Вы находитесь на странице: 1из 7

ARTICLES

PUBLISHED ONLINE: 20 JUNE 2016 | DOI: 10.1038/NMAT4677

Polysynthetic twinned TiAl single crystals for


high-temperature applications
Guang Chen1*, Yingbo Peng1, Gong Zheng1, Zhixiang Qi1, Minzhi Wang1, Huichen Yu2,
Chengli Dong2 and C. T. Liu3*
TiAl alloys are lightweight, show decent corrosion resistance and have good mechanical properties at elevated temperatures,
making them appealing for high-temperature applications. However, polysynthetic twinned TiAl single crystals fabricated
by crystal-seeding methods face substantial challenges, and their service temperatures cannot be raised further. Here we
report that Ti45Al8Nb single crystals with controlled lamellar orientations can be fabricated by directional solidification
without the use of complex seeding methods. Samples with 0 lamellar orientation exhibit an average room temperature
tensile ductility of 6.9% and a yield strength of 708 MPa, with a failure strength of 978 MPa due to the formation of extensive
nanotwins during plastic deformation. At 900 C yield strength remains high at 637 MPa, with 8.1% ductility and superior creep
resistance. Thus, this TiAl single-crystal alloy could provide expanded opportunities for higher-temperature applications, such
as in aeronautics and aerospace.

he development of high-temperature materials is mainly


driven by the strong need for applications in gas turbine
engines. TiAl intermetallic alloy in polycrystalline forms with
a high specific strength has been successfully applied to replace
Ni-based superalloys in the temperature range 650750 C, with
the benefit of a weight reduction of as much as 50% (refs 15).
As reported by Bewlay et al. in 2013, more than 40,000 TiAl lowpressure turbine (LPT) blades have been manufactured for the
GEnx 1B (Boeing 787) and the GEnx 2B (Boeing 747-8) turbineengine applications6 . To improve the performance and service
temperature of gas-turbine engines, the most effective way known
is to use single crystals with well-aligned crystal orientations7,8 .
To acquire TiAl polysynthetic twinned (PST) single crystals
with parallel lamellar orientations (the 0 lamellar orientation),
the seeded directional solidification (DS) technique has been
found to be necessary for manufacturing9,10 . However, there exist
insurmountable obstacles in both acquiring and applying these
PST single crystals produced by the seeding methods. The wellaligned PST single crystals so fabricated at the present time cannot
reach the expected elevated-temperature mechanical properties.
Other drawbacks include the limited compositions available
and complicated processing with high manufacturing costs11,12 .
Furthermore, a lack of sufficient ductility at ambient temperature is
another major concern for structural applications of these crystals1 .
The preferred orientation of directionally solidified -solidifying
TiAl alloys ( is the primary phase on solidification) is
perpendicular to the PST lamellar orientations, so the seeded
DS method is the only way to acquire well-aligned PST single
crystals. However, for -solidifying PST single crystals ( is the
primary phase on solidification), there is no possible way to control
the lamellar orientations due to the uncertain orientations of the
phase generated from the solid phase transformation11,13 .
In this study, we have successfully controlled the
transformationand thus well-aligned lamellar

orientationthrough regulation of the anisotropic interfacial


energies. Hence, the high performance of Ti45Al8Nb (at.%)
alloy with well-aligned 0 -orientation PST single crystals has been
successfully achieved without the use of the high-cost seeding
method; the crystals exhibit a yield strength as high as 637 MPa and
an excellent creep resistance at 900 C. These findings are expected
to provide a strong technical base for the potential applications of
these TiAl crystals as rotating parts in the aviation and aerospace
fields, as well as turbocharger rotors in the automotive industry.
The solid phase transformations in -solidifying TiAl alloys are
-Ti (body-centred cubic) to -Ti (hexagonal close-packed), and
then to 2 -Ti3 Al (hexagonal close-packed) and -TiAl (face-centred
cubic) phases. When the body-centred cubic phase solidifies
from the liquid during DS, the preferential growth direction is the
h001i direction. Based on the Burgers and Blackburn orientation
relationship14,15 , the 12 variants generated from the
solid phase transformation in TiAl alloys directly lead to crystal
orientation probabilities of 1/3 and 2/3 for obtaining 0 and
45 lamellar orientations, respectively16,17 . By carefully analysing
the transformation process, we realize that the different
variants growing along the 0 or 45 directions have the same
crystal parameters, except at the / phase interface. For the
0 -growing phase, the crystal planes on the sides of the interface
of the phase and (1120)
of the phase. For the
are (001)
45 -growing phase, the crystal planes on the sides of the interface
of the phase and (1012)
of the phase. So the /
are (001)
k (1120)
for 0 -growing phase and (001)
k
interface is (001)

