0% found this document useful (0 votes)
32 views18 pages

Pande 2009

This document reviews models of the Hall-Petch relationship in nanocrystalline materials. It discusses how the classical Hall-Petch model based on dislocation pile-ups does not fully explain experimental data showing abnormal softening with very small grain sizes less than 12 nm. A model incorporating both dislocation interactions and grain boundary sliding is more successful. The document also examines competing deformation mechanisms like Coble creep and provides a generalized expression for yield stress that considers realistic grain size distributions.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
32 views18 pages

Pande 2009

This document reviews models of the Hall-Petch relationship in nanocrystalline materials. It discusses how the classical Hall-Petch model based on dislocation pile-ups does not fully explain experimental data showing abnormal softening with very small grain sizes less than 12 nm. A model incorporating both dislocation interactions and grain boundary sliding is more successful. The document also examines competing deformation mechanisms like Coble creep and provides a generalized expression for yield stress that considers realistic grain size distributions.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 18

Progress in Materials Science 54 (2009) 689–706

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/

Nanomechanics of Hall–Petch relationship


in nanocrystalline materials
C.S. Pande, K.P. Cooper *
Materials Science and Technology Division, Naval Research Laboratory, Washington, DC 20375-5343, USA

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

Classical Hall–Petch relation for large grained polycrystals is


usually derived using the model of dislocation pile-up first investi-
gated mathematically by Nabarro and coworkers. In this paper the
mechanical properties of nanocrystalline materials are reviewed,
with emphasis on the fundamental physical mechanisms involved
in determining yield stress. Special attention is paid to the abnor-
mal or ‘inverse’ Hall–Petch relationship, which manifests itself as
the softening of nanocrystalline materials of very small (less than
12 nm) mean grain sizes. It is emphasized that modeling the
strength of nanocrystalline materials needs consideration of both
dislocation interactions and grain-boundary sliding (presumably
due to Coble creep) acting simultaneously. Such a model appears
to be successful in explaining experimental results provided a real-
istic grain size distribution is incorporated into the analysis.
Masumura et al. [Masumura RA, Hazzledine PM, Pande CS. Acta
Mater 1998;46:4527] were the first to show that the Hall–Petch
plot for a wide range of materials and mean grain sizes could be
divided into three distinct regimes and also the first to provide a
detailed mathematical model of Hall–Petch relation of plastic
deformation processes for any material including fine-grained
nanocrystalline materials. Later developments of this and related
models are briefly reviewed.
Prof. Frank Nabarro was a physicist by training, a metallurgist by
profession and a genius by nature, blessed with a unique ability to
treat everyone as his equal. During his later years he was very
much interested in the mechanical properties of nanocrystalline
materials. This review on that topic is our contribution to the spe-
cial issue of Progress in Materials Science honoring him.
Published by Elsevier Ltd.

* Corresponding author.
E-mail address: [email protected] (K.P. Cooper).

0079-6425/$ - see front matter Published by Elsevier Ltd.


doi:10.1016/j.pmatsci.2009.03.008
690 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690
2. Experimental background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690
3. Mechanisms of deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
4. Models using lattice dislocations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 693
5. Role of Coble creep as a competing mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695
6. A generalized expression for yield stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
7. Relationship between hardness and yield strength in metals and alloys . . . . . . . . . . . . . . . . . . . . . . . . 700
8. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703

1. Introduction

Nanocrystalline materials are polycrystalline materials consisting of grains in nanometer range.


They have the potential to exhibit outstanding physical, mechanical and chemical properties, which
could, in principle, lead to new applications and novel technologies (see Refs. [1–5]). There are now
several examples of real applications even at present where the outstanding physical mechanical
and chemical properties of nanomaterials are used for commercial products (e.g. soft magnets, coat-
ings, structural repair, etc.). These outstanding properties can be due to interface and nano-scale ef-
fects due to the high volume fraction of the interfacial phase (up to 50%), and smaller mean grain
sizes (not exceeding 100 nm). Of special importance and subject of this review are the unique mechan-
ical properties of nanocrystalline materials, especially yield stress, which are essentially different from
those of conventional coarse-grained polycrystalline materials. For example, bulk nanocrystalline
materials and thin nanocrystalline coatings in some cases show superhardness, extremely high
strength and good fatigue resistance [5–14], which are desired for many applications. The high
strength however is often accompanied by low ductility at room temperature, which may limit their
practical utility. (However, more recently, several researchers have claimed substantial strength as
well as high tensile ductility in nanocrystalline materials [15–18].) It has been also surmised that
some nanocrystalline ceramics and metallic alloys may even exhibit superplasticity at lower temper-
atures and faster strain rates than their coarse-grained counterparts [19–29]. There is also experimen-
tal evidence [30–33] that nanocrystalline materials may show anomalously fast diffusion, which may
in turn explain their deformation behavior. These developments lead one to hope that nanocrystalline
materials with unique combination of high strength and good ductility may provide new structural
and functional applications in the future. The main aim of this paper is to provide a brief overview
of the theoretical models of yield stress in nanocrystalline materials paying special attention to their
microscopic mechanisms. Once established, these mechanisms are then expected to provide theoret-
ical underpinnings to the mechanical behavior of nanocrystalline materials.

2. Experimental background

Mechanical behavior of nanocrystalline materials has been the theme of over 500 publications and
several review articles [3,34–50]. These articles conclude that yield stress and microhardness of nano-
crystalline materials can be 2–10 times higher than the corresponding coarse-grained polycrystalline
materials with the same chemical composition. Similarly, some published values of microhardness of
nanocrystalline composite coatings [13,14] are of the same order as microhardness (HV  70–90 GPa)
of diamond. In the range of grain sizes d above about 10 nm, the dependence of the yield stress s on d
deviates little from the classical Hall–Petch relationship given by the formula,
s ¼ s0 þ kd1=2 ð1Þ
with s0 and k being material constants [6–12]. Yield stress may also depend upon on the mode of pro-
cessing [51,52]. However, any further grain refinement may lead to lower yield stress. Thus, in the
C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 691

Fig. 1. Scaled yield stress as a function of (grain size)1/2 for several materials [53].

range of smaller grain sizes, heat-treated materials exhibit the so-called ‘inverse’ Hall–Petch behavior
(softening with further reduction of grain size). Masumura et al. [53] have plotted some of the avail-
able data (till 1998) in a Hall–Petch plot (see Fig. 1). It is seen that the yield stress-grain size exponent
for relatively large grains appears to be very close to 1/2, as in Eq. (1), and generally this trend con-
tinues until the very fine grain regime (100 nm) is reached. The large scatter of the data for grain
sizes below 100 nm could be attributed to problems in preparing these materials or to differences
in thermal treatments. With the advent of better prepared nanocrystalline materials whose grain sizes
are of nanometer (nm) dimensions, the applicability and validity of Eq. (1) as well as the underlying
mechanisms became of great interest. In addition, these nanocrystalline materials were found to
exhibit, in general, low tensile ductility at room temperature [6–12]. However more recent results
indicate that nanocrystalline materials with very low porosity [15] or with dendrite-like inclusions
[17,18] or with bimodal grain size distributions (consisting of both nano- and micron-sized grains)
[16] show better ductility. Some reports also indicate nanocrystalline materials with high-strain-rate
(tensile) superplasticity [21–29].
As far as microstructures in these materials are concerned, mechanically loaded nanocrystalline
materials are reported to show grain rotations [27,54], formation of shear bands [55–59], or emis-
sion of (usually) partial lattice dislocations by grain boundaries into grain interiors [27,29,60]. Fig. 2
shows a Hall–Petch plot for copper using early data from various researchers. It also defines the
three regions of the Hall–Petch plot. The limitation of the classical ideas of Hall–Petch plots is dra-
matically demonstrated in Fig. 3 by plotting the yield strength data in terms of grain size instead of
(grain size)1/2.

