÞ and q2 ðiÞ are the numerically
calculated values for all sub-slip systems.
To evaluate the three unknowns, ha , hc and hcþa the values
ðmÞ ðiÞ
b2 C is the measured b2 C of the sample. b represents of the abovementioned parameters Chk:0 , q1 ðiÞ and q2 ðiÞ are
the magnitude of Burgers vector. Table 2 shows the most required for solving the simultaneous equations of eq. (5).
common slip systems in hexagonal crystals. There are three They were published previously for the most common sub-
types of fundamental Burgers vectors represented as a type, slip systems.9) Using the average values of ha , hc and hcþa ,
c type and c þ a type in hexagonal crystals. In the analysis, b2 C can be determined from eq. (6).
therefore, averaged value of the magnitude of Burgers vector
is used. b2 represents the averaged value of square b. 2.2 Calculation of the values of q1 ðmÞ and q2 ðmÞ by the
ðmÞ
b2 C can be written as9) modified Williamson–Hall plot
ðmÞ XN The value of the full width at half-maximum (FWHM)
ðiÞ
b2 C ¼ fi C b2i ; ð2Þ obtained from each peak is substituted into the following
i¼1 modified Williamson–Hall equation:10)
where N is the number of different sub-slip systems. In the  2 2 1=2
case of hexagonal crystals, N is equal to 11, as shown K ¼ 0:9 þ M b 1=2
1=2 KC þ OðK 2 CÞ; ð8Þ
ðiÞ
in Table 2.5) C and bi are the average contrast factor and D 2
the Burgers vector corresponding to the ith sub-slip system, where K ¼ 2sin= and K ¼ 2cosðÞ=. Here, , 
respectively. fi arePNthe fractions of the particular sub-slip and  are the FWHM, diffraction angle and wavelength of the
systems, where i¼1 fi ¼ 1 and fi  0. For a hexagonal X-rays, respectively. In the case of Cu,  ¼ 0:15405 nm. D, 
crystal structure, eq. (2) can be written using three types of and b are the average particle size, the dislocation density and
fundamental Burgers vectors, defined as b1 ¼ 1=3h2110i the magnitude of the Burgers vector, respectively. M is a
(a type), b2 ¼ 1=3h0001i (c type) and b3 ¼ 1=3h2113i constant depending on both the effective outer cut-off radius
(c þ a type), of dislocations and the dislocation density. O stands for
1=2
ðmÞ X
Nhai X
Nhci X
Nhcþai higher-order terms in KC . Equation (8) is converted into
b2 Chk:l ¼ b21 fi CðiÞ þ b22 fj C ð jÞ þ b23 fn C ðnÞ ; the quadratic form, in which its high-order term becomes
i¼1 j¼1 n¼1 negligible. Modifying eq. (1) into the quadratic form, the
ð3Þ following equation is obtained:
where Nhai, Nhci and Nhc þ ai are the number of sub-slip  Chk:0 ð1 þ q1 x þ q2 x2 Þ;
½ðKÞ2  =K 2 ¼ ð9Þ
systems with Burgers vector types a, c and c þ a, respec-
tively. The values of fi were determined by the relative where  ¼ ð0:9=DÞ2 and  ¼ M 2 b2 =2. The value of  is
fractions of the population of the Burgers vectors types, a, c determined while keeping the left-hand side of eq. (9) as a
and c þ a, on the basis of the number of sub-slip systems. quadratic function of x, as shown in Fig. 1(a). As a result,
As a result, eq. (3) can be written as5,9) the measured values of q1 ðmÞ and q2 ðmÞ are obtained from the
X3 coefficients of x in the quadratic function.
ðmÞ ðiÞ
b2 Chk:l ¼ hi C b2i ; ð4Þ
i¼1 2.3 Evaluation of the dislocation density by the modified
where hi are the fractions of the dislocation population in the Warren–Averbach analysis
sample with the same Burgers vector, ba , bc and bcþa . Now, The modified Warren–Averbach equation is written as11)
ðiÞ
C are the average contrast factors over the sub-slip systems
Evaluation of Dislocation Density in a Mg-Al-Mn-Ca Alloy Determined by X-ray Diffractometry and Transmission Electron Microscopy 1069

Fig. 2 Measured data and deconvolution lines using a Lorenz function of


the (21.1) reflection.

Fig. 1 Schematic diagrams showing the method of analysis: (a) ½ðKÞ2 


=K 2 versus x plot following eq. (9); (b) the modified Warren–Averbach Table 3 The diffraction angles and FWHM of the diffraction profiles used
plot following eq. (11); (c) Y=L2 versus ln L. in this study.

