You are on page 1of 14

Acta Materialia 53 (2005) 4825–4838

www.actamat-journals.com

Transition of deformation mechanisms and its connection to


grain size distribution in nanocrystalline metals
B. Zhu, R.J. Asaro *, P. Krysl, R. Bailey
Department of Structural Engineering, University of California, San Diego, La Jolla, CA 92093, United States

Received 27 April 2005; received in revised form 23 June 2005; accepted 25 June 2005
Available online 1 September 2005

Abstract

A polycrystalline constitutive theory is developed based on the recently proposed model of Asaro, Krysl, and Kad (AKK) for
deformation mechanisms in nanocrystalline metals and the extended aggregate Taylor model of Asaro and Needleman (AN).
The AKK model is particularly concerned with mechanisms including the emission of perfect and partial dislocations, along with
deformation twins from grain boundaries, and grain boundary sliding, and with the grain size, strain rate, and temperature depen-
dent transition between these mechanisms. Thus of specific concern is the effect of grain size distribution as well as the mean grain
size on overall nanocrystalline aggregate response. Grain size distributions are simulated with log-normal distributions and incor-
porated into the AN aggregate model in the form of distributions of the volume fraction of grains with various sizes. Numerical
results for the case of electrodeposited Ni are shown to be in good agreement with available experimental data.
 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Nanostructured metals; Grain boundary emission of dislocations; Grain boundary sliding; Grain size distribution; Mechanical prope-
rties; Nanocrystalline modelling

1. Introduction and larger, size range plastic deformation occurs via


the generation and motion of intragranular slip, i.e., dis-
When the grain size of polycrystalline metals transits location motion. This process is evidently shut off at
through the micrometre down to the nanometre there grain sizes somewhat below a micron. This is readily
are accompanying transitions in the mechanisms of understood by simply noting that the crystallographic
inelastic deformation as well as significant changes in shear stresses required to generate and move dislocation
constitutive properties including, inter alia, levels of segments that exist within the well characterized net-
strength, strain rate sensitivity, and strain hardening. works which evolve during plastic flow are on the order
There is direct experimental evidence for these transi- of s  Gb/‘, where G is the shear modulus, b the magni-
tions (see e.g. [1–4]), theoretical evidence vis-à-vis tude of the Burgers vector, and ‘ is the segment length.
molecular dynamics simulations of nanocrystalline But if dislocations are to be confined to the intragranu-
deformation [5–8], as well as suspicions that arise from lar space – which might be taken as the definition of a
what is known about the mechanisms of plastic defor- grain – ‘ must be less than the grain diameter, d. In fact
mation in crystalline metals. For example, in face cen- simulations of the operation of Frank–Reed sources
tered cubic (fcc) metals with grain sizes in the micron, would suggest ‘ 6 d/4  d/3 (see e.g. [9]). This would
lead to the conclusion that s P (3  4)G(b/d), or
*
Corresponding author. Tel.: +1 858 534 6888; fax: +1 858 534
s/G P (3  4)(b/d). For pure Ni, since G  82 GPa, then
6373. if d  1 lm, s P 82 MPa, which is reasonable. If, how-
E-mail address: rasaro@san.rr.com (R.J. Asaro). ever, d  30 nm then s P 3280 MPa, which is too large

1359-6454/$30.00  2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2005.06.033
4826 B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838

by at least a factor of nearly 3! Not surprisingly, exper- and their characteristic grain sizes, are influenced by
imental evidence shows that, at grain sizes in this nano- temperature, lattice spacing, and stacking fault energy,
crystalline range, grains seem free of dislocations in their as illustrated below. In fact stacking fault energy, inter
interior (see e.g. [10,11]). Likewise molecular dynamic alia, is revealed to be a particularly important factor
simulations of deformation of such nanocrystalline in determining deformation mechanism and strength le-
polycrystals do not reveal intragranular dislocation vel. Moreover, it is shown that the proposed mecha-
activity as part of the deformation process [5–7]. nisms lead to an inherent strain rate sensitivity that is
As importantly, at the nanocrystalline grain size associated with the various mechanisms and that is also
scale, common fcc metals display an increasingly large part of deciding which mechanism is dominant at a par-
sensitivity to strain rate, a property that affords them ticular imposed strain rate.
ductility even though they have large strength. For later Asaro and Suresh [12] have recently explored mecha-
reference, and for perspective, we note that at a grain nisms involved in the nucleation of either perfect or par-
size of 20 nm the value of m from Fig. 1 is on the order tial dislocations as well as deformation twins from grain
of m  0.035 and 1/m  28, as would appear in a power boundaries for the specific purpose of rationalizing the
law relation between strain rate and stress, i.e., as in a above mentioned trends of increasing strain rate sensi-
relation of the form _ / r1=m . This may be compared tivity in nanocrystalline fcc metals. Their analysis re-
to values of 1/m for fcc metals with more traditional veals the influence of both the intrinsic stacking fault
grain sizes, also as shown in Fig. 1, of 50–100 or so. Lit- energy and the so-called unstable stacking energy intro-
erature data, compiled in Fig. 1, include data for nano- duced by Rice [14] in his analysis of dislocation emission
crystalline metals and Cu with nano-scale twins [12]. from a crack tip. Asaro and Suresh [12] indeed demon-
Recently, Asaro et al. [13] developed a simple, yet strate how at grain sizes below 20 nm, or so, the activa-
compelling, mechanistic model for the most likely sce- tion volume for such emission processes falls to low
nario for the anticipated transitions in deformation values on the order of 10–20b3 and, accordingly, the
mechanisms that appears to be quite consistent with strain rate sensitivity exponent, 1/m becomes on the or-
experimental observations that exist to date, and which der of 20–30 or less. We will not make specific use of
as we show herein leads to a complete constitutive the- their results in what follows except to note this fact
ory amenable to computational analysis. The model is found both experimentally and theoretically as, for
based on the notion, which is shown to follow naturally example, from their analysis.
from dislocation and partial dislocation mechanics, that It is evident, therefore, that there is a very strong sen-
at grain sizes below, say, 200–300 nm deformation oc- sitivity of mechanisms and overall response to grain size
curs via grain boundary dislocation emission; at grain and this will influence the response of nanostructured
sizes below, say, 20–30 nm deformation may occur in- polycrystals. This also means that in a given material
stead by the emission of partial dislocations or deforma- sample the distribution of grain sizes is vital to establish-
tion twins, and at even finer grain sizes deformation ing response. For example, and most particularly, we
occurs by grain boundary sliding. These mechanisms, will explore data obtained for electrodeposited pure Ni
that will explain that in a typical polycrystal with an
average grain size (by number count) of say 20 nm, the
Grain Size (nm) much fewer grains by number count in the size range
1000 100 50 20 >20 nm account for at least 50% of the volume fraction
0.04
– they, therefore, can dominate the behavior. In typical
nanocrystalline polycrystals, therefore, it is vital to pre-
cisely document the distribution of grain size as well as
0.03
to have an analytical (i.e., theoretical) framework for
interpreting the deformation mechanisms responsible
for the observed deformation. It is the development of
m