3. Mechanisms of deformation

As early as 1977, Armstrong and coworkers [61] noted the increase in yield stress on grain refine-
ment up to the beginning of the nanocrystalline (<100 nm) regime. Much effort has been spent to
theoretically describe the Hall–Petch relationship in nanocrystalline materials. Classically, high
692 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

Fig. 2. Compilation of yield stress data for pure copper from various publications [53].

Fig. 3. Plots showing limitation of standard Hall–Petch law at small grain sizes and existence of optimum grain size for yield
strength [49,53]. (a) Schematic of hardness or strength as a function of normalized grain size shows the limitation dramatically.
(b) Normalized yield strength plotted against (normalized grain size)1/2.

values for yield stress were considered to be related to the effect of increased grain boundaries pro-
viding additional obstacles for movement of lattice dislocations.
In early theoretical studies, models of nanocrystalline materials were considered as two-phase
composites consisting of nanograin interiors and grain-boundary regions (see, e.g. Refs. [55,62–
68]). Yield stress s is then accounted for using the so-called rule-of-mixture, yield stress s being
given by some weighted sum of the yield stresses characterizing the grain-interior and grain-bound-
ary phases. The ratio of the two phases, of course, strongly depends on the grain size d. In this
calculation, the yield stress of the grain-boundary phase is assumed to be lower than that of the
grain-interior phase, and with suitable adjustable parameters, the deviations from the conventional
Hall–Petch relationship can be described roughly in accordance with experimental data. It is
obvious that the ‘‘rule-of-mixture” approach is too approximate and arbitrary and sheds no light
on the actual mechanisms [11].
C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 693

Our goal is to briefly review a more precise physical mechanism of plastic flow in nanocrystalline
materials in terms of lattice dislocations, grain-boundary dislocations, vacancies and grain rotations
occurring in mechanically loaded nanocrystalline materials. At present, there are many theoretical
models of the abnormal Hall–Petch effect based on different deformation mechanisms and claiming
agreement with the corresponding experimental data from nanocrystalline materials. The following
plastic deformation mechanisms have been mentioned acting individually or in competition: (1)
grain-boundary sliding, (2) grain-boundary diffusional creep, (3) triple junction diffusional creep,
(4) rotational deformation (occurring through motion of grain-boundary disclinations) and (5) lattice
dislocations. The experimental data is usually not precise enough to allow one to select a theoretical
concept from a variety of theoretical models describing the same experimental data using various
mechanisms. In this context, in next sections of this review article, we will pay special attention to
theoretical models of plastic deformation mechanisms in nanocrystalline materials, that can account
for some additional microstructural results (either experimental or computational) and provide math-
ematical results rather than qualitative concepts. Needless to say, the subject is still a matter of some
controversy.

4. Models using lattice dislocations

The most obvious idea is to use conventional lattice dislocation slip model for nanocrystalline
materials, but taking into account the influence of smaller grain sizes and high-density ensembles
of grain boundaries on the formation of lattice dislocation pile-ups in grain interiors. Thus, this treat-
ment extends the classical derivation but assumes that there are very few dislocations available in any
one grain.
For this purpose it is instructive to start with a brief discussion of the models describing the clas-
sical Hall–Petch relationship (Eq. (1)) in coarse-grained polycrystals. Most of these models use the
concept of dislocation pile-ups (see review by Li and Chou [69]). In deriving the Hall–Petch relation,
grain boundaries here are considered as barriers to dislocation motion [70,71], causing stresses to con-
centrate and activating dislocation sources in the neighboring grains, thus initiating the slip from grain
to grain. In other type of models, though mentioned less often [72,73], the grain boundaries are re-
garded as dislocation barriers limiting the mean free path of the dislocations, thereby increasing strain
hardening and resulting in a Hall–Petch type relation. Several variations of these concepts are possible.
It is also possible that several dislocation processes could compete or reinforce the deformation
process.
Pande and Masumura [74] were the first to extend mathematically the classical derivation of Hall–
Petch relation to nanocrystalline materials. They assumed that the classical Hall–Petch dislocation
pile-up model is still dominant with the sole exception that the analysis must take into account the
fact that in nanocrystalline materials with small grain sizes, the number of dislocations in a pile-up
within a grain cannot be very large. In the limit at still smaller grain sizes, this mechanism should
cease when the grains are so small that there are only two dislocations in the pile-up. Mathematically,
the model utilizes the fact that the length of the pile-up is no longer proportional to the number of
dislocations in the pile-up if the pile-up is not large.
Then, Pande and Masumura [74], by considering the conventional Hall–Petch model, showed that a
dislocation theory for the Hall–Petch effect does not give a linear dependence of s on d1/2 when grain
sizes are in the nanometer range. When the number of dislocations in the pile-up falls to one, no fur-
ther increase in the yield stress is possible by this mechanism and it saturates. As mentioned before, if
the number of dislocations n in a pile-up is not too large, the length of the pile-up L is not linear in n.
Chou [75] gives the relation between L and n as:
  1=3 !
A  2n 
Lffi 4 n þ m  1  2i1 ; ð2Þ
2s  3 

where i1 = 1.85575 and mb is the Burgers vector of the lead dislocation in the pile-up. (The lead dis-
location could be in the grain-boundary itself and, hence, may have a Burgers vector different from
the rest of the dislocations.)
694 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

