Sie sind auf Seite 1von 8

Materials Science and Engineering A 375377 (2004) 3845

Nanostructured surface layer on metallic materials induced


by surface mechanical attrition treatment
K. Lu a, , J. Lu b
a

Institute of Metal Research, Chinese Academy of Sciences, Shenyang National Laboratory for Materials Science,
72 Wenhua Road, Shenyang 110016, China
b LASMIS, University of Technology of Troyes, 10000 Troyes, France

Abstract
In terms of the grain refinement mechanism induced by plastic straining, a novel surface mechanical attrition treatment (SMAT) was
developed for synthesizing a nanostructured surface layer on metallic materials in order to upgrade the overall properties and performance.
In this paper, the SMAT technique and the microstructure of the SMAT surface layer will be described. The grain refinement mechanism
of the surface layer during the SMAT will be analyzed in terms of the microstructure observations in several typical materials. Obvious
enhancements in mechanical properties and tribological properties of the nanostructured surface layer in different materials were observed.
Further development and prospects will be addressed with respect to the SMAT as well as the performance and technological applications of
the engineering materials with the nanostructured surface layer.
2003 Elsevier B.V. All rights reserved.
Keywords: Nanostructured materials; Surface; Mechanical attrition; Grain refinement; Properties

1. Introduction
In most cases, material failures occur on surfaces such
as fatigue fracture, fretting fatigue, wear and corrosion, etc.
These failures are very sensitive to the structure and properties of the material surface. Optimization of the surface
microstructure and properties is an effective approach to enhance the global behavior and service lifetime of materials.
With the intensive and extensive investigations on nanostructured materials in the past decades, more and more experimental evidence showed that this new class of materials possess novel properties and performances that are fundamentally different from their conventional coarse-grained polycrystalline counterparts, such as high hardness and strength
[13], enhanced physical properties [13], improved tribological properties [4], and superplasticity at low temperatures [5,6], etc. It is reasonable to expect to achieve surface
modification by generation of a nanostructured surface layer
so that the overall properties and behavior of the materials
are significantly improved.

Corresponding author. Tel.: +86-24-2390-6826;


fax: +86-24-2399-8660.
E-mail address: lu@imr.ac.cn (K. Lu).
0921-5093/$ see front matter 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2003.10.261

Conventionally, a nanostructured surface layer can be


made on a bulk material by means of various coating
and deposition technologies such as PVD, CVD, sputtering, electrodeposition, and plasma processing, etc. (as
illustrated in Fig. 1a). The coated materials can be either
nanometer-sized isolated particles or polycrystalline powders with nano-sized grains. The coated layer and the matrix can be different or made of the same kind of material.
The predominant factors in this process are the bonding of
the coated layer with the matrix and the bonding between
particles while maintaining the nanostructure.
An alternative approach to synthesis of a nanostructured
surface layer is to transform the original coarse-grained
surface layer of a bulk material into nano-sized grains
while keeping the overall composition and/or phases unchanged. Such a process may be referred as surface selfnanocrystallization of bulk materials, as schematically
shown in Fig. 1b.
So far, three kinds of techniques have been developed for
synthesizing bulk nanostructured materials [7]:
(1) Starting with isolated nanometer-sized particles. The
nanometer-sized particles can be generated by means
of various techniques including PVC, CVD, electrochemical and hydrothermal methods, precipitation from

K. Lu, J. Lu / Materials Science and Engineering A 375377 (2004) 3845

Fig. 1. Schematic illustration of three types of surface nanocrystallization processes: (a) surface coating or deposition; (b) surface
self-nanocrystallization; (c) hybrid surface nanocrystallization.

solution, and so on. The most widely applied PVD


method involves inert gas condensation. The second step
is to consolidate these ultrafine particles into bulk materials in which the particles become nano-sized grains.
(2) Starting with a noncrystalline structure. Nanocrystalline
materials are obtained by nucleating numerous crystallites in the glass by annealing or mechanical activation
[3]. These nuclei subsequently grow together (with a relatively low growth rate) and result in a nanostructured
material.
(3) Starting with a coarse-grained polycrystalline structure.
By increasing the free energy of the polycrystals and
generation of much more defects and interfaces (grain
boundaries) in various nonequilibrium processes, such
as ball-milling [8], severe plastic deformation [9], irradiation with high-energy particles, spark erosion, and
sliding wear [10], etc., the polycrystalline structure will
be transformed into nanocrystalline structures via different kinds of grain refinement mechanisms.