0.02
such a theoretical framework, a procedure for account-
ing for grain size distribution, and the analysis of exist-
ing data, that is the main goal of this series.
0.01
This paper is designed to be Part I of a series which ex-
Nanotwins in 500 nm grains
Literature data
plores the sensitivity of nanocrystalline aggregate re-
sponse to grain size distribution as well as to strain
0.00
rate, and temperature; the plan of Part I is as follows.
0.0 0.1 0.2 0.3
The Asaro–Krysl–Kad (AKK) model is reviewed and ex-
d-1/2 (nm-1/2) tended in the next section, where experimental data are
Fig. 1. Strain rate sensitivity exponent vs. grain size for nanocrystal- also described to illustrate its consistency. In Section 3,
line and ‘‘traditional’’ grain size fcc metals. the model is rendered into a constitutive theory for
B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838 4827

nanocrystalline aggregates, and in Section 4 some exam- in the side boundaries, or sbd dx ¼ 2ð12Gb2 Þdx. This leads
ples of calculations are described. We adapt the ex- to the remarkably simple result, s/G = (b/d). The d 1
tended Taylor model of Asaro and Needleman [15] for scaling of stress level derives simply from the fact that
this purpose and note that, within the context of that the area over which work can be performed by the ap-
aggregate model, grain distributions are to be expressed plied shear stress itself scales with d for a given dx. As
as volume fractions as described below. Grain size distri- noted by AKK, this leads to forecasted shear stresses
butions, as obtained for electrodeposited nanocrystal- that are too high for grain sizes less than, say 20–
line Ni by Dalla Torre et al. [16], are analyzed in 30 nm for typical fcc metals for which data exists.
detail and our constitutive model is used to analyze For the case of the emission of a partial dislocation,
experimental data for pure nanocrystalline Ni. Conclu- there is on the one hand a reduced requirement for work
sions follow in Section 5. Part II of this series will be associated with the residual segments in the side bound-
concerned with a more extensive statistical analysis of aries (owing to their lesser energy per unit length), but
available data for nanocrystalline Ni and Pd as well as now the additional requirement of creating a stacking
Cu and Al. fault with an energy C per unit area. The analysis of
AKK is not repeated here, but the result for the emis-
sion criterion is:
2. Background: the AKK model ða  1Þ ~
fðsms =GÞbsð1Þ =jbj þ ðsmz =GÞbð1Þ
z =jbjg ¼ C þ 1=3b=d;
a
Fig. 2 will serve as a point of departure in reviewing
ð1Þ
the model of Asaro–Krysl–Kad [13]. The figure illus-
trates a process of emission of a dislocation from a grain ~  C=Gb. Thus if we define
where a ” d/deq and C
boundary into the interior of a grain. Note that the fig-
sðaÞ  sms bsð1Þ =jbj þ smz bzð1Þ =jbj ð2Þ
ure describes details associated with the emission of a
partial dislocation, but it can also be used to explain we obtain
the result of the emission of a perfect dislocation. As ex- ða  1Þ ~
plained by AKK, as the segment is emitted into the sðaÞ =G  1=3b=d þ C; ð3Þ
a
grain it creates two trailing segments in the ‘‘side grain
boundaries’’, and in the case of a partial dislocation a which is AKKs equation (7). We also recall that deq is
stacking fault within the grain. The energy (per unit defined as the equilibrium spacing of Schockley partials
length) of these segments is taken as 1/2Gb2 for the per- in the absence of applied stress and is given as,
fect dislocation, and 1/6Gb2 for the partial dislocation, 1
respectively. The explanation for the latter value lies deq ¼ ðGb2 Þ=C. ð4Þ
12p
simply in the fact that for a Schockley partial dislocation
in a FCC crystal, the magnitude of the Burgers vector is The coordinate system of the dislocations slip system is
p
bpartial = 1/ 3bperfect. As is evident, b ” bperfect is the such that m is the unit normal to the slip plane, s is unit
magnitude of the perfect Burgers vector. Now in the vector along the direction of the perfect Burgers vector,
case of the emission of a perfect dislocation, the mini- b, and z is the third of a right-handed unit triad and lies
mum required resolved shear stress is that required to in the slip plane as well. AKK explore the effects of
perform the work of creating the two residual segments material parameters such as stacking fault energy, mod-
ulus, and lattice parameter, inter alia, and successfully
explain some key trends described in the existing
literature.
An example of experimental data is shown to illus-
trate the consistency of these predictions, and those to
follow, and what is known experimentally. Fig. 3 shows
stress vs. strain data for tension for electrodeposited Ni
with a mean grain size of approximately 20 nm. In the
first place, we note that for this Ni the ultimate strength
at the highest imposed strain rate (_ ¼ 4:5  103 s1 ) is
approximately rult  1800 MPa. If we take r  2s at
yield and just after, the above with data typical of Ni
(see Table 2), suggests that rult  2 · 1115 MPa =
2230 MPa, which is of the order of the observed ultimate
Fig. 2. Emission of a partial dislocation from an existing dislocation at
strength but is somewhat larger than the experimental
a grain boundary. As the leading partial dislocation enters the grain it value of 1800 MPa. We will reconcile this difference by
produces partial segments along the side boundaries. analyzing the polycrystal with a suitable grain size
4828 B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838

initiation of either perfect or partial dislocations from


grain boundaries that requires comment. Such discus-
sion is provided below.
Between the highest strain rate imposed and the low-
est, viz. _ ¼ 4:5  105 s1 , there is an approximately
200 MPa reduction in flow stress. Thus there is indeed
a significant sensitivity to strain rate which, if analyzed
via a power law as given above, would be estimated in
the range 1/m  25. This order of magnitude will
be used later vis-à-vis a phenomenological power law ki-
netic expression for the shear rate. A value of 1/m  25
is also consistent with the trends shown for the data of
Fig. 1, as noted in Section 1.
We now estimate the grain size, say for Ni, at which
there would be a transition from perfect dislocation
emission to partial dislocation emission by simply equat-
ing (b/d) to Eq. (3) as,
~
ðb=dÞ ¼ 1=3ðb=dÞ þ ~aC; ð5Þ
ða1Þ
where ~a  a
For Ni, ~a  1 and the estimate is that
.
the transition would occur at a grain size of approxi-
Fig. 3. Tensile behavior of nanocrystalline pure Ni with average grain mately d  70b, or d  20 nm! The implications of such
size of 20 nm. Electrodeposited nanocrystalline Ni was obtained in the transitions for polycrystalline response will be illustrated
form of 80 mm · 80 mm foils, 200 lm thick, from Goodfellow Metals.
below via our simulations.
Dog-bone shaped tensile specimens were cut, via EDM, with gauge
dimensions 500 lm · 200 lm. The gauge length was 12 mm. Tests were
conducted quasi-statically in a mechanically driven Instron test system
at 20 C at the strain rates indicated. 3. A constitutive model