Pande and Masumura [74] give an improved expression,


" #
A 1=2 i1 þ e
Lffi 2ðn þ m  1Þ  1=3 i1 ; ð3Þ
2s 12 ðn þ m  1Þ1=6

where e is a small correction term (e  1) and can be neglected. Then, following the classical analysis,
but using Eq. (3), they find that for small grain sizes there are additional terms to the Hall–Petch
relation,

s ¼ k1=2 þ c1 ðk1=2 Þ5=3 þ c2 ðk1=2 Þ7=3 ; ð4Þ


* *
where s = s/[m/s ], c1 = 0.6881, c2 = 0.1339 and l = Lms /2A. Eq. (4) is expected to be correct for all
grain sizes, as long as dislocation pile-up mechanism is operating. This model thus recovers the clas-
sical Hall–Petch at large grain sizes but for smaller grain sizes the s levels off. This mechanism there-
fore cannot explain a drop in s. If, on the other hand, the yield stress is source limited, s  Gb/d, i.e., the
yield stress should rise as d1. Thus, from these arguments, at smaller grain sizes, either the yield
stress should rises faster than d1/2 or it should saturate, but it should not decrease. The Hall–Petch
plot using Eq. (4) is given in Fig. 4 showing the leveling, but not the ‘inverse’ Hall–Petch curve.
Several researchers [76–78] have developed models similar to that of Pande and Masumura [74].
On the other hand, Malygin [79] has suggested that dislocations are absorbed in grain boundaries,
the effect being larger in nanocrystalline materials. The assumption is that grain boundaries act pre-
dominantly as sinks for dislocations (just the opposite to what was proposed by Li [80], who consid-
ered grain boundaries as sources for dislocation generation). Malygin’s model is attractive as a
dislocation mechanism and should be considered further. However, we point out two problems with
the model. First, it is doubtful if the dislocations play the same role whether the grains are large or
small. It is more likely that dislocations in ultrafine grains, if present at all, are confined to grain
boundaries [81]. Second, in Malygin’s model [79], the stress calculated is a work-hardened flow stress
rather than a yield stress and, as with Li [80], it merely assumes a relation between dislocation density
and yield stress, instead of proving it.
Lu and Sui [82] assume an enhancement of lattice dislocation penetration through the grain bound-
aries and the corresponding softening of nanocrystalline materials. Scattergood and Koch [83] assume
that the yield stress of fine-grained materials is controlled by the intersection of mobile lattice
dislocations with the dislocation networks at grain boundaries. Zaichenko and Glezer [84] have pro-
posed rotational defects (disclinations), formed at triple junctions, as sinks and sources of the lattice
dislocations moving in grain interiors and causing the plastic flow in nanocrystalline materials (see
also Ref. [85]).

Fig. 4. Plot of Hall–Petch dislocation model for nanocrystalline materials [74].


C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 695

All these models assume the existence of lattice dislocations, which play the same role in nanograin
interiors as with conventional coarse grains. This may not be energetically favorable [86,87] in nano-
crystalline materials. Transmission electron microscopy experiments [6,88] show reduced density or
even absence of dislocations in nanocrystalline materials. However, one cannot rule out that the mod-
els based on lattice dislocation slip may explain to some extent the deformation behavior of nanocrys-
talline materials with grain size in the intermediate range (d  30–100 nm).

5. Role of Coble creep as a competing mechanism

As we have seen, at sufficiently small grain sizes, the Hall–Petch model based upon the lattice dis-
locations may not be operative. Clearly, there is a need to consider additional mechanisms for very
small grain sizes (less than 10 nm). In this section, we consider the deformation mechanisms associ-
ated with enhanced diffusion along grain boundaries in nanocrystalline materials. Chokshi et al. [89]
were the first to suggest grain-boundary diffusional creep (Coble creep) as the dominant deformation
mode in nanocrystalline solids. In this region, they have proposed that even at room temperature Co-
ble creep may be operative in order to explain their results. Certainly, there is a qualitative order of
magnitude agreement with this mechanism and the trend is correct, however, the functional depen-
dence of s on d given by Choksi et al. [89] is incorrect as pointed out by Neih and Wadsworth [90].
Conventional Coble creep demands that s  d3/[d1/2]6, i.e., the s vs. d1/2 curve falls very steeply
as d1/2 increases. This is not found experimentally [76]. Chokshi et al. [89] showed that their data
fit better the relation,
s ¼ b  K 0  d1=2 ð5Þ
0 1/2 3
with b = 937 MPa and K = 0.027 MPa m instead of s  d . Eq. (5) cannot be related simply to any
known mechanism.
Even if the Coble creep argument is valid for grain sizes d < 20 nm, we still have to explain the
behavior in the 20–200 nm range. This is evidently the transition regime between the Hall–Petch
and Coble creep-like behavior. The transition regime can only be effectively described when a distri-
bution of grain size is taken into account, as has been done by Masumura et al. [53]. They provided a
unified model and developed an analytical expression for s as a function of the inverse square root of d
in a simple and approximate manner that could be compared with experimental data over the whole
range of grain sizes from very large to very small. As far as we know this is the only model that gives
an analytical expression for the Hall–Petch relation for the whole range of grain sizes. This model,
based on the idea of competition between lattice dislocation slip and grain-boundary diffusional creep
(Coble creep), is summarized below.
In this model, it is assumed that polycrystals with a relatively large average grain size obey the
classical Hall–Petch relation Eq. (1), and for very small grain sizes, it is assumed that Coble creep is
active and that the s vs. d relationship is given by,
sc ¼ A=d þ Bd3 ; ð6Þ
where B is both temperature and strain-rate dependent. The additional term A/d (the threshold term)
was added by Masumura et al. [53] on an ad hoc basis (see, however, the later part of this section for
justification on physical grounds). Contribution from this term can be significant if d is in the nanome-
ter range. For intermediate grain sizes, both mechanisms might be active simultaneously especially if
the specimen has a grain size distribution as is usually the case. The model is illustrated graphically in
Fig. 5.
The presence of a distribution of the grain sizes in a polycrystal is taken into consideration by using
an analysis similar to Kurzydlowski [91]. The volumes of the grains are assumed to be log-normally
distributed and are given by,
" #
1 ðln v  mln v Þ
f ðv Þ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp ; ð7Þ
v 2psln v s2ln v
where mln v and sln v are the mean value and standard deviation of ln v, respectively, and where the mv
is the mean volume of all the grains,
696 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

Fig. 5. Graphically illustrated model for grain size-dependence of material strength.

Z 1
mv ¼ v f ðv Þdv ¼ exp½mln v þ ðsln v Þ2 ; ð8Þ
0

mv can also be written as mv ¼ g d 3 where d  is mean grain size and with g being a geometrical shape
factor considered for this analysis to be equal to 1. Following Masumura et al. [53], other authors
have also used this type of averaging, e.g. see Morita et al. [92], Zhu et al. [93] and Phaniraj et al.
[94].
Finally, it is assumed that a grain size d* exists at which value the classical Hall–Petch mechanism
switches to the Coble creep mechanism, shp = sc at d = d*. Using Eqs. (1) and (6), we have,

kðd Þ1=2 ¼ A=d þ Bðd Þ3 ð9Þ

from which d* can be determined.