39

In terms of the underlying mechanism of the method


(3), i.e., grain refinement induced by plastic deformation,
we developed a new technique, namely surface mechanical
attrition treatment (SMAT), to synthesize a nanostructured
surface layer on bulk metallic materials [11]. Owing to the
plastic deformation in the surface layer induced by the mechanical attrition, the coarse-grained structure in the surface
layer is refined into the nanometer scale without change of
the chemical compositions. This SMAT has been successfully applied in various kinds of materials including pure
metals, steels, and alloys [1216], on which a nanostructured
surface layer up to 50 m thick has been obtained. And the
obvious enhancement in the overall properties and performance of the materials is observed after the SMAT treatment.
In this paper, we will report the processing of the SMAT,
the microstructure and grain refinement mechanism of the
surface layer induced by the SMAT, as well as properties of
the nanostructured surface layer in different materials.

2. Surface mechanical attrition treatment


The key point for realizing the surface self-nanocrystallization of a bulk material is to introduce a large amount of
defects and/or interfaces into the surface layer so that its microstructure is transformed into nano-sized crystallites. Or in
other words, a grain refinement process into the nano-scale
is needed in the surface layer while the structure of the
coarse-grained matrix remains unchanged. Surface mechanical attrition is an effective technique to realize the surface
self-nanocrystallization on metallic materials.
Fig. 2 illustrates the experimental set-up of the SMAT.
Spherical steel balls with smooth surface (or of other materials such as glass and ceramics) are placed in a reflecting
chamber that is vibrated by a vibration generator. Typical
ball sizes are 110 mm in diameter and that can be different
for different materials. The vibration frequency of the chamber is in the range from 50 Hz to 20 kHz. When the balls are
resonated, the sample surface to be treated is impacted by a

Vacuum
Sample

sample

Vibration
generator
(a)

(b)

Fig. 2. (a and b) Schematic illustration of the surface mechanical attrition treatment set-up and the repeated multidirectional plastic deformation in the
sample surface layer induced by impact of the flying balls.

40

K. Lu, J. Lu / Materials Science and Engineering A 375377 (2004) 3845

large number of flying balls over a short period of time. The


velocity of the balls is about 120 m/s, depending upon the
vibration frequency, the distance between the sample surface to the balls, and the ball size. The impact directions of
the balls onto the sample surface are rather random due to
the random flying directions of the balls inside the vibration chamber. Each impact will induce plastic deformation
with a high strain rate in the surface layer of the sample, as
schematically shown in Fig. 2b. As a consequence, the repeated multidirectional impacts at high strain rates onto the
sample surface result in severe plastic deformation and grain
refinement progressively down to the nanometer regime in
the entire sample surface. The temperature rise at the sample surface induced by the repeated impacts was measured
in a Fe sample, being about 50100 C, which varies with
the intensity of impacts and the materials treated.
Comparing the SMAT with other conventional surface
treatment such as shot peening, one may find obvious differences in several aspects. Much larger balls (a few mm)
are used in the SMAT than in the shot peening (0.21 mm
in diameter). Spherical balls with smooth surface are necessary for obtaining a nanostructured surface layer in the
SMAT. Balls with rough surface (as in shot peening) will
wear and damage the nanostructured surface layer during
the treatment. The velocity of the balls in the SMAT process
is much lower (120 m/s) compared with the conventional
shot peening (typically about 100 m/s). Conventional shot
peening is a directional process in which the angle between
the shot jet and the sample surface is normally fixed, close
to 90 in many case. But in the SMAT, random directional
impacts of the balls onto the sample are needed in order to
facilitate the grain refinement process.
Our experimental results on a number of materials including pure metals and alloys show that the nanostructured surface layer can be up to 50 m thick, in which
grain sizes vary gradually from a few nanometers (in the
top treated surface layer) to about 100 nm. Underneath is
a refined structured layer (up to 100 m thick) consisting
of submicrometer-sized crystallites or cells separated by either grain boundaries or subboundaries. In deeper layers are
deformed coarse grains with various kinds of dislocation
configurations such as dense dislocation walls, dislocation
tangles, and dislocation cells. Fig. 3 shows a scheme for a
cross-sectional view of the SMAT sample. The thickness of
the nanostructured surface layer (as well as the refined structure layer) depends very much upon the material treated and
the processing parameters (such as ball size, vibration frequency, temperature, etc.).
In fact other mechanical treatment techniques are also
possible to achieve surface nanostructure layers on metallic materials when large strains with a high strain rate are
achieved, such as shot peening, hammer peening, laser shock
treatment, surface rolling, and high-speed machining. However, the thickness of the nanostructured surface layer may
be different for various treatment techniques due to the different strains and strain rates applied in the surface layer.