(1012) for 45 -growing case. To analyse the difference between these


two types of the / phase interface, we use the Bramfitt planar
disregistry equation18 :
|d
(hkl)s
(hkl)n
=

3
X
i=1

[uvw]is

cos d

|
[uvw]in

[uvw]in

100

1 Engineering

Research Center of Materials Behavior and Design, Ministry of Education, Nanjing University of Science and Technology, Nanjing 210094,
China. 2 Aviation Key Laboratory of Science and Technology on Materials Testing and Evaluation, Science and Technology on Advanced High Temperature
Structural Materials Laboratory, Beijing Key Laboratory of Aeronautical Materials Testing and Evaluation, Beijing Institute of Aeronautical Materials, Beijing
100095, China. 3 Centre for Advanced Structural Materials, Department of Mechanical and Biomedical Engineering, CSE, City University of Hong Kong,
Hong Kong, China. These authors contributed equally to this work. *e-mail: gchen@njust.edu.cn; chainliu@cityu.edu.hk
NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials

2016 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT4677

ARTICLES
a

Growth direction

a
c

2 mm

b
200 nm

200 nm

2 mm

A
B
C
D

100 m

100 m

Figure 1 | Optical micrographs of directionally solidified Ti45Al8Nb PST


single crystals at different withdrawal rates. a, Longitudinal section of an
ingot at a withdrawal rate lower than Vc , showing a PST single crystal with a
lamellar orientation aligned parallel to the growth direction. b, Longitudinal
section of an ingot at a withdrawal rate higher than Vc , showing a PST single
crystal with a lamellar orientation aligned at 45 to the growth direction.
c, Magnified lamellar microstructure in the area marked in a, showing more
clearly the parallel lamellar microstructure. d, Magnified area marked in b.

Based on this equation and the crystallographic parameters of the


TiAl alloy (see Supplementary Table 1), the calculated values in
k (1120)
are
the case of (001)

(001)

(1120)

|d[100] cos d[1100]


|

d[1100]

|d[010] cos d[0001] |


d[0001]

|d[110] cos d[1101]


|

d[1101]

100

k (1012)
are
and in the case of (001)

(001)

(1012)

|d[100] cos d[21 10]


|

d[21 10]

|d[010] cos d[0111]


|

d[0111]

|d[110] cos d[2423]


|

d[2423]

100

k
The calculated planar disregistry values of the 0 -growing (001)

(1120) interface and the 45 -growing (001) k (1012) interface


are 32.4% and 35.6%, respectively. The planar disregistry of
the 0 -directed interface is 3.2% lower than that of the 45 directed interface, which means the 0 -directed case has a lower
interfacial energy than the 45 -directed case (see details in the
Supplementary Information). Anisotropy of the interfacial energy
plays an important role in nucleation and growth during the
solid phase transformation. According to solid phase
transformation theory19 , the / interfacial energy is a resistance
to nucleation but a driving force for grain growth. With a different
interfacial energy, the 0 grains nucleate more easily than the
45 grains. So, if the nucleation driving force is sufficiently large
for the 0 grains to nucleate, there will be no 45 grains nucleating
in the matrix, leading to the parallel lamellar structures. However,
if the nucleation driving force is large enough for both 0 and
45 grains, the 45 grains grow more easily and faster than the
0 grains, resulting in the 45 -oriented lamellar structure.
The nucleation driving force is closely related to the undercooling
effect during transformation, while the degree of undercooling
2

10 nm

Figure 2 | Lamellar microstructure of a Ti45Al8Nb well-aligned PST


single crystal before and after tensile test at ambient temperature.
a, Bright-field TEM image of a tensile specimen showing the original 2 /
lamellar structure before the test. b, After the tensile deformation, the
bright-field TEM image reveals the ultrafine twinning and lamellar
structure, with the selected area electron diffraction pattern (inset).
c, A high-resolution (HR) TEM image of the deformed specimen showing
multiple-twinned structures containing three twin boundaries A /B , B /C
and C /D .