So far, the modeling is fairly simple and straightforward. It is the averaging requirement over a dis-
tribution of grain sizes that complicates the mathematics. However, this averaging is necessary other-
wise Hall–Petch plot will switch abruptly from one mechanism to the next.
Then the yield stress after averaging is given as,

hs  s0 i ¼ F hp þ F c ; ð10Þ

where
Z 1
1
F hp ¼ shp dv ð11Þ
mv v
and
Z 1
1
Fc ¼ sc dv : ð12Þ
mv v
We define,

d
n¼ and r ¼ sln v ð13Þ
d
C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 697

along with,
 
r2
Kðn; r; aÞ ¼ n3a exp aða þ 1Þ ð14Þ
2
and,
( 1=2    )
1 v r2 2
Uðv ; r; aÞ ¼ erf ln  ð2a þ 1Þ ; ð15Þ
2r2 mv 2

where erf {} is the error function. Further defining additional normalized variables,

F c1 ¼ Kð2d Þ1 fc1 ;
fc1 ¼ Kðn; r; 1=3Þ½Uðn; r; 1=3Þ þ 1 ð16Þ
and

ðd Þ3
F c2 ¼ B fc2 ; ð17Þ
2

fc2 ¼ Kðn; r; 1Þ½Uðn; r; 1Þ þ 1 ð18Þ


and

A=d

ð19Þ
Bðd Þ3

we have a normalized form of the yield stress, sn, as a function of the scaled grain size n, grain size
parameter r and p,
2 pfc1 þ fc2
sn ¼
hs  s0 i ¼ fho : ð20Þ
kðd Þ1=2 1þp

The parameter p is the ratio of Coble threshold stress to conventional stress evaluated at d* where
the transition from Coble creep to Hall–Petch strengthening occurs. For each p and r universal curve is
obtained with the form and shape of the curve similar to experimental data. Eq. (20) is the key result
of their analysis. In Fig. 6, the yield stress data for NiP of McMahon and Erb [95], after normalization, is
compared with Eq. (20). The threshold stress as compared to the Coble creep stress is small in this
material. The value for the optimum grain size d* = 5.5 nm, as determined from this analysis, is in
agreement with the original hardness data where the hardness (or stress) begins to decrease with
decreasing grain size at a grain size of 5–6 nm.
Thus this model uses conventional Hall–Petch strengthening for larger grains and Coble creep with
a threshold stress for smaller grains. In a material with a distribution of grain sizes, a fraction of grains
deforms by a lattice dislocation slip process and the rest by vacancy transport. As the average grain
size decreases, the fraction deforming by slip decreases and the overall response changes from
strengthening to softening. The exact form of the yield stress against grain size curve depends on
the relative values of Hall–Petch slope k, the conventional Coble constant B, the threshold constant
A and the width of the grain size distribution b. For ease of analysis, it was assumed in this model that
one mechanism (i.e., classical Hall–Petch dislocation mechanism) operates on large grains and Coble
creep on the smaller grains. In practice, in all grains both mechanisms will compete and the faster one
will dominate. This is how strain-rate sensitivity will become a factor.
The model of Masumura et al. [53] is supported by results [96] of computer modeling of plastic
deformation processes in nanocrystalline materials, indicating that the essential contribution of Coble
creep to these processes cannot be discounted. Also, it should be noted that Masumura et al. [53] have
suggested a new general approach, backed by detailed mathematical analysis, to describe the mechan-
ical characteristics of deformed nanocrystalline materials, which takes into account a distribution in
grain size and suggests simultaneous action of different deformation mechanisms in a mechanically
loaded sample. Later researchers in several theoretical works have exploited this approach as a
698 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

Fig. 6. Graph showing comparison of NiP experimental data [95] with NRL model of normalized yield stress as a function of
(normalized average grain size)1/2.

method for description of competing deformation modes, which is more accurate than the ‘‘rule-of-
mixture” approach.
Fedorov et al. [97] suggested a theoretical model describing the yield stress dependence on grain
size in fine-grained materials, based upon competition between conventional dislocation slip, grain-
boundary diffusional creep and triple junction diffusional creep. As in Masumura et al. [53], the model
[97] takes into account a distribution in grain size.
It has been found experimentally [51,52] that the s(d) relationship in nanocrystalline materials of-
ten shows two different behaviors, depending on their processing. Heat-treated materials exhibit ‘in-
verse’ Hall–Petch behavior, while the yield stress of as-prepared materials slightly increases or
saturates at grain size d 6 10 nm showing little or no ‘inverse’ Hall–Petch behavior. Gutkin et al.
[98] postulate that this difference in the deformation behavior between heat-treated and as-prepared
nanocrystalline materials may be related to the difference between their defect structures, leading dif-
ferent deformation modes occurring due to grain refinement, the contribution of grain-boundary slid-
ing being higher in as-prepared materials commonly characterized by high density of lattice and
grain-boundary dislocations. They showed that,
s ¼ k1 þ k2 =d; ð21Þ
where k1 and k2 are constants. This could explain the threshold term that was used by Masumura et al.
[53].
It is however by no means certain that inverse Hall–Petch is observed only in heat-treated nanom-
aterials and not in as-prepared materials. There are several examples in the literature that clearly
show inverse Hall–Petch for as-prepared materials also. (For example, Erb et al. [99] give several
examples of inverse Hall–Petch in as-prepared nanocrystalline electrodeposits.)

6. A generalized expression for yield stress

From the above discussion a generalized form of yield stress applicable to any polycrystal is ob-
tained as,
B0 3
s ¼ s0 þ kd1=2 þ k1 þ þ Bd ; ð22Þ
d
where k1 and B0 are as yet undetermined constants. Several points regarding this equation should be
noted. Firstly, the first two terms on the left hand side are to be used for grains larger that a given
C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 699

critical size d*. For grains smaller than d* the last three terms on the right hand side should be appli-
cable. As before d* is obtained by the following equation,
B0
s0 þ kðd Þ1=2 ¼ k1 þ 3
þ Bðd Þ : ð23Þ
d
Secondly, not all the three terms on the right hand side will be significant for every situation. If Co-
ble creep is dominant, the last two terms will be dominant; whereas if grain-boundary dislocations are
active, the first two terms may be more significant. Eq. (22) differs from the corresponding result used
in Masumura et al. [53] in two respects. First, it has an additional term k1 and second, the B0/d term,
which was introduced in an ad hoc fashion in Masumura et al. [53], is identified with grain-boundary
dislocation reactions. Finally, this equation is true provided all the grains are of the same size. In a real
polycrystal, we will have a distribution of grain sizes. Hence the averaging procedure given by Masum-
ura et al. [53] should be used.
We define a volume average of the system of Eq. (23) as,
Z 1 Z d  
1 1=2 1 B0 3
ðsÞav e  so ¼ kd v f ðv Þdv þ k1 þ þ Bd v f ðv Þdv ; ð24Þ
mv d mv 0 d
where v is the grain volume, f(v) is the grain volume distribution function and is assumed to be log-
normal. The grain volume average mv is defined by,
Z 1
mv ¼ v f ðv Þdv  cd3 ; ð25Þ
0

where d is the mean grain size and c is a geometrical factor of the order of one. All the integrals in Eqs.
(24) and (25) can be integrated exactly in terms of error functions [53].
For practical purposes it might be convenient to have an approximate expression, but simpler than
developed in Masumura et al. [53], for the yield stress as a function of grain size. Towards that goal, we
proceed as follows. Eq. (24) can be re-written as,
Z 1 Z d  
1 1=2 1 1=2 B0 3
ðsÞav e  so ¼ kd v f ðv Þdv  kd  k1   Bd v f ðv Þdv
mv 0 mv 0 d
or
ðsÞav e  so ¼ I1  I2 ; ð26Þ
where I1 is the first integral and I2 is the second integral.
When using a lognormal distribution, the first integral can be evaluated and is equal to,