3.1. Perspective on discreteness


distribution as fitted to experimentally determined size
distributions. We also note the essential absence of
To begin, it is revealing to inquire as to the actual
strain hardening which may be also forecast from the
number of dislocations or stacking faults that would
AKK model as described by Eq. (1). There is, however,
need to be emitted into a typical grain whose dimensions
much more that requires comment concerning the
are on the order of, say 20 or 30 nm. Fig. 4 presents a

Table 1 s
List of faulting systems
Slip Perfect Burgers Leading partial, Trailing partial, d
s = nb
plane vector b+ b
n = no. dislocations/
(1 1 1) 
½1 0 1 
½1 1 2 ½2 1 1

stacking faults
(1 1 1) ½0 
1 1 ½1 2 1 ½1 1 2
suppose b = 0.3nm, then how many disl's/SF ?
(1 1 1) ½
1 1 0 ½2 1 1 ½1 2 1
ð
1 1 1Þ [1 1 0] [2 1 1] ½1 2 1
ð
1 1 1Þ ½0 
1 1 ½1 2 1 ½1 1 2 grain
ð
1 1 1Þ ½
10 1 ½1 1 2 ½2 1 1
 d = 10nm 20nm 30nm
size
ð1 
1 1Þ [1 1 0] [1 2 1] ½2 1 1 γ = 0.05 n = 1.5 3 4.5
ð1 
1 1Þ ½0 
11 ½1 1 2 ½1 2 1

γ = 0.08 n = 2.4 4.8 7.2
ð1 
1 1Þ ½
1 0 1 ½2 1 1 ½1 1 2
ð1 1 1Þ [1 0 1] ½2 1 1 [1 1 2]
ð1 1 1Þ ½0 
11 ½1 1 2 ½1 2 1
Fig. 4. How many dislocations or stacking faults are emitted into a
ð1 1 1Þ ½
1 1 0 ½1 2 1 ½2 1 1
nanocrystalline grain?

Table 2
Material data used in the simulations
G (GPa) b (nm) b (nm) C/Gb deq (nm) DF (kJ/mol) v (m3) m c_ 0
29
80 0.249 0.144 1/100 0.5 97.5 1.54 · 10 0.037 103
B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838 4829

simple analysis of this for perspective. The important direction. This is because each perfect slip system natu-
point to be illustrated here is that to induce strains that rally defines two partial systems, one for slip along the
are less than, say 10%, the number of defects involved in parent Æ1 1 0æ direction and the second for slipping in
a given grain is itself quite modest and typically less than the anti direction. Thus slip in what may be referred to
10. This, in turn, illustrates the rather discrete nature of as the + and  directions, in any {1 1 1} plane, can be de-
the deformation process that occurs in grains of this scribed by 24 partial systems. Note that partial slip
size. We will, nonetheless, view the process in what fol- events occur in particular sequences, e.g., slipping over
þ
lows as occurring in a continuous manner despite its the perfect Burgers vector, ~b, occur in the sequence, ~ b

obviously discrete nature. For analyzing polycrystalline followed by ~b . Slipping in the direction ~b on the other
 þ
regions, this is viewed as representing suitable averaging. hand occurs in the sequence ~b followed by ~b . Now
In other words, we take the view that the discrete slip we recall that slip events involving partial dislocations,
increments given grains undergo over a short time scale i.e., the emission of stacking faults into the grains, in-
on the order of 106 s can be time-stretched over much volve emitting only the leading partial dislocation.
longer times, provided the average ensembles strain Therefore, from the perfect slip system, (m, b) we obtain
increment is the same in both cases. This means that the two partial slip systems, (m, b+) and (m, b). Table 2
in the analysis, and at any given moment, the number lists the 24 such partial slip systems, each associated with
of defects emitted will be non integer, yet the average a perfect slip system.
plastic strain in a representative polycrystalline aggre- The restriction of allowing only the ‘‘leading’’ partial
gate will be equal to what would have been produced dislocation to be emitted derives from several rationales.
by discrete slip events in a discrete group of grains. First, we recall from the dislocation analysis above that
on energetic grounds the emission of a single partial dis-
3.2. Dislocation and partial dislocation slip systems location is favored over the emission of a perfect dislo-
cation at sufficiently small grain sizes, Secondly, and
The slip systems associated with perfect dislocations entirely consistent with this view, are the results of a ser-
are the usual 12 (or 24) fcc systems involving the octahe- ies of molecular dynamics simulations performed by
dral {1 1 1} planes, and the face diagonal Æ1 1 0æ directions Van Swygenhoven and co-workers [5–7], which show
lying within those planes. The choice of 12 or 24 is simply the same phenomena as summarized below.
one of allowing the slipping rate to possess algebraic sign Before reviewing this, however, the complete list of
or to be strictly positive, respectively. In the latter case, partial dislocation, i.e., stacking fault based, slip systems
slip systems describing slip in positive and negative direc- is provided (see Table 1).
tions are required. Fig. 5 illustrates the kinematics asso- These will provide the basis for developing the kine-
ciated with partial dislocations, which are themselves matics of slip via the emission of partial dislocations
associated with parent perfect dislocations. The parent, from grain boundaries. We will not include, in what fol-
i.e., perfect, dislocation Burgers vector is denoted as ~b. lows, the specific mechanism of deformation twinning
Partial systems are associated with each parent fcc slip as, for example, analyzed by Asaro and Suresh [12],
system, and for our purposes we may restrict attention but note that in the context of the present analysis the
to slipping along only a single sense along each Æ1 1 0æ criteria for twin emission would be similar to that used
here for partial dislocation emission.