With a nanostructured surface layer prepared, it is possible to modify the composition and/or phases of the nanostructured surface layer by exposure of the treated surface to
different media that can be solid, liquid or gaseous. Solid
solutions, compounds, or composites might be formed in
the nanostructured surface layer so that specific properties
can be obtained. This type of surface nanocrystallization
process, i.e., hybrid surface nanocrystallization (shown in
Fig. 1c) may provide an efficient way to enhance the surface properties. This process can be possibly performed by
an in situ process combining the SMAT with the chemical
reaction simultaneously.

3. Surface nanocrystallization mechanism


Understanding of the formation process of nanocrystallites during the SMAT process is crucial for development
of the SMAT. Owing to the gradient variation of the strain
and strain rate from the treated top surface (both are very
large) to the deep matrix (essentially zero, as schematically
shown in Fig. 3), a gradient grain size distribution from a few
nanometers (in the top surface layer) to several micrometers
is developed in the SMAT sample, that provides a unique
opportunity to examine the microstructure characteristics at
different levels of strain and strain rate. Therefore, the underlying mechanism for deformation-induced grain refinement
in the micrometernanometer regime can be deduced.
Analogous to the grain refinement mechanism during
plastic deformation of bulk metals, formation of nanostructures from the coarse-grained polycrystals in the surface
layer upon the SMAT involves various dislocation activities
and development of grain boundaries. Plastic deformation behaviors and dislocation activities in metals depend
strongly on the lattice structure and the stacking fault energy (SFE). For example, in those materials with high SFEs,
dislocation walls and cells will be formed to accumulate
strains and subboundaries are formed to subdivide coarse
grains. While for those with low SFEs, plastic deformation
mode may change from dislocation slip to mechanical twins
(especially under high strain rate and/or low temperature)
[17]. In the following, we take three typical examples with
different lattice structures and SFEs to demonstrate the surface nanocrystallization process upon the SMAT treatment.
3.1. Fe [16]
Fe is a typical bcc metal with a high SFE of about
200 mJ/m2 . After a SMAT to a pure Fe bulk sample, obvious
grain refinement was observed in the surface layer in which
the top layer consists of ultrafine crystallites of about 10 nm
in size, as indicated in Fig. 4. The grain/cell size increases
gradually from 10 nm to several micrometers at about
60 m deep. Detailed cross-sectional TEM observations
of the as-treated Fe sample revealed the grain refinement
process, which involves the following elemental processes:

K. Lu, J. Lu / Materials Science and Engineering A 375377 (2004) 3845

41

Fig. 3. Schematic illustration of microstructure characteristics and distributions of strain and strain rate along depth in the surface layer subjected to the
SMAT. Based on the velocity of the balls and the measured depth of the pit caused by an individual impact, the strain rate at the sample surface was
estimated to be as much as 1023 s1 .