k exp 5r2 =72 K
I1 ¼ ¼ ; ð27Þ

ðdÞ 1=2 
ðdÞ1=2
where r is the standard deviation in ‘nðv Þ. The second integral can be expressed in terms of error func-
tions as mentioned before, but we find that for the range of grain sizes from 0 to d*, a good approxi-
mation is given by,
2 ( !)2 3
M1 
d
I2 ¼ exp 4M 2 ‘n 5; ð28Þ
 3
ðdÞ d

where M1 and M2 strictly are functions of d*. But we find that M2  1 and M1 can be obtained from
1=2
using the fact that the curve of ðsÞav e vs: d should peak around d*. This gives M1 as,
K 5=2
M1 ¼ ðd Þ : ð29Þ
6
Thus finally,
2 ( !)2 3
 3 
K K d d
ðsÞav e ¼ so þ   exp 4M 2 ‘n 5: ð30Þ
 1=2 6ðd Þ1=2 d
ðdÞ d
700 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

We propose this approximate, semi-empirical equation as a convenient and simple equation to de-
scribe the whole range of grain sizes. The first two terms, of course, give the classical Hall–Petch result.
The last term is responsible for the ‘inverse’ Hall–Petch effect. M2 is about 1, K (the Hall–Petch slope)
can be obtained experimentally and d* being given by Eq. (24).
It is interesting to note that in this scheme, the mechanisms responsible for the ‘inverse’ Hall–Petch
effect can, as a first approximation, be represented by d* alone. In principle, d* can be obtained in terms
 from the peak, d , of the experimental curve of s vs. ðdÞ
of d  1=2 . If the experimental grain size distri-
p
* *
bution is very narrow, dp  d and dp or d is given by Eq. (24). If, however, the grain size distribution is
not narrow, the actual peak dp of s vs. ðdÞ  1=2 curve may not coincide with d*. By using various grain
size distributions and using the averaging procedure given by Masumura et al. [53] it is easy to show
that dp P d*. In this case, Eq. (30) is only approximately true.
Taking these factors into account, we write Eq. (30) as,

snorm ¼ n  M1 n6 exp½4M2 f‘nðnÞg2 ; ð31Þ

where
!1=2
ðsÞav e  so 
d
snorm ¼ and n ¼ ; ð32Þ
Kðdp Þ1=2 dp

and because of the reasons stated before, M1  1/6 and M2  1. We plot Eq. (32) using these values of
M1 and M2. The right side of the curve will be affected if somewhat different values of M1 and M2 are
used. Fig. 7 illustrates this model by plotting normalized yield stress as a function of normalized grain
size.
It is seen that Fig. 7 and Eq. (32) are able to account for both the conventional and ‘inverse’ Hall–
Petch regions. We should point out that Eq. (32) is semi-empirical, and is a rough approximation to the
actual function obtained by the detailed averaging procedure. Eq. (32) uses just two constants, M1 and
M2 whose values are approximately given and uses the normalization of the average grain size with
the peak of the curve, which is easily obtained from the experimental data.
As in Masumura et al. [53], Conrad and Narayan [100] assume that at larger grain sizes the hard-
ness increases with decreasing grain size according to classical Hall–Petch relation and there is a
change in mechanism below some grain size. They propose a thermally-activated grain-boundary
shearing model. The grain-boundary shear was associated with many independent, atomic shear
events. Employing the concept of thermally-activated shear, they suggest that the macroscopic shear
rate is produced by independent, atomic shear events at the grain boundaries.
Although many computer simulations (see Section 8 below) suggest the validity of Coble creep as a
small grain size deformation mechanism, there is, at present, no strong experimental evidence to com-
pletely establish their validity. For alternative views, see Kottada and Chokshi [101] and Li et al.
[102,103].

7. Relationship between hardness and yield strength in metals and alloys

Many of the early studies on the mechanical properties on nanomaterials showed hardness data
rather than tensile properties (e.g. yield strength and tensile strength). This was mainly due to the
size limitations of materials available for testing in these early studies. Many experimental and
modeling studies have assumed this relationship. The yield strength-hardness relationship was used
for the experimental data for NiP shown in Fig. 6 in this review. While many authors do not explic-
itly state what relationship they have used, all indications are that the factor of one third was used
to arrive at the strength data. Whenever possible we tried to scale the data to some given value so
as the get a functional dependence. For example, we used the yield stress for large grains as a scale
in Fig. 1.
Strictly speaking there is no simple relationship, linear or nonlinear, between hardness and yield
strength in metals and alloys. Hardness correlates better with tensile strength than yield strength,
the indenter deforms sufficient material until the flow stress and area of contact cannot sustain the
C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 701

Fig. 7. Semi-empirical formulation for yield stress (Eq. (32)).

indenter load. Unless one is able to model the deformation under the indenter, and have detailed
knowledge of the flow curve (which means one already knows the yield strength from a separate
test), one does not know the yield strength except on a very qualitative basis. This is indeed the
case then the analysis of strength data as presented here and in the literature, in general, becomes
questionable in view of recent results which clearly show that for nanocrystalline electrodeposits
(e.g. the type of materials analyzed in Fig. 6) this relationship does not apply (Brooks et al.
[104]). In this analysis, it was shown that for about 200 nanocrystalline Ni- and Co-based test
materials, such a relationship overestimates the yield strength significantly. For all samples ana-
lyzed, yield strength was always much less than 1/3 of the hardness and, more importantly, there
was no single conversion factor that could describe the relationship for all 200 samples that were
analyzed. The latter then creates a serious problem in any modeling work as variable conversion
factors for different materials make a generalized analysis impossible. In other words, any conclu-
sions made in modeling work can only really apply to hardness and not the yield strength. This,
however, creates another problem, as hardness values cannot easily be correlated with deformation
processes in the materials. There are at least two other studies, Ebrahimi et al. [105] and Dalla Tor-
re et al. [106], which showed significant differences in the hardness to strength ratio for electrode-
posited Ni.
On the other hand, Pavlina and Van Tyne [107] have compiled hardness as well as yield and tensile
strength values for over 150 nonaustenitic, hypoeutectoid steels having a wide range of compositions
and a variety of microstructures. The microstructures include ferrite, pearlite, martensite, bainite, and
complex multiphase structures. The yield strength of the steels ranged from approximately 300 MPa
to over 1700 MPa. Tensile strength varied over the range of 450–2350 MPa. Regression analysis was
used to determine the correlation of the yield strength and tensile strength to the diamond pyramid
hardness values for these steels. Both the yield strength and tensile strength of the steels exhibited a
linear correlation with the hardness over the entire range of strength values. Only a weak effect of
702 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

strain-hardening potential on the hardness-yield strength relationship was observed. So, for large
grained materials, there is some justification in the use of the conventional conversion relations. All
this points to the fact that one has to be very careful in comparing any nanocrystalline experimental
data with theory, and also much of the agreement reported between various models and data should
be considered as qualitative and tentative.