3.3. Perspectives from molecular dynamics simulations

The simulations performed by Swygenhoven and co-


workers, noted above, have provided much insight into
the process of stacking fault emission in sufficiently
small grains. Fig. 6 illustrates an example of stacking
fault emission in Ni taken from their simulations. The
stacking fault is indicated in orange and the grain
boundary atoms are colored as blue. The fault is shown
at a particular time as it passed through the grain as it,
in fact, did in this simulation. There are several aspects
to the computed phenomenology that are directly rele-
vant to the development of a continuum theory for such
deformation.
The partial dislocations observed in such simulations
are generally observed to transit entirely across individ-
Fig. 5. Partial dislocation slip systems. ual grains, yet on occasion multiple faults in a grain will
4830 B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838

on the a fault system within a given grain during the


time period T, where each contributes a strain D(i),
then the average strain rate they contribute during that
time interval would be,
nðaÞ
1 X
p
_ ¼ DðiÞ . ð6Þ
T i¼1
Next realize that, as described by Table 2, there are
some 24 distinct fault systems that are, in general, acti-
vated in parallel. The total strain rate is then, following
the example of relation Eq. (6),
1 Xn
_ p ¼ nðaÞ DðaÞ ; ð7Þ
T a¼1 g
where, after examining Fig. 4, we identify D(a) as
1
DðaÞ ¼ bðaÞ mðaÞ . ð8Þ
d
In Eq. (8), b(a) is the Burgers vector of the ath fault sys-
tem, and m(a) is the normal to the plane that fault acts
Fig. 6. Emission of a stacking fault from molecular dynamics [5].
on. At this point the approximation is made of taking
the time average over a single nano-grain, or an ensem-
interact and appear to block each other. Such occur-
ble average over many such nano-grains, to obtain,
rences may, in fact, give rise to at least modest levels
of strain hardening. When faults are preferred (over Xn
1
_ p ¼ n_ gðaÞ bðaÞ mðaÞ . ð9Þ
for example grain boundary sliding which is preferred d
a¼1
at very fine grain sizes), only the leading partials are ob-
served, as described and rationalized above. The leading The above relation is to be calibrated vis-à-vis the data
edge of faults, i.e., the partial dislocations themselves, is shown in Fig. 3.
generally not perpendicular to the ‘‘side boundaries’’ as To this end we recall the data shown in Fig. 3 which
is readily apparent in the figure. This suggests that there reveals a clear rate sensitivity. We assume that this arises
is resistance to motion of the partial dislocation within from the rate dependent process of producing what we
the side boundaries. This, in turn, may derive from the have referred to as the side boundary partial dislocation
need to create segments of partial dislocations in the side segments whose energy was described as, 16Gb2 above in
boundaries. Imagine for example, that the partial dislo- addition to the nucleation of dislocations as described,
cation were observed to be tangent to the side boundary. for example, by Asaro and Suresh [12]. At sufficiently
This would mean the partial dislocation would exert a low rates of deformation we assume that the need for
line tension force on the boundary equal to its energy such segments is precluded. Thus, we postulate a kinetic
per unit length, which is as noted above, equal to law for n_ ðaÞ as
1
Gb2partial . This leads to the same picture as outlined 8 n o
2 < ðaÞ ~ 1=m
above for the required shear stress for partial motion. ðaÞ n_ 0 s =G~ ðaÞ
aC ~ > 0;
if sðaÞ =G  ~aC
n_g ¼ g ð10Þ
But this particular simulated event shows an angle much :
0 otherwise;
greater than 0 which, of course, indicates that less work
is being performed at these side boundaries. This has di- where
rect relevance to the discussion associated with Fig. 2
and the effect of the first term in Eq. (3) and with a pos- 1
gðaÞ ðnÞ ¼ b=d þ f ðnÞ. ð11Þ
sible, albeit partial, explanation of the rate sensitivity 3
seen in Fig. 3. Some additional explanation is required at this stage of
development. In Eq. (10), the rate of generation of
3.4. A continuum theory for deformation via the emission faults, n_g ðaÞ , represents the rate of generating faults in
of stacking faults the entire grain; accordingly we can define the rate
n_ ðaÞ ¼ d1n_ gðaÞ as that per unit length of grain boundary.
With the perspectives offered above, a continuum the- If, however, faults are preferentially generated at grain
ory for deformation of nanocrystalline metals, via the boundary triple points, there may well be a grain size
emission of stacking faults, can be constructed as fol- dependence to this generation rate. Now, fault genera-
lows. Suppose that n(a) stacking faults have been emitted tion is only possible if the driving stress is sufficient to
B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838 4831

do the work of creating the associated stacking faults


and this is reflected by the fact that n_ gðaÞ , and n_ ðaÞ would
vanish if sðaÞ =G  ~ aC~ 6 0. At sufficiently low rates of
deformation, fault generation would ensue as long as,
as noted sðaÞ =G  ~ aC~ > 0, but at high deformation rates
ðaÞ
s =G  ~ ~
aC would have to be a significant fraction of
g(a); how significant would, of course, depend on the
strain rate sensitivity parameter, 1/m. As noted below,
during the discussion of calibration, the numerology
associated with the data of Fig. 3 suggests that
1=m  Oð20Þ or Oð30Þ; for that matter the same data
suggests that n_ 0  Oð5  102 Þ  Oð5  101 Þ. The inter-
sections of emitted faults represent impediments to con-
tinued fault propagation, and this represents a form of
strain hardening. To account for this, the phenomeno-
logical term f(n) P has been included. Here, and as an
example, n  na¼1 nðaÞ and f(n) describes an increase in
the resistance to generating and fully emitting faults as
faults accumulate within a given grain. Finally, let
b(a) = b(a)s(a), i.e., let s(a) be a unit vector along b(a) and
b(a) is the magnitude of the partial dislocation Burgers Fig. 7. A composite model for SF emission and GBS.
vectors, that we immediately note are all crystallograph-
ically equivalent in a fcc crystal. Thus, since bðaÞ ¼ b, for
all systems we can write,
a rationale for the transition between dislocation in-
Xn X n
duced deformation and GBS. This will turn out, not
_ p ¼ n_ ðaÞ 
bsm ¼ c_ ðaÞ sm; ð12Þ
a¼1 a¼1
surprisingly, to lead to a picture that is strain rate depen-
dent. Fig. 7 illustrates the scheme, consistent with their
where model. They construct what they refer to as the macro-
8 n ðaÞ o
<1 ~ 1=m scopic shearing rate as,
bn_ 0 s =G~ aC
if sðaÞ =G  ~
aC~ > 0;
c_ ðaÞ ¼ d g ðaÞ
ð13Þ
: 6btD
0 otherwise; c_ ¼ sinhðvs=kT Þ expðDF =RT Þ; ð15Þ
d
and where we may define a reference shearing rate as
c_ 0  d1n_ 0 
b so that Eq. (13) becomes, where tD is a typical lattice vibrational frequency
8 n (1013), v is an atomic volume (taken as b3 by them),
o
~ 1=m
< ðaÞ and DF is an activation energy for localized lattice/grain
ðaÞ c_ 0 s =G~ ðaÞ
aC
if sðaÞ =G  ~
aC~ > 0;
c_ ¼ g ð14Þ boundary diffusion. We further propose to interpret this
:
0 otherwise; shearing rate as an effective plastic strain rate within the
framework of an isotropic flow theory, in particular
3.5. A composite model for grain boundary sliding J2-flow theory. qInffiffiffiffiffiffiffiffiffiffiffi
Eq. ffi(15), s is an effective shear stress
defined as s ¼ 12r0ij r0ij , so that sc_ represents the rate
As noted above, when the grain size is sufficiently of working per unit volume of material. J2-flow theory
small deformation may occur via grain boundary sliding assumes that the plastic part of the rate of deformation,
(GBS). This was described in the AKK model via the Dps, is directed so that Dpsir 0 , or given the definition of s
phenomenological kinetic relations proposed by Conrad that
and Narayan [17]. Fig. 7 above illustrates a composite
Dps kos=or. ð16Þ
scenario for such dual deformation mechanisms. This
model assumes that that as individual nanocrystalline The flow law for this part of the deformation rate then
grains deform by the emission of stacking faults, they becomes,
may also flow past each other via sliding along their mu-
1
tual boundaries. The result would be the formation of Dps ¼ c_ r0 =s. ð17Þ
shear zones, highlighted in ‘‘red’’ in the figure. As noted, 2
Conrad and Narayan [17] have recently provided a cor- At this point we assume that the total plastic rate of
relation between a simple phenomenological model of deformation is given as the sum of what is contributed
grain boundary sliding in nanostructured metals and by SF emission and by GBS, i.e., we take for the rate
available data – we use that model to attempt to provide of deformation due to SF emission _ p ) Dpsf . Thus
4832 B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838