increases with the increased cooling rate. Because the cooling rate
(C) is calculated as the product of the thermal gradient (G) and
the withdrawal rate (V ), C = G V during the DS process20 , the
nucleation of the grains can be controlled by regulating V while G
remains constant. Based on these analyses, there must exist a critical
withdrawal rate Vc , below which the 0 grains can nucleate but
the 45 ones cannot, thus demonstrating that well-aligned TiAl PST
single crystals can be obtained without the use of seeding crystals.
Based on our theoretical analysis, unidirectional solidification
experiments are conducted on a Bridgman-type apparatus (see
Supplementary Fig. 3 and Methods) for the Ti45Al8Nb alloy and
the results are shown in Fig. 1. The 0 -oriented PST single crystals
can be obtained at a withdrawal rate lower than Vc (Fig. 1a,c),
while the 45 -oriented PST single crystals can be obtained at a
withdrawal rate higher than Vc (Fig. 1b,d). These results verify our
prediction well.
Tensile tests were performed at ambient temperature on our
non-seeded Ti45Al8Nb PST single crystals with controlled
orientations. The tensile ductility and yield strength of the wellaligned (0 ) alloys reach as high as 7.6% and 735 MPa, respectively
(see full data in Supplementary Table 2). In previous studies, the
ambient-temperature ductility of TiAl alloys with polycrystalline
structures is usually less than 2% (ref. 21), except for Appel et al., who
obtained tensile elongations of 3% in a novel laminate structure
in Ti(40-44)Al8.5Nb alloys22 . But our well-aligned PST single
crystals have a superior tensile ductility of 6.37.6%, indicating that
the microstructure and alloy composition both play an important
role in improving the ambient-temperature ductility. There exist
three kinds of dislocation in high Nb-TiAl crystals during

deformation at ambient temperature: the 1/2[110]


and 1/2[112]

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials

2016 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT4677


a

Service temperature

Our ST
150

0 yield strength
0 elongation
45 yield strength
45 elongation

800

125
100
75

600

50

Twins
Elongation (%)

Yield strength (MPa)

1,000

ARTICLES

Dislocations

400
25
500 nm

200
0

200

400

600

800

1,000

Temperature (C)

8,000
C

6,000

600
4,000
300

True stressstrain curve


2,000

0
0.00

0.02

0.04

0.06

800

15,000
10,000

600
A

5,000

400
200
0
0.00

0.08

Work hardening rate

B
True stressstrain curve

0
C

Work hardening rate (MPa)

900

c
Work hardening rate

True stress (MPa)

1,200

Work hardening rate (MPa)

True stress (MPa)

5,000
0.02

0.04

True strain

0.06

0.08

0.10

0.12

True strain

Figure 3 | Mechanical properties of Ti45Al8Nb PST single crystals as a function of temperature and the microstructure after tension. a, Mechanical
properties as a function of temperature. The well-aligned PST single crystal maintains a high yield strength of 637 MPa at 900 C; a temperature much
higher than the 650750 C reported for polycrystalline alloys4 (see the pink-colour region in the figure). b,c, The true stressstrain curve and the work
hardening rate obtained at ambient temperature (b) and 900 C (c). d, TEM microstructure of well-aligned PST single crystals after tension tested at
900 C. Twins and dislocations appear simultaneously after the elevated-temperature deformation.

superdislocation23 . Usually, the


screw dislocations as well the h011]
superdislocation is able to decompose into two superpartials
h011]
accompanied by twinning deformation. The
of type 1/2[011],
dissociation of the superdislocations has been proposed as the
possible reason for the enhanced ductility24 . Furthermore, splitting
of the superdislocations and twinning deformation have been
detected previously in polycrystalline TiAl + Nb alloys by Chen
and colleagues25 .
The lamellar structures in the as-cast crystals were examined
carefully by transmission electron microscopy (TEM), and the
coarse structural features before deformation are shown in Fig. 2a.
Refined nanotwinned (nt) structures with a twin thickness (the
spacing between adjacent twin boundaries) of 10 nm (Fig. 2b,c)
are formed essentially during plastic deformation at ambient
temperature. A comparison of the structural features before
and after the deformation clearly indicates that the deformation
twinning refines the twin spaces. A similar nt-structure has been
shown to have huge advantages in copper, as reported by Lu and
Tu (refs 2630). The nt-Cu can enhance the strength without the
loss of ductility, and it can also maintain an electrical conductivity
comparable to that of pure copper2628 . It also exhibits a higher
electro-migration resistance than regular Cu grains, and greatly
reduces the formation of Kirkendall voids29,30 . For all these beneficial
reasons, nt-structures have drawn a great deal of the recent attention
for alloy design. The initial state of twin nucleation is the passage
partial dislocation loop, which involves an
of the first 1/6h112]
intrinsic stacking fault. This increases the surface energy; thus, a
large part of the twin formation energy comes from the creation
of the twin boundary energy. Succeeding loops for the twin growth