8. Concluding remarks

Theoretical studies of plastic deformation processes in nanocrystalline materials were initially


dominated by the ‘‘rule-of-mixture” approach. In the framework of this approach, grain interiors
and grain boundaries (and, sometimes, triple junctions of grain boundaries) are considered as con-
stituent phases characterized by different values of the yield stress [11,55,62–68], the yield stress
of a nanocrystalline material being given as some weighted sum of the yield stresses of grain
interiors and grain boundaries. Deformation mechanisms, operating in nanocrystalline materials
at a more refined level, are described by lattice dislocations and creep invoking atomic level phe-
nomena. Here, even though it is commonly assumed that one deformation mechanism dominates
in practice, several mechanisms may simultaneously contribute to plastic flow. A grain size distri-
bution must be taken into account in calculation of the yield stress as the averaged (over grain
sizes) value with different deformation mechanisms acting in grains with sizes being in different
ranges.
The importance of triple junctions and disclination defects in the deformation of nanocrystalline
materials, in particular when it comes to the inverse Hall–Petch relationship, was addressed in the
work by Palumbo et al. [108]. Disinclinations are definitely effective in a description of the rotational
deformation mode in nanocrystalline solids [109–114]. Motion of grain-boundary disclinations in
plastically deformed solids is commonly treated as that associated with absorption of lattice disloca-
tions (that are generated and move in grains under the action of mechanical load) by grain boundaries
[112,113,115]. This micromechanism, according to Ref. [115], is responsible for experimentally ob-
served grain rotations in polycrystalline materials during superplastic deformation. Recently, a theo-
retical model [110,111,116] has been suggested to describe the rotational deformation in
nanocrystalline solids as the processes occurring mostly via the motion of grain-boundary disclina-
tions and their dipoles, associated with the emission of dislocation pairs from grain boundaries into
the adjacent grain interiors. It is an example of interacting deformation modes in which the rotational
deformation and lattice dislocation slip support each other.
Wolf et al. [117] have reviewed the results of recent molecular-dynamics simulations of the struc-
ture and deformation behavior of nanocrystalline materials, i.e., polycrystalline materials with a grain
size of typically less than about 100 nm. The scale of their simulations is sufficient and sophisticated
enough to cover almost the entire range of grain sizes over which transition from a dislocation-based
deformation mechanism to one involving grain-boundary processes takes place as suggested by
Masumura et al. [53]. Their simulations provide novel insights into the intricate interplay between
the dislocation and grain-boundary processes responsible for ‘inverse’ Hall–Petch crossover. However,
in spite of these agreements, these simulations are inherently limited by extremely high deformation
rates and short time scale inherent in any molecular-dynamics simulations.
Some other features not considered in this review are superplasticity in solids at relatively high
strain rates and low temperatures [21–29]. The dominant mode of superplasticity in nanocrystalline
materials is viewed to be grain-boundary sliding [27,29,118,119]. Other works worthy of providing in-
sights to the micro mechanism of Hall–Petch law for nanocrystalline materials are as follows. Van
Swygenhoven et al. [120,121] showed by molecular-dynamics simulations that grain-boundary
sliding is the primary deformation mechanism in nanocrystalline materials. A discussion of grain-
boundary sliding was introduced by Meyers et al. [122]. Recent observations by Ma and coworkers
[123–126] suggest that nanosized grains rotate during plastic deformation and can coalesce along
directions of shear, creating larger paths for dislocation movement between them, providing a path
for more extended dislocation motion. One way by which grains can rotate is by disclination motion.
Murayama et al. [127] first observed disclinations in a mechanically milled Fe sample and suggested
C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 703

Fig. 8. Bimodal grain size distribution concept.

that the generation of partial disclination defects provides an alternative mechanism to grain-bound-
ary sliding, which could result in rotation.
Most nanocrystalline metals while exhibiting enhanced yield strength often have a reduction in
ductile strength. An attractive processing route for enhancing the ductility of high-strength nano-
crystalline metals and alloys is to develop a bimodal grain size distribution, in which, supposedly,
the finer grains provide strength, and the coarser grains maintain or even enhance ductility. This
is schematically shown in Fig. 8. In practice, Wang et al. [128] first reported a successful application
of this concept in nanostructured, ultrafine-grained Cu. They rolled the Cu to 93% at liquid nitrogen
temperatures and then annealed it at low temperatures, up to 200 °C. The annealing treatment opti-
mized strength and ductility producing a mixture of nano-scale, ultrafine grains (80–200 nm) along
with about 25% volume fraction of coarser grains (1–3 lm). The coarser grains were probably the
result of secondary recrystallization (see also Ma [129]). Since then several researchers have re-
ported similar results. For a more recent work on microstructural evolution and mechanical charac-
teristics in a nanocrystalline material (Ni instead of Cu) with a bimodal grain size distribution see
the work of Prasad et al. [130].

Acknowledgements

This work was supported, in part, by the Office of Naval Research. We thank Dr. Robert A. Masum-
ura for his valuable analysis and guidance.

References

[1] Roco MC, Williams RS, Alivisatos P, editors. Nanotechnology research directions. Dordrecht: Kluwer; 2000.
[2] Chow G-M, Ovid’ko IA, Tsakalakos T, editors. Nanostructured films and coatings. NATO science series. Dordrecht: Kluwer;
2000.
704 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

[3] Gleiter H. Acta Mater 2000;48:1.