tions generated in previous deformation. Therefore, the


grains can be subdivided efficiently by the DDWs and
DTs in the treatment.
(2) Transformation of DDWs and DTs into subboundaries
with small misorientations: Formation of subgrain
boundaries subdividing the original grains is resulted
from development of DDWs and DTs by accumulating
of more and more dislocations with increasing strains.
At a certain strain level, for minimizing the total system energy, dislocation annihilation and rearrangement
occur in DDWs and DTs, which will transform into
subboundaries (with larger misorientations across relative to the DDWs) separating individual cells. The subgrain boundaries are actually formed by recombination

(1) Development of dense dislocation walls (DDWs) and


dislocation tangles (DTs): In order to accommodate
plastic strains, dislocation activities are motivated in
the original coarse grains, forming DDWs along (1 1 0)
slip planes and random DTs in some grains (depending on the grain orientations). Development of these
dislocation configurations results in subdivision of
original grains by forming individual dislocation cells
primarily separated by DDWs and DTs. The repeated
multi-directional impact onto the sample surface may
lead to a change of slip systems with the strain path
even inside the same grain. The dislocations not only
interact with other dislocations in the current active
slip systems, but also interact with inactive disloca-

0.15
5

10

0.10

10

0.05

10

10

0.00

10

Mean microstrain (%)

Grain/cell size (nm)

10

10

20

30

40

50

60

70

80

Distance from surface (m)


Fig. 4. Variations of the grain/cell size with the depth from the SMAT surface of the Fe sample determined by means of XRD analysis (), TEM (()
equiaxed, (H) short-axis, (N) long-axis) and SEM () observations. The mean microstrain () was also determined at different depths by means of XRD
analysis.

42

K. Lu, J. Lu / Materials Science and Engineering A 375377 (2004) 3845

of a high density of dislocations and usually have small


misorientations of a few degrees, i.e., small-angle grain
boundaries.
(3) Evolution of subboundaries to highly-misoriented grain
boundaries: With further increasing strains, more dislocations are generated and annihilated in the subboundaries, raising up the energy of the subboundary
and progressively increasing its misorientations. The
orientations of the grains with respect to their neighboring grains become completely random, forming
highly-misoriented grain boundaries. The increment of
misorientations between neighboring grains can be realized by accumulating more dislocations with different
Burgers vectors in grain boundaries, or alternatively,
by rotation of grains (or grain boundary sliding) with
respect to each other under certain strains. The grain rotation process would be much facilitated when the size
of grains is reduced due to the obvious size dependence
[18].
With further straining, DDWs and DTs could form inside
the inner of the refined subgrains or grains, and these refined grains could be further subdivided following the similar
mechanism. With increasing strain in the top surface layer,
the subdivision takes place on a finer and finer scale. When
dislocation multiplication rate is balanced by the annihilation rate, the increase of strains could not reduce the subgrain size any longer, and a stabilized grain size is resulted.
In an Al alloy with a high SFE subjected to the SMAT, a
similar grain refinement mechanism was found [15].
3.2. Cu [19]
Cu is an fcc metal with a medium SFE (about 78 mJ/m2 ).
After the SMAT to a bulk Cu specimen, a nanostructured
surface layer of about 35 m thick was formed. Microstructure examination by means of cross-sectional TEM observations revealed a different grain refinement mechanism:
(1) Development of equiaxed dislocation cells (DCs): For
the fcc structure, more dislocation slip planes exist compared with that in bcc Fe. The strain-induced dislocation
activities lead to formation of equiaxed DCs (instead of
DDWs or DTs as in Fe) where high density of dislocations locate at the cell boundary with few dislocations
inside the cell. The cell size observed ranges from a few
micrometers to several tens of nanometers, depending
upon the strain and strain rate. With increasing strains,
the DC size decreases and forming a DC network that
subdivides the original coarse grains.
(2) Formation of twins and subboundaries with small misorientations: At a certain strain and strain rate level,
mechanical twinning is activated in some grains with
favorable orientations. Twins appear at depth from 10
to 100 m from the top surface, implying the strains
and strain rate at this depth is appropriate for formation of twins in Cu with the medium SFE. At the same