form significantly more easily24 . Moreover, the reduction of stacking


fault energy plays a key role in the generation of the nt-structure26 .
Since Nb is known to reduce the stacking fault energy and enhance
the mobility of superdislocations, all of which make the dislocations
sweep more easily to form twins during deformation, the high
Nb content in TiAl crystals generally promotes the formation of
high-density nanoscale twins31 . The nt-structure has been shown
to be effective in improving the ductility and strength by blocking
dislocation motion. It is found that both the ductility and strength
increase with decreasing twin thickness26 . Lu et al. reported that the
yield strength follows an empirical HallPetch relationship with the
twin thickness27,28 , while Liu et al. observed a similar relationship
with the lamellar spacing at an even early time32 . Also, it has been
reported that alloying of Mn additions is effective in promoting
the formation of nanotwins and the enhancement of ductility in
TiAl alloys33,34 .
The temperature dependences of the mechanical properties
of Ti45Al8Nb PST single crystals with different lamellar
orientations are shown in Fig. 3a. Surprisingly, we discover that there
is no sharp decrease in the yield strength of our well-aligned crystals
from ambient temperature to 900 C (see full data in Supplementary
Table 3). The single-crystal material maintains a yield strength
of 637 MPa at 900 C, indicating that these crystals exhibit an
unprecedented high-temperature performance. Note that the yield
strength of our crystals at 1,000 C decreases drastically to 238 MPa,
while the ductility rises sharply from 8.1% (900 C) to 76.3%. The
brittle-to-ductile transition temperature (BDTT) of typical TiAl
alloys is usually in the range of 650820 C (ref. 21), indicating that
their service temperatures are relatively much lower. It is worthwhile

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials

2016 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT4677

ARTICLES

0.30
0.30
0.25

0.25

Creep strain

0.15

150 MPa

210 MPa

0.20
0.15

0.20

150 MPa

0.10
0.05

210 MPa
100 MPa

0.00

0.10
100 MPa
0.05
0.00

200

400

600

800

1,000

Time (h)

b
105
Minimum creep rates (s1)

mentioning that Heilmaier predicted the service temperature of


TiAl alloys can be extended by approximately 200 C by optimizing
the composition and microstructure35 . Based on our current work,
it is important to point out that the service temperature of our
well-aligned PST single crystals can be potentially raised to as high
as 900 C, which is substantially higher than that of the alloys
and crystals developed previously. The high performance at both
ambient and elevated temperatures of our crystals is consistent with
the prediction by Heilmaier and colleagues.
To understand the ductility effect of our Ti45Al8Nb alloys, we
have carefully assessed the work hardening rate (d/d) of wellaligned PST single crystals from the true stressstrain curves obtained at ambient temperature and 900 C (see Fig. 3b,c). The work
hardening rate shows three distinct stages: stage A, characterized
by a decreasing work hardening rate with the true strain; followed
by stage B, with an increasing work hardening rate; and finally
stage C, with a decreasing work hardening rate again. It has been
recognized that when deformation is dominated by dislocation slips,
the true stressstrain curve will be characterized by a higher yield
point, followed by a continuously decreased strain hardening rate.
However, as indicated in Fig. 3b, the work hardening rate can be
distinctly divided into three regions A, B and C. Region A, with a
continuous decrease in the work hardening rate with the true strain,
corresponds to dislocation-slip-dominated plastic deformation. The
increase in the hardening rate with strain in region B is due to
twinning-controlled plastic deformation. In region C, dislocationslip-dominated plastic deformation operates again36 . On the basis
of the work hardening analysis, we conclude that both dislocation
slip and twinning deformation take place during plastic deformation
at ambient temperature, and the twinning-controlled deformation
operates in the middle stage of the deformation, resulting in a high
tensile ductility at room temperature.
Figure 3c also indicates the operation of the twinning-controlled
deformation in our single crystal at 900 C after about 8% plastic
deformation dominated by dislocation slip. Note that region B in
Fig. 3c is not as prominent as the equivalent region in Fig. 3b,
and the reason for this difference is that the dislocations largely
activated at elevated temperatures eventually weaken the effect
of twins during plastic deformation. Nevertheless, twins and
dislocations still appear simultaneously at the elevated-temperature
deformation, as shown in Fig. 3d, and this result is consistent with
other work37 .
Creep resistance and structural stability, which are important
prerequisites for high-temperature applications of TiAl alloys,
essentially determine the service temperature range in competition
with other structural materials38 . Uniaxial tensile creep tests of
Ti45Al8Nb well-aligned PST single crystals with the 0 lamellar
orientation were performed under constant stresses of 100, 150 and
210 MPa, respectively, at 900 C in air. Since the Ti48Al2Cr2Nb
(4822) polycrystalline alloy is currently used in GEnx engine, we
have also tested 4822 polycrystalline alloys with fully lamellar
structures at the same conditions for comparison. The creep lifetime
tests of our single crystals have achieved a rupture time of 363 h
and 116 h, which are higher by a factor of 6668 than those of the
4822 alloys under 150 and 210 MPa, respectively. In addition, our
crystal still maintains steady-state creep up to 800 h, while the creep
lifetime of the 4822 alloy is only 76.4 h under the same stress of
100 MPa. All the creep data from our well-aligned PST TiAl + Nb
single crystals (with the 0 lamellar orientation) obtained at 900 C
under different stresses are shown in Fig. 4a. In comparison, the
creep curves of the 4822 polycrystalline specimens are also included
in Fig. 4a,b at the same testing conditions at 900 C. The minimum
creep rates of our single crystals during the steady-state creep are
measured to be 1.03 108 s1 , 5.79 108 s1 and 1.74 107 s1
under stresses of 100, 150 and 210 MPa, respectively. However,
the minimum creep rates of 4822 specimens are measured to be