[4] Komarneni S, Vaja RA, Lu GQ, Matsushita J-I, Parker JC, editors. Nanophase and nanocomposite materials IV. MRS
symposium proceedings, vol. 703. Warrendale, PA: MRS; 2003.
[5] Berndt CC, Fischer T, Ovid’ko IA, Skandan G, Tsakalakos T, editors. Nanomaterials for structural applications. MRS
symposium proceedings, vol. 740. Warrendale, PA: MRS; 2003.
[6] Siegel RW, Fougere GE. Nanostruct Mater 1995;6:205.
[7] Hahn H, Padmanabhan KA. Nanostruct Mater 1995;6:191.
[8] Farkas D, Kung H, Mayo M, VanSwygenhoven J, Weertman J. Structure and mechanical properties of nanophase materials
– theory and computer simulations vs. experiment. Warrendale, PA: MRS; 2001.
[9] Koch CC, Morris DG, Lu K, Inoue A. MRS Bull 1999;24:54.
[10] Mohamed FA, Li Y. Mater Sci Eng A 2001;298:1.
[11] Gutkin MYu, Ovid’ko IA, Pande CS. Rev Adv Mater Sci 2001;2:80.
[12] Padmanabhan KA. Mater Sci Eng A 2001;304–306:200.
[13] Verpek S, Argon AS. J Vac Sci Technol 2002;20:650.
[14] Niederhof A, Bolom T, Nesadek P, Moto K, Eggs C, Patil DS, et al. Surf Coat Technol 2001;146–147:183.
[15] Champion Y, Langlois C, Guerin-Mailly S, Langlois P, Bonnentien J-L, Hytch M. Science 2003;300:310.
[16] Wang Y, Chen M, Zhou F, Ma E. Nature 2002;419:912.
[17] He G, Eckert J, Loeser W, Schultz L. Nat Mater 2003;2:33.
[18] Ma E. Nat Mater 2003;2:7.
[19] Mayo MJ. Nanostruct Mater 1997;9:717.
[20] Mayo MJ. In: Chow G-M, Noskova NI, editors. Nanostructured materials: science and technology. Dordrecht: Kluwer;
1998. p. 361.
[21] Mishra RS, Valiev RZ, McFadden SX, Mukherjee AK. Mater Sci Eng A 1998;252:174.
[22] McFadden SX, Misra RS, Valiev RZ, Zhilyaev AP, Mukherjee AK. Nature 1999;398:684.
[23] Islamgaliev RK, Valiev RZ, Mishra RS, Mukherjee AK. Mater Sci Eng A 2001;304–306:206.
[24] Mishra RS, Valiev RZ, McFadden SX, Islamgaliev RK, Mukherjee AK. Philos Mag A 2001;81:37.
[25] Zhan G-D, Kuntz JD, Wan J, Mukherjee AK. In: Berndt CC, Fisher T, Ovid’ko IA, Skandan G, Tsakalakos T, editors.
Nanomaterials for structural applications. MRS symposium proceedings, vol. 740. Warrendale, PA: MRS; 2003. p.
49–54.
[26] Valiev RZ, Song C, McFadden SX, Mukherjee AK, Mishra RS. Philos Mag A 2001;81:25.
[27] Mukherjee AK. Mater Sci Eng A 2002;322:1.
[28] Valiev RZ, Alexandrov IV, Zhu YT, Lowe TC. J Mater Res 2002;17:5.
[29] Mukherjee AK. In: Mishra RS, Earthman JC, Raj SV, editors. Creep deformation: fundamentals and
applications. Warrendale, PA: TMS; 2002. p. 3–19.
[30] Horvath J, Birringer R, Gleiter H. Solid State Commun 1987;62:391.
[31] Schaefer H-E, Wurschum R, Gessmann T, Stockl G, Scharwaechter P, Frank W, et al. Nanostruct Mater 1995;6:869.
[32] Kolobov YuR, Grabovetskaya GP, Ratochka IV, Ivanov KV. Nanostruct Mater 1999;12:1127.
[33] Kolobov YuR, Grabovetskaya GP, Ivanov KV, Valiev RZ, Lowe TC. In: Lowe TC, Valiev RZ, editors. Investigations and
applications of severe plastic deformation. NATO science series. Dordrecht: Kluwer; 2000. p. 261.
[34] Gleiter H. Nanocrystalline materials. Prog Mater Sci 1989;33:223–315.
[35] Birringer R. Mater Sci Eng A 1989;117:33.
[36] Gleiter H. Nanostruct Mater 1992;1:1.
[37] Suryanarayana C. Nanocrystalline materials. Int Mater Res 1995;40:41–64.
[38] Lu K. Mater Sci Eng 1996;R16:161–221.
[39] Weertman JR, Farkas D, Hemker K, Kung H, Mayo M, Mitra R. MRS Bull 1999;24:44.
[40] Suryanarayana C, Koch CC. Hyperfine Interact 2000;130:5–44.
[41] Valiev RZ, Islamgaliev RK, Alexandrov IV. Prog Mater Sci 2000;45:103–89.
[42] Furukawa M, Horita Z, Nemoto M, Langdon TG. J Mater Sci 2001;36:2835–43.
[43] Mohamed FA, Li Y. Mater Sci Eng A 2001;298:1–15.
[44] Kumar KS, Van Swygenhove H, Suresh S. Acta Mater 2003;51:5743–74.
[45] Veprek S, Veprek-Heijiman MGJ, Karvankova P, Prochazka J. Thin Solid Films 2005;476:1–29.
[46] Wolf D, Yamakov V, Phillipot SR, Mukherjee A, Gleiter H. Acta Mater 2005;53:1.
[47] Weertman J. In: Koch CC, editor. Nanostructured materials. Norwich, NY: Noyes Publications; 2002. p. 393–417.
[48] Meyers MA, Mishra A, Benson DJ. Prog Mater Sci 2006;51:427–556.
[49] Ovid’ko IA, Pande CS, Masumura RA. Grain boundaries in nanomaterials. In: Gogotsi YG, editor. Handbook on
nanomaterials. CRC: Florida; 2005. p. 531–52 [chapter 18].
[50] Gutkin MYu, Ovidko IA, Pande CS. Physical mechanisms of plastic flow in nanocrystalline materials. In: Nalwa HS, editor.
Nanoclusters and nanocrystals. California: American Science Publications; 2003. p. 225–52.
[51] Weertman JR, Sanders PG. Solid State Phenom 1994;35–36:249.
[52] Volpp T, Göring E, Kuschke W-M, Arzt E. Nanostruct Mater 1997;8:855.
[53] Masumura RA, Hazzledine PM, Pande CS. Acta Mater 1998;46:4527.
[54] Nieman GW, Weertman JR, Siegel RW. J Mater Res 1991;6:1012.
[55] Carsley JE, Ning J, Milligan WW, Hackney SA, Aifantis EC. Nanostruct Mater 1995;5:441.
[56] Witney AB, Sanders PG, Weertman JR, Eastman JA. Scripta Metall Mater 1995;33:2025.
[57] Andrievskii RA, Kalinnikov GV, Kobelev NP, Soifer YaM, Shtansky DV. Phys Solid State 1997;39:1661.
[58] Andrievskii RA. In: Chow G-M, Noskova NI, editors. Nanostructured materials: science and
technology. Dordrecht: Kluwer; 1998. p. 59–65.
[59] Wei Q, Jia D, Ramesh KT, Ma E. Appl Phys Lett 2002;81:1240.
[60] Kumar KS, Suresh S, Chisholm MF, Norton JA, Wang P. Acta Mater 2003;51:387.
[61] Hughes GD, Smith SD, Pande CS, Johnson HR, Armstrong RW. Scripta Mater 1986;20:93–7.
C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706 705