O
the total plastic part of the rate of deformation is given
s ¼ s_  X  s þ s  X ¼ L : D þ D  s þ s  D . ð26Þ
as,
In Eqs. (24) and (25) the two Jaumann stress rates are
Dp ¼ Dpsf þ Dps . ð18Þ
O O Xn

These specifications for the plastic parts of the deforma- s ¼ s þ bðaÞ c_ ðaÞ ; ð27Þ
a¼1
tion rate will now be inserted into a finite deformation
kinematic and full constitutive framework. where
bðaÞ ¼ WðaÞ  s  s  WðaÞ ; ð28Þ
3.6. Constitutive framework

The material is assumed to flow through the partial 3.7. An aggregate model
dislocation slip systems and via GBS, whereas elastic
deformation occurs via deformation of the lattice itself We follow here the construction of Asaro and Nee-
(with material embedded on it). Rigid body rotations dleman [15] and imagine a collection of grains all of
are included in this lattice part. Hence as in [18] or in which are characterized by a constitutive law such as gi-
the context of crystalline slip, in [19], the deformation ven in Eq. (24). The aggregate occupies a volume V
gradient is decomposed as, bounded by a surface Sext. The region is subjected to
all around displacement conditions that are taken to
F ¼ F  Fp ; ð19Þ
be such as to cause homogeneous deformations in what
p
where F* and F are the lattice and plastic parts of F, is assumed to be an initially homogeneous sample. Aver-
respectively. Note that the crystallographic vectors aging theorems that rely on strict equilibrium and com-
introduced earlier, i.e., s and m convect with the lattice patibility are available, but here we will have to relax the
so that in the deformed state they become, requirement of point-wise equilibrium as follows.
Introduce the nominal (1st Piola–Kirchhoff stress) as
s ¼ F  s; m ¼ m  F1 . ð20Þ
n ¼ F1  s. ð29Þ
Note that as s and m are originally orthogonal, so are s*
and m*. Using Eq. (19) to form the velocity gradient we Note that n is related to the force, dP, transmitted
uncover its lattice and plastic parts, which can be ex- through a surface element m dS in the reference state
pressed in via dP = m Æ n dS. Now let F be a deformation gradient
that satisfies compatibility in each grain. Then using
 X
n
1 the divergence theorem,
F_  F1  F_  F1 ¼ Dp þ Xp ¼ c_ ðaÞ s m þ c_ r0 =s;
a¼1
2 Z XZ Z I
ð21Þ n : F dV ¼ n : F dV ¼ DT  u dS þ T  u dS.
V g Vg S int

where all terms have meanings as described above. Now ð30Þ


define
In Eq. (30), Vg represents the volume of a grain, Sint
ðaÞ 1 1 represents the internal interfaces between grains (i.e.,
P ¼ fs m þ m s g and WðaÞ ¼ fs m  m s g.
2 2 grain boundaries), and DT is the traction difference
ð22Þ between the grain boundaries. Note that if point wise
Then equilibrium held, DT = 0. Note also that an identity of
the form Eq. (30) holds with F or n, or both replaced
X
n
1 X
n
Dp ¼ c_ ðaÞ PðaÞ þ c_ r0 =s and Xp ¼ c_ ðaÞ WðaÞ ; by their corresponding rates since the equilibrium equa-
a¼1
2 a¼1 tions expressed in terms of n are, nij,j = 0 and in rate
ð23Þ form n_ ij;j ¼ 0 in the absence of body forces. Here, as typ-
ical, commas denote differentiation and the summation
where we have explicitly made the assumption that convention is used.
GBS, expressed within a J2-flow theory framework, con- The homogeneous displacement boundary conditions
tributes nothing to the plastic spin. As shown in [15], the to be imposed take the form u ¼ F   x on Sext. Substitut-
resulting constitutive relation for a crystal is given as ing this into the last integral, and rearranging terms, in
Xn   Eq. (30) gives
O 1 0
s¼L:D RðaÞ c_ ðaÞ  L : c_ r =s ; ð24Þ
2 XZ Z I
a¼1
n : F dV  DT  u dS ¼ f xT dSg : F. 
where g Vg S int
S ext
ðaÞ ðaÞ ðaÞ ðaÞ ðaÞ ðaÞ
R ¼L:P þb ; b ¼W ssW ; ð25Þ ð31Þ
B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838 4833