Ti45Al8Nb, 0
Ti48Al2Cr2Nb, FL
n = 6.7

106

107
n = 3.8

108
100

120

140

160

180

200

220

Stress (MPa)

Figure 4 | Creep properties of well-aligned Ti45Al8Nb PST single


crystals with the 0 lamellar orientation and the commercial
Ti48Al2Cr2Nb polycrystalline alloy at different stresses at 900 C in
air. a, Creep strainlifetime curves. The red lines are for the Ti45Al8Nb
PST single crystals and the blue lines for the 4822 alloys. The inset is the
creep strainlifetime curves of the 4822 commercial alloy under stresses of
150 and 210 MPa. b, Comparison of the minimum creep rates of the single
and polycrystalline materials.

1.00 107 s1 , 2.01 106 s1 and 1.46 105 s1 under the same
stress range. It can be seen in Fig. 4b that the minimum creep
rate of our single crystal with the 0 lamellar orientation is lowered
dramatically by one to two orders of magnitude, as compared with
the results from the 4822 polycrystalline alloy at 900 C (see full
creep data in Supplementary Table 4). Furthermore, we calculated
the stress exponent n of our well-aligned PST single crystals and the
4822 specimens at 900 C by means of the creep rate equation:
= A n
The linear relationship in Fig. 4b demonstrates that the creep
behaviour of both materials can be described well by the above rate
equation. The n value is measured to be 3.8 for our single crystals
and n of 6.7 for the commercial 4822 alloys. The distinctly higher
value for the 4822 alloys clearly indicates that the polycrystalline
intermetallic alloy is much more sensitive to the load conditions
at 900 C.
Since the high-temperature creep performance of our TiAl + Nb
PST single crystals has been shown to be highly promising, another
factor required for consideration in industrial applications of TiAl
alloys is the fatigue resistance. Fortunately, many studies have

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials

2016 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT4677


demonstrated that TiAl alloys exhibit excellent fatigue resistance at
ambient and elevated temperatures6,21,3848 (see detailed discussion
of the TiAl fatigue properties in Supplementary Information). Our
PST single crystals have a (2 + ) lamellar microstructure with
a good combination of the strength and ductility at both room
and elevated temperatures; thus it is reasonably to judge that these
PST single crystals probably have an even better fatigue-resistance
performance. Nevertheless, an evaluation of the fatigue behaviour
of PST single crystals is highly recommended as a future study.
In conclusion, our study of directionally solidified PST single
crystals of TiAl + Nb with different lamellar orientations reveals
that it is the anisotropic interfacial energy that controls the final
lamellar orientations of these intermetallic crystals. Based on
our analyses, we have successfully fabricated PST single crystals
with controlled lamellar orientations at a low production cost,
without the requirement of complicated and costly crystal-seeding
methods. The well-aligned PST single crystals with the 0 lamellar
orientation exhibit a combination of a substantially enhanced
ductility of 6.37.6% and a high strength of 9301,035 MPa at
ambient temperature. The impressive mechanical performance of
the present TiAl PST single crystals can be attributed to the
formation of high-density nanoscale twins and the refined lamellar
structure during plastic deformation. More importantly, the crystals
possess a superior yield strength of 637 MPa and a high creep
resistance at 900 C in the stress range 100210 MPa in air. The creep
lifetime and minimum creep rates of our PST single crystals are
superior to those of commercial 4822 TiAl polycrystalline material
by more than one order of magnitude. Consequently, the service
temperature of this crystal material can be potentially raised to
as high as 900 C. Our studies have demonstrated an important
advance in the design and fabrication of TiAl + Nb PST single
crystals with controlled orientations, and the promising mechanical
performance of these alloys indicates that they have significant
potential in service environments at higher temperatures.