time, in other grains, DCs transform into subboundaries


with small misorientations with further straining, which
is analogous to the transformation from DDWs as in
Fe. These twin boundaries and subboundaries divide the
original coarse grains into ultrafine (sub)grains.
(3) Evolution of subboundaries to highly-misoriented grain
boundaries: Increasing strains will drive the transformation from subgrain boundaries into conventional grain
boundaries with large misorientations, similar to the process as in Fe.
3.3. AISI 304 stainless steel [20]
AISI 304 stainless steel is a widely used engineering material with an fcc austenite structure and a very low SFE
(about 17 mJ/m2 ). Due to the low SFE, dislocation activities
in plastic deformation are much different from the above
two cases. The grain refinement mechanism was found as
follows:
(1) Formation of planar dislocation arrays and twins: The
strain-induced dislocations in the austenite phase slip
mainly on their respective {1 1 1} planes, forming regular dislocation grids, instead of irregular DCs (as in fcc
Al alloy, Cu) or DDWs (as in bcc Fe). Such a difference
originates from the low SFE that makes it difficult for
partial dislocations to cross-slip for forming DCs, rather
favorites formation of dislocation arrays and twins in
{1 1 1} planes.
(2) Grain subdivision by twins and martensite transformation: With increasing strains, twins in different {1 1 1}
planes intersect with each other and subdivide the original austenite grains into refined blocks. A strain-induced
martensite transformation is a prevailing phenomenon
in the plastic deformed AISI 304 stainless steel. In
the SMAT-treated sample, we observed formation of
martensite phase at intersections of the twins, of which
the size ranges from several nanometers to submicrometers. Such a phase transformation is a unique process
that facilitates grain refinement procedures.
(3) Formation of nanocrystallites: In the top surface layer,
manometer-sized martensite crystallites are formed by
means of further twin-twin intersection and martensite phase transformation under high strains and strain
rates. Analogous to other systems, randomly-oriented
nanocrystallites are formed via grain rotation or other
grain boundary activities (motion or sliding) to accommodate the high strains.

4. Properties of surface nanostructured layers


Mechanical property measurements indicated a significant increment of hardness and strength in the surface layer
with nanostructures after the SMAT. Fig. 5 shows a variation of hardness along the depth from the treated surface

K. Lu, J. Lu / Materials Science and Engineering A 375377 (2004) 3845

43

Hardness (GPa)

as-treated
as-annealed at 923 K

1
0

50

100

150

200

Depth from surface (m)


Fig. 5. Measured hardness as a function of depth from the top surface for the SMAT Fe sample and for the annealed (at 593 K for 30 min) sample.

as determined by using nanoindentation tests in the SMAT


Fe sample. In the top surface nanostructured layer, hardness
reaches as high as 3.8 GPa, which is about twice that for the
coarse-grained matrix. There is no change in hardness profile after the sample was annealed 593 K for 1 h for residual
stress relaxation, implying that the high hardness in the surface layer is not resulted from the residual stress induced in
the SMAT. After annealing at 923 K for 1 h for recrystallization of the nanostructures and forming coarse grains, hardness in the surface layer drops to that of the coarse-grained
matrix. This observation indicates that the hardness increment is due to grain refinement into the nanometer scale, instead of the alloying (contamination) effect from the SMAT
media (balls and other impurities). Plotting the hardness values as a function of measured grain size along the depth,
one may check the validity of Hall-Petch relation in a wide
grain size range (from 10 nm to 50 m) in one sample. A
normal Hall-Petch relation is observed in Fe as far as the
grain size is around 10 nm.
Tensile yield strength of a SMAT low-carbon steel sheet
(1.5 mm thick) was found to be enhanced by about 35% relative to the untreated sample, while the elongation-to-failure
remains unchanged. Such a large increment in yield strength
can be reasonably attributed to the strong nanostructured
surface layers (about 20 m thick) on both sides of the sample. The yield strength of the surface nanostructured layer is
estimated to be about 23 times that of the coarse-grained
form. A similar phenomenon was observed in a 316L
stainless steel sample in which both the yield strength and
the fracture strength are enhanced. Fig. 6 shows a tensile
stress-strain curve for a SMAT 316L stainless steel sample (1 mm thick with 10 m thick nanostructured surface
layers on both sides, gauge length of 46 mm and width of
5 mm), in comparison with an untreated original sample