  Bragg reflections 2/deg FWHM/deg


b2 2 Re
ln AðLÞ 
¼ ln AS ðLÞ   L ln ðK 2 CÞ 00.2 34.47 3:97  102
2 L
 2 2     10.1 36.76 4:32  102
b R1 R2 2
þQ L4 ln ln ðK 4 C Þ: ð10Þ 11.0 57.71 6:22  102
2 L L
20.0 67.50 7:17  102
For hexagonal crystals, eq. (10) can be arranged as 20.1 70.20 7:46  102
  21.1 97.15 1:15  101
 2 Re
ln AðLÞ 
¼ ln A S
ðLÞ   L ln ðK 2 b2 CÞ
2 L
 2     down to 0.3 mm. Further, electropolishing was performed to
 R1 R2 2
þQ L4 ln ln ðK 4 b2 C Þ: ð11Þ remove the extra dislocations introduced into the sample
2 L L
surface by the earlier mechanical polishing.
Here, AðLÞ is the real part of the Fourier coefficients of For the XRD analysis, the diffraction profiles of the
the XRD profiles. Superscript S in the term ln AðLÞ indicates (00.2), (10.1), (11.0), (20.0), (20.1) and (21.1) reflections
the crystallite size, and L is the Fourier variable. L is defined were measured with a conventional diffractometer (Rigaku
as12) UltimaIV X-ray diffractometer), using Cu K1 and K2
radiation operating at 40 kV and 40 mA at a scan speed of
L ¼ na3 ; ð12Þ
0.25 min1 .
where a3 ¼ =f2ðsin2  sin1 Þg, n are integers starting from In this study, the modified Williamson–Hall as well as the
0 and (2  1 ) is the angular range of the measured modified Warren–Averbach plots were employed to obtain
diffraction profiles.12) Re is the effective outer cut-off radius the microstructural details involved in dislocation densi-
of dislocations. Q, R1 and R2 are all constants. ty.9,10,13) In these plots, the peaks obtained from only Cu K1
ln AðLÞ can be considered as a function of K 2 b2 C from radiation were needed. Therefore, each peak was separated
eq. (11), and hence the dislocation density is determined into K1 radiation and K2 radiation using a Lorenz function,
ðmÞ
from the ln AðLÞ versus K 2 b2 Chk:0 plot, as shown in and the FWHM of all peaks were obtained. A representative
Fig. 1(b), as the coefficient of the second term in eq. (11) Y, result of peak deconvolution is shown in Fig. 2.
which is arranged as
Y   4. Results and Discussion
2
¼  ln Re   ln L: ð13Þ
L 2 2
The XRD profile data obtained from AX52 alloy are
The value of  is evaluated from the gradient of the linear listed in Table 3. Using these data, analysis was performed
relationship between Y=L2 and ln L, as shown in Fig. 1(c). according to eqs. (8) and (9). As a result,  ¼ 4:29 
105 nm2 , q1 ðmÞ ¼ 0:266 and q2 ðmÞ ¼ 0:0184 are obtained
3. Experimental Procedures from the ½ðKÞ2  =K 2 versus x plot, as shown in Fig. 3.
Although the plot data is localised, a smooth quadratic curve
A sample of AX52 die-cast alloy was cut into the proper was obtained.
size and polished mechanically with emery papers down From the value of , the average particle size D can be
to #2000, followed by buff polishing with Al2 O3 powders determined as: D ¼ 137 nm. The value of D is apparent grain
1070 T. Shintani, Y. Murata, Y. Terada and M. Morinaga

Fig. 3 The relationship between ½ðKÞ2  =K 2 and x in eq. (9). Fig. 5 Y=L2 versus ln L plot according to eq. (13). The gradient of the
regression line provides the dislocation density. The effective outer cut-off
radius of dislocations is also obtained from the intercept value of the
horizontal scale.

Table 4 The comparison between the result of this study and others.

Dislocation
Specimen Method of analysis density
=1013 m2
This modified Warren-Averbach
AX52 die-cast alloy 5.3
study plot
Itoh’s
AX52 die-cast alloy TEM observation 2.0
study
AZ91 alloy
Mathis’s
(heat treated 18 h MWP fitting procedure 4.0
study
at 686 K)