Methods
Methods and any associated references are available in the online
version of the paper.
Received 9 July 2015; accepted 24 May 2016;
published online 20 June 2016

References
1. Liu, C. T. & Stiegler, J. O. Ductile ordered intermetallic alloys. Science 226,
636642 (1984).
2. Kear, B. H. & Thompson, E. R. Aircraft gas turbine materials and processes.
Science 208, 847856 (1980).
3. Clemens, H. & Kestler, H. Processing and applications of intermetallic
-TiAl-based alloys. Adv. Eng. Mater. 2, 551570 (2000).
4. Lasalmonie, A. Intermetallics: why is it so difficult to introduce them in gas
turbine engines? Intermetallics 14, 11231129 (2006).
5. Mukherji, D. et al. The effects of boron addition on the microstructure and
mechanical properties of CoRe-based high-temperature alloys. Scr. Mater. 66,
6063 (2012).
6. Bewlay, B. P., Weimer, M., Kelly, T., Suzuki, A. & Subramanian, P. R. The
science, technology, and implementation of TiAl alloys in commercial aircraft
engines. MRS Proc. 1516, 4958 (2013).
7. Taub, A. I. & Fleischer, R. L. Intermetallic compounds for high-temperature
structural use. Science 243, 616621 (1989).
8. Perepezko, J. H. The hotter the engine, the better. Science 326,
10681069 (2009).
9. Johnson, D. R., Inui, H. & Yamaguchi, M. Directional solidification and
microstructural control of the TiAl/Ti3 Al lamellar microstructure in TiAlSi
alloys. Acta Mater. 44, 25232535 (1996).
10. Takeyama, M. et al. Lamellar boundary alignment of DS-processed TiAlW
alloys by a solidification procedure. Mater. Sci. Eng. A. 329331, 712 (2002).
11. Johnson, D. R., Masuda, Y., Inui, H. & Yamaguchi, M. Alignment of the
TiAl/Ti3 Al lamellar microstructure in TiAl alloys by growth from a seed
material. Acta Mater. 45, 25232533 (1997).