with the same geometry. The yield strength of the sample


at a strain rate of 103 s1 increases from 280 MPa (original) to 550 MPa after the SMAT, while the ultimate tensile
stress is increased by about 13% (from 620 to 700 MPa).
Analysis of the fracture mechanism of the SMAT-treated
sample revealed that the nanostructured surface layers on
both sides of the sample, which are much stronger than the
coarse-grained matrix, obstruct the slip bands developed in
the coarse grains in the inner part of the sample. So that
crack nucleation will not occur in the sample surface due to
the much high yield strength of the surface layer, but in the
sub-surface layers. Therefore, the stress-to-failure is much
enhanced for the sample with nanostructured surface layers.
Wear and friction properties of the low-carbon steel sheet
after the SMAT were measured by using a reciprocating friction tester with a diamond tip in comparison with that of the
original steel sheet [21]. As shown in Fig. 7, the wear volume
loss of the nanostructured surface layer in the SMAT sample
is lower than that for the untreated original one. The friction
coefficient values at different applied loads for the as-treated
sample are evidently smaller (about a half) than those for the
original sample. These observations indicate that the friction and wear properties of the low-carbon steel can be improved by means of formation of the nanostructured surface
layer. The improved wear and fraction properties can be attributed to the strong surface layer with nanograins and a
gradient variation in the microstructure and properties along
the depth from top surface.
As grain boundaries may act as fast-diffusion channels,
atomic diffusivity in the nanostructured surface layer will be
considerably enhanced compared to the conventional polycrystalline materials. This behavior can be utilized to upgrade the traditional surface chemical treatment techniques,
such as nitriding process of steel, by enhancing the diffusion

44

K. Lu, J. Lu / Materials Science and Engineering A 375377 (2004) 3845

800

Treated material
600

Stress (MPa)

Base Material

400

200

0
0

10

15

20

25

30

Strain (%)
Fig. 6. Tensile test stress-strain curves for the SMAT 316L stainless steel sheet sample (1.5 mm thick) with nanostructured surface layers (of about 15 m
thick) on both sides, and for the untreated sample (base material).

0.15

Friction Coefficient

Wear Volume Loss, mm

0.4

as-treated
original

0.10

0.05

as-treated
original

0.3

0.2

0.1

0.0

0.00
0

(a)

10

Load, N

(b)

10

Load, N

Fig. 7. Variations of the wear volume loss with load (a) and variations of the coefficient of friction with load (b) for the SMAT and the original low-carbon
steel samples.

kinetics or reducing the diffusion temperature. Our preliminary results showed that the nitriding temperature of iron can
be reduced to about 300 C after the SMAT to form a nanostructured surface layer, in contrast to more than 500 C for
conventional polycrystalline Fe and steels [22]. Enhanced
diffusivity of other elements (Cr, Al) in the nanostructured
surface layer is also observed experimentally.

5. Summary and prospects


The evidences obtained so far have already indicated that
the nanostructured surface layer synthesized by means of

the SMAT to metallic materials provides a plenty of unique


opportunities in both basic scientific research and technological applications as well, including:
(i) To enhance the surface mechanical, tribological, chemical and corrosion properties of bulk materials;
(ii) To investigate the strain-induced grain refinement
mechanism in a wide grain size scale (micro-nanometer);
(iii) To study the structureproperty relationship of solid
in a wide grain/cell size range (several nanometers
micrometers) in ONE sample. The gradient structure
in the surface layer allows one to prepare porosity-free
microsamples for property measurements with different