The basic assumption underlying Asaro and Needle- subject to statistical uncertainty that we specifically ad-
mans [15] averaging model is that the contributions dress here. For grain size distributions, the numbers of
arising from any lack of equilibrium across the grain grains of various diameters in nanocrystalline materials
boundaries will be small compared to the integral taken can usually be well represented by a log-normal distribu-
over the external surface, an approximation difficult to tion function,
precisely measure in general. Invoking this assumption, "  2 #
however, yields 1 1 lnðD=D0 Þ
P ðDÞ ¼ exp  ; ð40Þ
8 9 ð2pÞ1=2 Dr 2 r
XZ <I =
n : F dV ¼ xT dS : F. ð32Þ where the symbol D is used for the grain diameter and
V : ;
g g
S ext D0 and r are constant parameters describing the median
 in and shape parameters (i.e., the standard deviation
In particular, if we specify F to have the uniform F of
each grain R 1 lnD) of the distribution, respectively, and
P ðDÞ dD ¼ 1. The arithmetic mean size D  can be
XZ 0
calculated [20] as
n : F dV ¼ V  
n : F; ð33Þ  
g Vg
 1 2
D ¼ D0 exp r . ð41Þ
where 2
Z
1 X X ~, the variance of D, is the average value of the quantity
r

 n dV ; V ¼ V g. ð34Þ  2 , and is calculated from
ðD  DÞ
V g Vg g

 is trivially the average of F in the aggregate, ~ ¼ D20 expðr2 Þðexpðr2 Þ  1Þ.


r ð42Þ
Since F
the average Kirchhoff stress is Commonly, the arithmetic mean grain size is often gi-
 ven by material suppliers or can be obtained from the
s ¼ F n. ð35Þ
experimental measurement; if we assume the variance
To put the constitutive description in terms of the ~ is known, then the parameters, D0 and r which de-
r
nominal stress and rate of change of the deformation scribe the log-normal distribution can be obtained by
gradient we need only recall from Eq. (29) that inverting Eqs. (41) and (42), which yields
1
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n_ ¼ F1  s_ þ F_  s; ð36Þ D4
D0 ¼ ; ð43Þ
and from the definition of the Jaumann rate of Kirch- ~þD
r 2
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi!
hoff stress that,
r
~
O r ¼ exp ln þ1 . ð44Þ
s_ ¼ s þX  s  s  X. ð37Þ 2
D
These two expressions allow the constitutive relation, Most of the relevant available literature considers
Eq. (24), to be rephrased as that the mean grain size, internal strains and texture ob-
n_ ¼ K : F_  B;
_ ð38Þ tained from TEM and XRD well describe the structure
of the sample material (e.g. [16]). However, there is usu-
and if we designate individual grains with the super- ally considerable uncertainty associated with both of
script g, the constitutive relation for the grain g these methods (see e.g. the discussion in [21]). For exam-
becomes, ple, the question of whether a true characteristic re-
ðgÞ ðgÞ gion(s) of the sample was investigated for TEM can be
n_ ðgÞ ¼ KðgÞ : F_  B_ . ð39Þ
routinely raised. Therefore, an uncertainty analysis on
We recognize that when any such relation is constructed the grain size distribution is necessary, in order to pro-
for an individual grain that it is most naturally done vide a more accurate description of the actual behavior
with respect to a set of axes that are native to the crys- of the material, especially with respect to interpreting
talline grain in question, i.e., on crystal axes belonging experimental specimen to specimen scatter.
to that grain itself. The uncertainty analysis of this paper is conducted as
follows. We assume that the arithmetic mean grain size
3.8. Grain distributions is given, and what is uncertain is the variance of the
grain size distribution. According to Eqs. (43) and
Experimental techniques available for the full charac- (44), we obtain the parameters (the median and vari-
terization of structure at the nano-scale have been pri- ance) of the log-normal distributions for a certain arith-
marily based on transmission electron microscopy metic mean grain size, using reasonable standard
(TEM) and X-ray diffraction, from which the average deviations. An example is shown in the left side figure
grain size and grain size distributions can be estimated, of Fig. 8. In this the mean grain size is taken to be
4834 B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838

20 nm, and the variance varies from 50 to 200. As noted 3.9. Calculational procedure
below, this mean grain size is consistent with that mea-
sured for electrodeposited Ni as tested by Dalla Torre 3.9.1. Boundary conditions
et al. [16] and ourselves. Denote the axes of a fixed Cartesian ‘‘laboratory’’
However as mentioned earlier, for the current analy- frame by ei and the cube axes in the grain g by ai; for
sis of dislocation emission and grain boundary sliding, example, if a tensile test were being set up the tensile axis
the volume-weighted grain distribution is considered to could be the e2 axis. For a FCC crystal the e2 axis would
be more directly relevant than the number-weighted fall somewhere within the standard triangle whose ver-
grain distribution as described by Eq. (40). If one as- texes are the [1 0 0], [1 1 0], [1 1 1] crystal directions. The
sumes each grain in the sample has the same shape, a initial orientation of the ‘‘laboratory’’ axes on with re-
sphere for example, then the total volume of the sample spect to the cube axes of a the typical grain g are de-
is scribed via the transformation,
Z 1 ðgÞ
ei ¼ UðgÞ  ai . ð47Þ
V ¼ kD3 P ðDÞ dD; ð45Þ
0 Note that this involves describing the orientation of
where k is a constant describing the shape of the grains each grain in a polycrystalline aggregate by the specifi-
(e.g., k ¼ 43p213 for a sphere), and the volume-weighted cation of its transformation tensor, U(g). Methods for
grain size distribution is then simply constructing such distributions are discussed in detail
by Asaro and Needleman [15]. For the calculations
kD3 P ðDÞ presented herein, we have taken the polycrystal as
Pv ¼ . ð46Þ
V being isotropic and thus initially composed on a uni-
We can then convert the log-normal distribution to a formly random distribution of grain orientations. Sup-
volume fraction-weighted distribution according to Eq. pose that there are N such grains, and further suppose
(46). The right side figure in Fig. 8 shows the volume- that each has the same initial volume, i.e., each occu-
weighted grain distributions for the above example, pies the same initial volume fraction of the aggregate.
whereas the left side figure illustrates the frequency dis- Then when averages are taken as specified in Eq. (34)
tribution. These distributions are illustrated via different for example the aggregates constitutive response is de-
forms below and in specific connection to the simulated scribed by,
experimental results.
The resulting volume-weighted grain distributions are n_ ¼ K _  B;
 :F _ ð48Þ
then used in the aggregate model of Asaro and Needle-
where
man [15]. For efficiency, the volume-weighted distribu-
tions are grouped into histograms with suitable 1 XN XN XN
subintervals, and the grains inside the aggregate are as- n_ ¼ n_ ðgÞ ; ¼ 1
K KðgÞ ; _ ¼ 1
B
ðgÞ
B_ .
N g¼1 N g¼1 N g¼1
signed the different grain sizes according to the relative
proportions described by the histograms. ð49Þ

Fig. 8. Grain size distributions based on an assumed mean grain size of 20 nm with to different variances.
B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838 4835