mentioned in section 1, the dislocation-free zone affects the


Fig. 4 The modified Warren–Averbach plot following eq. (11).
number of dislocations and results in low density. Further-
more, in TEM contrasts, it is not necessarily the case that all
size measured from the modified Williamson–Hall plot. This dislocations can be measured, because of the extinction law
value reflects the grain size but is not always identical to the in TEM observations. The true value of  in the sample
real size. Therefore, it should be considered to be a measure must be larger than that reported by Itoh et al. With this
of the grain size.14) interpretation, our result is consistent with the earlier results
The values of q1 ðmÞ and q2 ðmÞ are substituted into eq. (5), obtained by TEM observation and is considered to be
and then the fractions hi of the three Burgers vector types are practicable.
determined: ha ¼ 0:84, hc ¼ 0:10 and hcþa ¼ 0:06. Using hi , Máthis et al. calculated the dislocation density of AZ91
the modified Warren–Averbach plot was prepared according alloys heat treated for 18 h at 686 K4) by XRD analysis. Their
to eq. (11). The result is shown in Fig. 4. A series of reported value of  is very close to that in this study. They
quadratic curves was obtained using several L values. As a used the MWP fitting procedure, which is the principal
result,  ¼ 5:3  1013 m2 and Re ¼ 152 nm were estimated, analysis method for obtaining the value of the dislocation
as shown in Fig. 5. density from XRD profiles.2) Although the alloy used in this
To discuss the validity of the results obtained in this study study, AX52 is different in chemical composition from
and the method employed, values reported previously on the AZ91, the reported value is in good agreement with our
dislocation density of magnesium alloys1,4) are listed in result. Therefore, we conclude that the modified Warren–
Table 4, together with the value obtained in this study. Averbach analysis is effective because of its simplicity and
Itoh et al. reported the dislocation density of AX52 die- no arbitrary input data compared to the MWP fitting method.
cast alloy,1) which is the same alloy used in this study.
They measured the density with TEM techniques, and their 5. Conclusion
reported value of  is smaller than that of this study.
In general, it is known that some dislocations leave the The modified Warren–Averbach analysis method was
specimen surface, and a dislocation-free zone is formed in a revised and applied to the estimation of the dislocation
very thin layer at the surface. The thickness of conventional density value in AX52 (Mg-Al-Mn-Ca) die-cast alloy from
TEM samples is roughly several tens of nm; hence, as XRD profiles, and the validity of this analysis method was
Evaluation of Dislocation Density in a Mg-Al-Mn-Ca Alloy Determined by X-ray Diffractometry and Transmission Electron Microscopy 1071

discussed. We found that this method is credible and Lukáč: Acta Mater. 52 (2004) 2889–2894.
4) K. Máthis, J. Gubicza and N. H. Nam: J. Alloy. Compd. 394 (2005)
accurate. A similar procedure using AX52 (Mg-Al-Mn-Ca)
194–199.
die-cast alloy in this study is also applicable to other 5) I. C. Dragomir, D. S. Li, G. A. Castello-Branco, H. Garmestani, R. L.
hexagonal crystals, such as Ti, Zr and Zn, as well as MWP Snyder, G. Ribárik and T. Ungár: Mater. Charact. 55 (2005) 66–74.
fitting. 6) T. Ungár, M. G. Glavicic, L. Balogh, K. Nyilas, A. A. Salem, G.
Ribárik and S. L. Semiatin: Mater. Sci. Eng. A 493 (2008) 79–85.
Acknowledgement 7) T. Ungár, O. Castelnau, G. Ribárik, M. Drakopoulos, J. L. Béchade, T.
Chauveau, A. Snigirev, I. Snigireva, C. Schroer and B. Bacroix: Acta
Mater. 55 (2007) 1117–1127.
This work was supported by Grant-in-Aid for Scientific 8) T. Ungár, J. Gubicza, G. Ribárik and A. Brobély: J. Appl. Cryst. 34
Research of Japan Society for the Promotion of Science (2001) 298–310.
(JSPS), Japan. 9) I. C. Dragomir and T. Ungár: J. Appl. Cryst. 35 (2002) 556–564.
10) T. Ungár and G. Tichy: Phys. Status. Solid. 171 (1999) 425–434.
11) T. Ungár and A. Borbély: Appl. Phys. Lett. 69 (1996) 3173–3175.
REFERENCES 12) T. Kunieda, M. Nakai, Y. Murata, T. Koyama and M. Morinaga: ISIJ
Int. 45 (2005) 1909–1914.
1) D. Itoh, Y. Terada and T. Sato: Mater. Trans. 49 (2008) 1957–1962. 13) T. Ungár, J. Gubicza, G. Ribárik and A. Borbély: J. Appl. Cryst. 34
2) G. Ribárik, T. Ungár and J. Gubicza: J. Appl. Cryst. 34 (2001) 669– (2001) 298–310.
676. 14) Y. Murata, I. Nakaya and M. Morinaga: Mater. Transact. 1 (2008)
3) K. Máthis, K. Nyilas, A. Axt, I. Dragomir-Cernatescu, T. Ungár and P. 20–23.

Potrebbero piacerti anche