K. Lu, J. Lu / Materials Science and Engineering A 375377 (2004) 3845

grain sizes (in surface layers at different depths) but the


same composition.
(iv) New alternative approaches to functional surface structures by means of nanostructure-selective reaction due
to the much enhanced atomic diffusivity and chemical
reactivity of the nanostructured surface layer.
(v) Flexibility and low-cost procedures of the SMAT will
greatly facilitate the widespread application of this
technique to various industry areas. The process is also
feasible to obtain a localized nanostructured surface
layer in bulk materials and to realize surface nanocrystallization of components with complex shapes.
Other advantages of this technique include the fact that
many existing processes can be used to obtain the nanostructured surface layer with a high productivity. Surface
nanocrystallization of metallic materials will certainly provide a complementary process to the nanocrystallization process for bulk materials. With increasing investigations on
the processing and properties of the nanostructured surface
layer, industrial applications of this new technique to upgrade the traditional engineering materials and technologies
in the near future can be anticipated.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]

Acknowledgements
Financial support from the National Science Foundation
of China, the Ministry of Science and Technology of China
(Grant G1999064505), NEDO International Joint Research
Grant Program (01MB5), and Ministry of Research of
France (Grant 2001882, CPER EN2040) is acknowledged.
The authors thank G. Liu, N.R. Tao, M.L. Sui, H.W. Zhang,
K. Wang, and W.P. Tong for their valuable contributions to
the present work.

45

[19]
[20]
[21]
[22]

H. Gleiter, Prog. Mater. Sci. 33 (1988) 223.


C. Suryanarayana, Int. Mater. Rev. 40 (1995) 41.
K. Lu, Mater. Sci. Eng. R16 (1996) 161.
D.G. Morris, Mechanical Behaviour of Nanostructured Materials,
Trans. Tech. Publications Ltd., Switzerland, 1998, p. 70.
S.X. McFadden, R.S. Mishra, R.Z. Valiev, A.P. Zhilyaev, A.K.
Mukherjee, Nature 298 (1999) 684.
L. Lu, M.L. Sui, K. Lu, Science 287 (2000) 1463.
H. Gleiter, in: R.W. Cahn, P. Haasen (Eds.), Physical Metallurgy,
Elsevier, Amsterdam, 1996, p. 843.
C.C. Koch, Nanostruct. Mater. 2 (1993) 109.
R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Prog. Mater. Sci. 45
(2000) 103.
D.A. Hughes, N. Hansen, Phys. Rev. Lett. 87 (2001) 135503-1.
K. Lu, J. Lu, J. Mater. Sci. Technol. 15 (1999) 193.
N.R. Tao, M.L. Sui, J. Lu, K. Lu, Nanostruct. Mater. 11 (1999) 433.
G. Liu, J. Lu, K. Lu, Mater. Sci. Eng. A286 (2000) 91.
G. Liu, S.C. Wang, X.F. Lou, J. Lu, K. Lu, Scripta Mater. 44 (2001)
1791.
X. Wu, N. Tao, Y. Hong, B. Xu, J. Lu, K. Lu, Acta Mater. 50 (2002)
2075.
N.R. Tao, Z.B. Wang, W.P. Tong, M.L. Sui, J. Lu, K. Lu, Acta
Mater. 50 (2002) 4603.
B. Bay, N. Hansen, D.A. Hughes, D. Kuhlmann-Wilsdorf, Acta
Metall. Mater. 40 (1992) 205.
H. van Swygenhoven, D. Farkas, A. Caro, Phys. Rev. B 62 (2000)
831.
K. Wang, G. Liu, J. Lu, K. Lu, in press.
H.W. Zhang, Z.K. Hei, G. Liu, J. Lu, K. Lu, Acta Mater. 51 (2003)
1871.
Z.B. Wang, X.P. Yong, N.R. Tao, S. Li, G. Liu, J. Lu, K. Lu, Acta
Metall. Sinica 37 (2001) 1252.
W.P. Tong, N.R. Tao, Z.B. Wang, J. Lu, K. Lu, Science 299 (2003)
686.

Das könnte Ihnen auch gefallen