ARTICLES
12. Yokoshima, S. & Yamaguchi, M. Fracture behavior and toughness of PST
crystals of TiAl. Acta Mater. 44, 873883 (1996).
13. Yamaguchi, M., Johnson, D. R., Lee, H. N. & Inui, H. Directional solidification
of TiAl-base alloys. Intermetallics 8, 511517 (2000).
14. Burgers, W. G. On the process of transition of the cubic-body-centered
modification into the hexagonal-close-packed modification of zirconium.
Physica 1, 561586 (1934).
15. Blackburn, M. J. The Science, Technology, and Application of Titanium
(Pergamon, 1970).
16. Kim, M. C. et al. Composition and growth rate effects in directionally solidified
TiAl alloys. Mater. Sci. Eng. A. 239240, 570576 (1997).
17. Jung, I. S., Jang, H. S., Oh, M. H., Lee, J. H. & Wee, D. M. Microstructure
control of TiAl alloys containing stabilizers by directional solidification.
Mater. Sci. Eng. A. 329331, 1318 (2002).
18. Bramfitt, B. L. The effect of carbide and nitride additions on the heterogeneous
nucleation behavior of liquid iron. Metall. Trans. 1, 19871995 (1970).
19. Porter, D. A., Easterling, K. E. & Sherif, M. Y. Phase Transformations in Metals
and Alloys 3rd edn (CRC Press, 2009).
20. Elliott, A. J. et al. Directional solidification of large superalloy castings with
radiation and liquid-metal cooling: a comparative assessment. Metall. Mater.
Trans. A 35, 32213231 (2004).
21. Kim, Y. W. Ordered intermetallic alloys, part III: gamma titanium aluminides.
JOM 7, 3039 (1994).
22. Appel, F., Oehring, M. & Paul, J. D. H. A novel in situ composite structure in
TiAl alloys. Mater. Sci. Eng. A 493, 232236 (2008).
23. Yuan, Y. et al. Dissociation of super-dislocations and the stacking fault energy
in TiAl based alloys with Nb-doping. Phys. Lett. A 358, 231235 (2006).
24. Paul, J. D. H., Appel, F. & Wagner, R. The compression behavior of niobium
alloyed -titanium aluminides. Acta Mater. 46, 10751085 (1998).
25. Chen, G. L. & Zhang, L. C. Deformation mechanism at large strains in a
high-Nb-containing TiAl at room temperature. Mater. Sci. Eng. A 329331,
163170 (2002).
26. Lu, L., Shen, Y. F., Chen, X. H., Qian, L. H. & Lu, K. Ultrahigh strength and
high electrical conductivity in copper. Science 304, 422426 (2004).
27. Lu, K., Lu, L. & Suresh, S. Strengthening materials by engineering coherent
internal boundaries at the nanoscale. Science 324, 349352 (2009).
28. Li, X., Wei, Y., Lu, L., Lu, K. & Gao, H. J. Dislocation nucleation governed
softening and maximum strength in nano-twinned metals. Nature 464,
877880 (2010).
29. Chen, K. C., Wu, W. W., Liao, C. N., Chen, L. J. & Tu, K. N. Observation of
atomic diffusion at twin-modified grain boundaries in copper. Science 321,
10661069 (2008).
30. Hsiao, H. Y. et al. Unidirectional growth of microbumps on (111)-oriented and
nanotwinned copper. Science 336, 11071010 (2012).
31. Wang, J. G., Zhang, L. C., Chen, G. L. & Ye, H. Q. Deformation-induced 2
phase transformation in a hot-forged Ti45Al10Nb alloy. Mater. Sci. Eng. A
252, 222231 (1998).
32. Liu, C. T. & Maziasz, P. J. Microstructural control and mechanical properties of
dual-phase TiAl alloys. Intermetallics 6, 653661 (1998).
33. Tsujimoto, T., Hashimoto, K. & Nobuki, M. Alloy design for improvement of
ductility and workability of alloys based on intermetallic compound TiAl.
Mater. Trans. JIM 33, 9891003 (1992).
34. Hanamura, T., Uemori, R. & Tanino, M. Mechanism of plastic deformation
of Mn-added TiAl L10-type intermetallic compound. J. Mater. Res. 3,
656664 (1988).
35. Heilmaier, M., Handtrack, D. & Varin, R. A. Creep properties of boron-doped
dual-phase Ti-rich L12 -based trialuminide intermetallics. Scr. Mater. 48,
14091414 (2003).
36. Rohatgi, A., Vecchio, K. S. & Iii, G. T. G. The influence of stacking fault energy
on the mechanical behavior of Cu and CuAl alloys: deformation twinning,
work hardening, and dynamic recovery. Metall. Mater. Trans. A 32,
135145 (2001).
37. Zhang, W. J., Liu, Z. C., Chen, G. L. & Kim, Y. W. Deformation mechanisms
in a high-Nb containing TiAl alloy at 900 C. Mater. Sci. Eng. A 271,
416413 (1999).
38. Appel, F., Paul, J. D. H. & Oehring, M. Gamma Titanium Aluminide Alloys:
Science and Technology 1st edn (Wiley-VCH, 2011).
39. Sastry, S. M. L. & Lipsitt, H. A. Fatigue deformation of TiAl base alloys. Metall.
Mater. Trans. A 8, 299308 (1977).
40. Larsen, J. M. Assuring reliability of gamma titanium aluminides in
long-term service. In Gamma Titanium Aluminides (eds Kim, Y.-W. et al.)
463472 (TMS, 1999).
41. Balsone, S. J., Wayne, J. J. & Maxwell, D. C. Fatigue crack growth in a cast
gamma titanium aluminide between 25 and 954 C. In Fatigue and Fracture
of Ordered Intermetallics Materials (eds Soboyejo, W. et al.) 307318
(TMS, 1994).
42. Chan, K. S. & Shih, D. S. Fundamental aspects of fatigue and fracture in a TiAl
sheet alloy. Metall. Mater. Trans. A 29, 7387 (1998).