To use the tensile test as an example, we let the tensile and that F_ 22 be specified, that is
axis be the e2 ‘‘laboratory’’ axis; thus the line of material
particles initially along the e2 axis remains so aligned so F_ 22 ¼ specified stretching rate. ð54Þ
that,
For plane strain loading the third of Eqs. (52) is re-
  e2 ¼ F 22 e2 ;
F ð50Þ placed by
and
F 33 ¼ 0 ) F_ 33 ¼ 0. ð55Þ
F 12 ¼ F 32 ¼ 0; ð51Þ
Other types of boundary conditions are easily specified
so that tensile fiber remains along the e2 axis. by following the example of tensile loading described
The remaining boundary conditions for axisymmetric above.
loading are:
T 1 ¼ 
n11 ¼0) n_ 11 ¼ 0 on x1 ¼ constant; 3.9.2. Grain distributions for nanocrystalline Ni
T 1 ¼ n
31 ¼0)n _ 31 ¼ 0 on x3 ¼ constant; The grain size distributions represented in Fig. 8 are
shown again in different form in Fig. 9 as functions of

T3 ¼ n 33 ¼0) n_ 33 ¼ 0 on x3 ¼ constant;
ð52Þ the variance of the distribution. In Fig. 10 are shown

T2 ¼  n12 ¼0)n _ 12 ¼ 0 on x1 ¼ constant; the two distributions as measured and reported by
T 2 ¼ 
n32 ¼ 0 ) n _ 32 ¼ 0 on x3 ¼ constant; Dalla Torre et al. [16] for electrodeposited Ni and
our theoretical fit using the log-normal distribution
F 21 ¼ 0 ) F_ 21 ¼ 0. give by Eq. (40) above. In the simulations described
Note that Eqs. (50) and (51) also require that below, we use these distributions to describe the grain
size along with the material data listed in the next
F_ 12 ¼ F_ 32 ¼ 0; ð53Þ subsection.

0.07 0.06
Volume weighted Number weighted
Number weigthed Volume weighted
0.06 Variance = 50 0.05
Variance = 100
Variance = 150
0.05 Variance = 200
0.04
Frequency
Frequency

0.04
0.03
0.03
0.02
0.02

0.01 0.01

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
(a) Grain size(nm) (b) Grain size (nm)

Fig. 9. (a) Comparison of the distributions with different variance and (b) histogram of the distribution with mean = 20 nm and variance = 100.

70 60 0.5
Percentage of grains per 10 nm

Batch A Batch B
Number of grains per 2 nm

Number of grains per 2 nm

60 50
0.4
50
40
0.3
40
30
30
0.2
20
20
0.1
10 10

0 0 0
0 20 40 60 80 0 20 40 60 80 0 20 40 60 80
(a) Grain size (nm) (b) Grain size (nm) (c) Grain size (nm)

Fig. 10. (a), (b) Number-weighted grain size distributions according to Dalla Torre et al. [16]; (c) a simulated fit to the histogram of (b) ([16] Batch B).
4836 B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838

3.10. Material data Torre et al. [16] (i.e., their batch B) with a mean grain
size of 20 nm and our fitted variance of 100. The simu-
The applications below are carried out for the case of lated curves are clearly below the experimental curves.
nanocrystalline Ni whose properties are listed in Table 2 The right side figure in Fig. 11 shows the contributions
along with the grain size distributions described in the from the three deformation modes, viz. perfect and par-
previous subsection. tial dislocation emission, and grain boundary sliding, to
The phenomenological hardening term, f(n), in Eq. (11) the overall tensile deformation. With this distribution,
is neglected in the simulated cases presented below. the deformation is forecast to primarily involve the
emission of perfect dislocations.
The left side figure in Fig. 12 shows another simu-
4. Numerical simulations lated stress vs. strain curve, but with an assumed mean
grain size of 23 nm and variance of 70. Note that now
Fig. 11 shows the simulated results using the grain the agreement with experiment is much better except
size distributions described above for uniaxial tension for an apparent larger ultimate tensile stress as simu-
at the three different strain rates we performed our uni- lated than what the experimental curves appear to dis-
axial quasi-static tension tests for electrodeposited nano- play. This apparent discrepancy is discussed in the
crystalline Ni. The experimental uniaxial stress vs. strain next section, but here we claim the agreement is in fact
curves from Fig. 3 have been superposed for compari- quite good using a grain size distribution that nearly
son. Note that as described above, the mean grain size equally fits the Dalla Torre et al. [16] data as does that
here was taken from the best fit to the data of Dalla used to obtain the results shown in Fig. 11. Note also

2000 0.2
Strain rate 4.5e-3
Strain rate 4.5e-4
Strain rate 4.5e-5
1500 0.15 Full dislocation emission
Partial dislocation emission
GBS
Stress (MPa)

Shear strain

1000 0.1

Strain rate 4.5e-3


500 Strain rate 4.5e-4 0.05
Strain rate 4.5e-5
Numerical prediction
Experiment
0 0
0 1 2 3 4 0 0.05 0.1 0.15
Strain (%) Accumulated effective plastic strain

Fig. 11. Computed stress vs. strain response (left side figure) for nanocrystalline Ni with average grain size of 20 nm and variance of 100 as in
Eq. (42). Contribution to the overall plastic strain arising from the three deformation mechanisms considered. All simulations are carried out at 20 C.

2000
0.15 Strain rate 4.5e-3
Strain rate 4.5e-4
Strain rate 4.5e-5
1500 Full dislocation emission
Partial dislocation emission
0.1 GBS
Stress (MPa)

Shear strain

1000

Strain rate 4.5e-3 0.05


500 Strain rate 4.5e-4
Strain rate 4.5e-5
Numerical prediction
Experiment
0 0
0 1 2 3 4 0 0.05 0.1 0.15
Strain (%) Accumulated effective plastic strain

Fig. 12. Computed stress vs. strain response (left side figure) for nanocrystalline Ni with average grain size of 23 nm and variance of 60 as in Eq. (42).
Contribution to the overall plastic strain arising from the three deformation mechanisms considered but for a simulated nanocrystal with a mean
grain size of 10 nm and a variance of 50. All simulations are carried out at 20 C.
B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838 4837