[62] Gryaznov VG, Gutkin MYu, Romanov AE, Trusov LI. J Mater Sci 1993;28:4359.
[63] Gutkin MYu, Ovid’ko IA. Nanostruct Mater 1993;2:631.
[64] Kim HS. Scripta Mater 1998;39:1057.
[65] Konstantinidis DA, Aifantis EC. Nanostruct Mater 1998;10:1111.
[66] Ovid’ko IA. Nanostruct Mater 1997;7:149.
[67] Wang N, Wang Z, Aust KT, Erb U. Acta Metall Mater 1995;43:519.
[68] Kim HS, Estrin Y, Bush MB. Acta Mater 2000;48:493.
[69] Li JCM, Chou YT. Metall Trans 1970;1:1145.
[70] Armstrong RW, Head AK. Acta Met 1965;13:759.
[71] Pande CS, Masumura RA. In: Valluri SR, Taplin DMR, Rama Rao P, Knott JF, editors. Proc of sixth international conference
on fracture; 1984. p. 857–61.
[72] Evans AG, Hirth JP. Scripta Metall Mater 1992;26:1675.
[73] Pande CS, Masumura RA, Armstrong RW. Nanostruct Mater 1993;2:323.
[74] Pande CS, Masumura RA. In: Suryanarayana C, Singh J, Froes FH, editors. Processing and properties of nanocrystalline
materials. Warrendale, PA: TMS; 1996. p. 387.
[75] Chou YT. J Appl Phys 1967;38:2080.
[76] Nazarov AA, Romanov AE, Valiev RZ, Baudelet B. Strength of materials. Japan: JIMIS; 1994. p. 877.
[77] Nazarov AA. Scripta Mater 1996;34:697.
[78] Lian J, Baudelet B, Nazarov AA. Mater Sci Eng A 1993;172:23.
[79] Malygin GA. Phys Solid State 1995;37:1248.
[80] Li JCM. Trans TMS-AIME 1963;227:247.
[81] Evans AG, Hirth JP. Scripta Metall 1992;26:1675.
[82] Lu K, Sui ML. Scripta Metall Mater 1993;28:1465.
[83] Scattergood RO, Koch CC. Scripta Mater 1992;27:1195.
[84] Zaichenko SG, Glezer AM. Phys Solid State 1997;39:1810.
[85] Ovid’ko IA. J Phys D 1994;27:999.
[86] Gryaznov VG, Kaprelov AM, Romanov AE. Scripta Metall 1989;23:1443.
[87] Romanov AE. Nanostruct Mater 1995;6:125.
[88] Siegel RW. In: Trigg GL, editor. Encyclopedia of applied physics, vol. 11. Weinheim: VCH; 1994. p. 1.
[89] Chokshi AH, Rosen A, Karch J, Gleiter H. Scripta Metall 1989;23:1679.
[90] Nieh TG, Wadsworth J. Scripta Metall Mater 1991;25:955.
[91] Kurzydlowski KJ. Scripta Metall Mater 1990;24:879.
[92] Morita T, Mitra R, Weertman JR. Mater Trans 2004;45:502–8.
[93] Zhu B, Asaro RJ, Krysl P, Zhang K, Weertman JR. Acta Mater 2006;54:3307–20.
[94] Phaniraj MP, Prasad MJNV, Chokshi AH. Mater Sci Eng A 2007;463:231–7.
[95] McMahon G, Erb U. Microstruct Sci 1989;17:447.
[96] Yamakov V, Wolf D, Phillpot SR, Gleiter H. Acta Mater 2002;50:61.
[97] Fedorov AA, Gutkin MYu, Ovid’ko IA. Scripta Mater 2002;47:51.
[98] Gutkin MYu, Ovid’ko IA, Pande CS. Yield Stress of nanocrystalline materials: role of grain boundary dislocations, triple
junctions and Coble creep. Philos Mag 2004;84:847–63.
[99] Erb U, Palumbo G, Zugic R, Aust KT. In: Suryanarayana C, Sing J, Froes FH, editors. Processing and properties of
nanocrystalline materials. Warrendale, PA: TMS; 1996. p. 93–110.
[100] Conrad H, Narayan J. Scripta Mater 2000;42:1025–30.
[101] Kottada RS, Chokshi AH. Scripta Mater 2005;53:887–92.
[102] Li YJ, Mueller J, Hoppel HW, Goken M, Blum W. Acta Mater 2007;55:5708–17.
[103] Li YJ, Kapoor R, Wang JT, Blum W. Scripta Mater 2008;58:53–6.
[104] Brooks I, Lin P, Palumbo G, Hibbard GD, Erb U. Mater Sci Eng A 2008;491:412–9.
[105] Ebrahimi F, Bourne GR, Kelly MS, Matthews TE. Nanostruct Mater 1999;11:343–50.
[106] Dalla Torre F, Van Swygenhoven H, Victoria M. Acta Mater 2002;50:3957–70.
[107] Pavlina EJ, Van Tyne CJ. J Mater Eng Perform 2008. ISSN 1544-1024.
[108] Palumbo G, Erb U, Aust K. Scripta Metall Mater 1990;24:2347–50.
[109] Ovid’ko IA. Science 2002;295:2386.
[110] Gutkin MYu, Kolesnikova AL, Ovid’ko IA, Skiba NV. Philos Mag Lett 2002;82:651.
[111] Gutkin MYu, Ovid’ko IA. Physical mechanics of deformed nanostructures, vol. 1. St. Petersburg: Yanus; 2003.
[112] Romanov AE, Vladimirov VI. In: Nabarro FRN, editor. Dislocations in solids, vol. 9. Amsterdam: North-Holland; 1992. p.
191.
[113] Seefeldt M. Rev Adv Mater Sci 2001;2:44.
[114] Gryaznov VG, Trusov LI. Prog Mater Sci 1993;37:289.
[115] Valiev RZ, Langdon TG. Acta Metall 1993;41:949.
[116] Gutkin MYu, Ovid’ko IA, Skiba NV. Mater Sci Eng A 2003;339:73.
[117] Wolf D, Yamakov V, Phillpot SR, Mukherjee A, Gleiter H. Acta Mater 2005;53:1–40.
[118] Gutkin MYu, Ovid’ko IA, Skiba NV. J Phys D 2003;36:L47.
[119] Pande CS, Masumura RA. Mater Sci Eng A 2005;409:125.
[120] Van Swygenhoven H, Derlet PM. Phys Rev B 2001;64:224015.
[121] Van Swygenhoven H, Spaczer M, Caro A. Acta Mater 1999;47:561.
[122] Meyers M, Benson D, Fu HH. In: Chung YW, Dunard D, Liaw P, Olson G, editors. Advanced materials for the 21st
century. Warrandale, PA: TMS; 1999. p. 537.
[123] Jia D, Wang YM, Ramesh KT, Ma E, Zhu YT, Valiev RZ. Appl Phys Lett 2001;79:611–3.
[124] Wang YM, Ma E, Chen MW. Appl Phys Lett 2002;80:2395–7.
[125] Jia D, Ramesh KT, Ma E. Scripta Mater 2000;42:73.
706 C.S. Pande, K.P. Cooper / Progress in Materials Science 54 (2009) 689–706

[126] Jiang X, Jia CL. Appl Phys Lett 1996;69:3902–4.


[127] Murayama M, Howe JM, Hidaka H, Takaki S. Science 2002;295:2433–5.
[128] Wang YM, Chen M, Zhou F, Ma E. Nature 2002;419:912.
[129] Ma E. J Mater Eng Perform 2005;14:1059.
[130] Prasad MJNV, Suwas S, Chokshi AH. Mater Sci Eng A 2009;503:86–91.

You might also like