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials

2016 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT4677

ARTICLES
43. Hnaff, G. & Gloanec, A. L. Fatigue properties of TiAl alloys. Intermetallic 13,
543558 (2005).
44. Dahms, M. Gamma titanium aluminide research and applications in Germany
and Austria. Adv. Perform. Mater. 1, 157182 (1994).
45. Gloanec, A. L., Hnaff, G., Bertheau, D., Belaygue, P. & Grange, M. Fatigue
crack growth behaviour of a gamma-titanium-aluminide alloy prepared by
casting and powder metallurgy. Scr. Mater. 49, 825830 (2003).
46. Sadananda, K. & Vasudevan, A. K. Fatigue crack growth behaviour in titanium
aluminides. Mater. Sci. Eng. A 192193, 490501 (1995).
47. Rao, K. T. V., Kim, Y. W., Muhlstein, C. L. & Ritchie, R. O. Fatigue-crack growth
and fracture resistance of a two-phase ( + 2 ) TiAl alloy in duplex and
lamellar microstructures. Mater. Sci. Eng. A 192193, 474482 (1995).
48. Recina, V., Lundstrm, D. & Karlsson, B. Tensile, creep, and low-cycle fatigue
behavior of a cast -TiAl-based alloy for gas turbine applications. Metall. Mater.
Trans. A 33, 28692881 (2002).

Acknowledgements
The authors are grateful for financial support from the National Key Basic Research
Program of China (grant 2011CB605504), and a Project Funded by PAPD of Jiangsu

Higher Education Institutions. We thank F. T. Kong, X.-Q. Chen, Y. Chen, G. D. Tang,


H. Nan, J. C. Li and X. Y. Cheng for beneficial discussions. C.T.L. was supported by
internal funding from City University of Hong Kong, Kowloon, Hong Kong.

Author contributions
G.C. designed and supervised the project. Y.B.P. synthesized the PST samples. G.Z.
conducted the quantitative calculation of the anisotropy of interfacial energy. Z.X.Q.
and M.Z.W. performed experiments and analysed the data. H.C.Y. and C.L.D. performed
high-temperature mechanical tests. C.T.L. assessed the outcomes. G.C., Y.B.P., G.Z.,
Z.X.Q., M.Z.W. and C.T.L. wrote the paper. All authors discussed the results and
commented on the manuscript.

Additional information
Supplementary information is available in the online version of the paper. Reprints and
permissions information is available online at www.nature.com/reprints.
Correspondence and requests for materials should be addressed to G.C. or C.T.L.

Competing financial interests


The authors declare no competing financial interests.

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials

2016 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT4677


Methods
Our Ti45Al8Nb alloy was prepared from 99.999% high-purity Ti, 99.99% Al and
99.95% Nb. The master ingots were prepared by cold-crucible induction-float
melting, and cast to rods from the ingot by suction casting. Then, the rod was put
in a high-purity alumina crucible, which is 5 mm in inner diameter with an yttria
coating for directional solidification. A liquid metal cooling Bridgman-type
apparatus was employed to produce directionally solidified bars under the
protection of 380 Pa high-purity argon (see Supplementary Fig. 3). After heating to
1,823 K and holding for 25 min, each rod was directionally solidified at different
withdrawal rates. An optical floating zone furnace (Model: FZ-T-4000-HSPC-NUST) was employed to prepare TiAl PST single crystals for mechanical tests.
Structural characterization. An Olympus GX-41 optical microscope was used to
characterize the microstructure. The microstructures of the polished specimens
were etched with the agent comprising 90 ml distilled water, 5 ml HNO3 and 5 ml

ARTICLES
HF. Structural characterization of the twins was carried out by means of an FEI
TecnaiG2 20 LaB6TEM and HRTEM. The sample preparation for TEM
observations followed the standard route of cutting, ion thinning and
final polishing.
Mechanical testing. After homogenizing and stress relief annealing of PST
samples, multiple mechanical tests were performed based on the load and
elongation measurement. Tensile specimens with a gauge section of
1 2.5 10 mm were prepared by low-stress grinding and polishing with 2000 grit
sand paper. Ambient-temperature tensile tests were performed at a strain rate of
2 104 s1 using an Instron 5982 testing machine. The high-temperature tensile
tests were performed on an Instron 5500R testing machine at a strain rate of
1 103 s1 . The uniaxial tensile creep tests were performed on an Instron 5982
testing machine under constant loads of 100, 150 and 210 MPa at 900 C in air.
The high-temperature specimens have a gauge section of 83 15 mm.

NATURE MATERIALS | www.nature.com/naturematerials

2016 Macmillan Publishers Limited. All rights reserved

Вам также может понравиться