that the predicted trend in strain rate sensitivity is in of strain hardening. In fact, it is most probable that most
good accord with experiment. This agreement has been recorded stress vs. strain response is terminated prema-
accomplished using a value of 1/m in complete accord turely due to the onset of fracture brought on by the pres-
with the data of Fig. 1, viz. 1/m  27. ence of small defects in the specimens.
The right side figure in Fig. 12 shows the computed Finally, we note that a comparison of the right side
contributions to the overall tensile plastic strain for a figures in Figs. 11 and 12 shows that when the grain size
simulated uniaxial tension test for a nanocrystal with a is at the transitional grain size range of 20 nm and just
mean grain size of 10 nm and a variance of 50. Note that below, as estimated for nanocrystalline Ni, the expected
with this smaller grain size there is noticeably more con- deformation modes indeed begin to change. What this
tribution from deformation modes including partial dis- means, however, is that at a grain size of 20 nm, and
locations (i.e., stacking fault emission) and grain with a grain size distribution typical of electrodeposited
boundary sliding. Of course, as the temperature is in- Ni, observation of faults or twins would be anticipated
creased we expect contributions from modes such as to be relatively rare events. In fact, as the variance of
grain boundary sliding to increase further [17,22]. It the grain size distribution increases beyond say 100, as
should be recalled from Eq. (5) that, given the material we have defined this parameter, it is unlikely that any
properties of Ni, the grain size where we expect a tran- deformation twinning or fault emission would be ob-
sition from perfect to partial dislocation emission is served at all. At mean grain sizes below 15 nm such
approximately 20 nm – this is what these simulations modes become prevalent enough to be anticipated via
show, at least at ambient temperatures like 20 C. TEM observation but only in the smallest of the grains
In Part II of this series we explore, in far more detail, in the distribution. Again this shows the need for mate-
the influence of grain size distribution, material proper- rials synthesized with more uniform grain sizes than are
ties, temperature, and strain rate on material response. currently available. With grain sizes below 15 nm, grain
This is also analyzed vis-à-vis stress and strain state boundary sliding is also seen to contribute to the overall
via example simulations of plane strain vs. uniaxial ten- deformation as seen by the right side of Fig. 12.
sion and compression deformation. Part II of this series will explore the effect of grain size
distribution, strain rate, and temperature in far more de-
tail and use additional data for nanostructured Ni and
5. Discussion and conclusions Pd, along with the more limited data available for Cu
and Al.
The examples presented above have illustrated the
main focus of this discussion, viz. the vital influence of
grain size distribution on the response of nanocrystalline Acknowledgments
metals. As an example, we have shown that for electro-
deposited Ni with a mean grain size in the range of R.J.A. and P.K. thank the National Science Founda-
20 nm (typical for materials that have been reported tion NIRT initiative, specifically under Grant #0210173.
on recently) the response is sensitive to grain size distri-
bution, yet not simply to the mean grain size, but also to
References
the spread about the mean grain size as measured by the
variance about the mean, as defined in Eq. (42). This [1] Gleiter H. Nanocrystalline materials. Prog Mater Sci
illustrates, inter alia, that future experiments need to 1989;33(4):223–315.
be diligent with respect to documenting grain size distri- [2] Wang YM, Ma E. Strain hardening, strain rate sensitivity, and
bution, but also need to develop methodology for pre- ductility of nanostructured metals. Mater Sci Eng A 2004;375–
paring nanocrystalline metals with better controlled 77:46–52.
[3] Gertsman VY, Hoffman M, Gleiter H, Birringer R. The study of
grain sizes that allow for more definitive discrimination grain-size dependence of yield stress of copper for a wide grain-
amongst the various potential mechanisms. size range. Acta Metall Mater 1994;42:3539.
The comparison between the simulated tensile stress [4] Ebrahimi F, Bourne GR, Kelly MS, Matthews TE. Mechanical
vs. strain curves is seen to be quite good except that properties of nanostructured nickel produced by electrodeposit-
the experimental curves display limited ductility and ter- ion. Nanostruct Mater 1999;11(3):343–50.
[5] Van Swygenhoven H, Spaczer M, Caro A. Microscopic descrip-
minate after quite modest plastic strains. This is clear in tion of plasticity in computer generated metallic nanophase
the curves shown in Fig. 3 which are repeated on the left samples: a comparison between Cu and Ni. Acta Mater
side of Figs. 11 and 12. The experimentally observed fail- 1999;47:3117–26.
ure mode involved abrupt necking and was ductile in that [6] Van Swygenhoven H, Derlet PM. Grain-boundary sliding in
the fracture surface was dimpled. These results will be de- nanocrystalline fcc metals. Phys Rev B 2001;64. 224105-1,5.
[7] Van Swygenhoven H, Caro A, Farkas D. A molecular dynamics
scribed in further detail in Part II. Similar curves in the study of polycrystalline fcc metals at the nanoscale: grain
literature likewise often display limited elongation to boundary structure and its influence on plastic deformation.
failure owing to the very high strength and low levels Mater Sci Eng A 2001;309:440–4.
4838 B. Zhu et al. / Acta Materialia 53 (2005) 4825–4838

[8] Hasnaoui A, Van Swygenhoven H, Derlet PM. On non-equilib- [15] Asaro RJ, Needelman A. Texture development and strain
rium grain boundaries and their effect on thermal and mechanical hardening in rate dependent polycrystals. Acta Metall
behavior: a molecular dynamics computer simulation. Acta Mater 1985;33:923–53.
2002;50:3927–39. [16] Dalla Torre F, Van Swygenhoven H, Victoria M. Nanocrystalline
[9] Hirth JP, Lothe J. Theory of dislocations. New York (NY): Mc- electrodeposited Ni: microstructure and tensile properties. Acta
Graw-Hill; 1968. Mater 2002;50:3957–70.
[10] Kumar KS, Suresh S, Chisholm MF, Horton JA, Wang P. [17] Conrad H, Narayan J. On the grain size softening in nanocrys-
Deformation of electrodeposited nanocrystalline nickel. Acta talline materials. Scr Mater 2000;42:1025–30.
Mater 2003;51:387–405. [18] Lee EH. Elastic–plastic deformation at finite strains. J Appl Mech
[11] Kumar KS, Van Swygenhoven H, Suresh S. Mechanical behavior 1969;36:1.
of nanocrystalline metals and alloys. Acta Mater [19] Asaro RJ, Rice JR. Strain localization in ductile single crystals. J
2003;51(19):5743–74. Mech Phys Solid 1977;25:309–38.
[12] Asaro RJ, Suresh S. Mechanistic models for the activation volume [20] Evans M, Hastings N, Peacock B. Statistical distributions. 3rd
and rate sensitivity in metals with nanocrystalline grains and ed. New York (NY): Wiley; 2000.
nano-scale twins. Acta Mater 2005;53(12):3369–82. [21] Krill CE, Birringer R. Estimating grain-size distributions in
[13] Asaro RJ, Krysl P, Kad B. Deformation mechanism transitions in nanocrystalline materials from X-ray diffraction profile analysis.
nanoscale FCC metals. Philos Mag Lett 2003;83(12):733–43. Philos Mag A 1998;77(3):621–40.
[14] Rice JR. Dislocation nucleation from a crack tip – an analysis [22] Nieman GW, Weertman JR, Siegel RW. Mechanical
based on the Peierls concept. J Mech Phys Solids behavior of nanostructured metals. Nanostruct Mater
1992;40(2):239–71. 1992;1:185–90.

You might also like