You are on page 1of 13

Materials Science & Engineering A 689 (2017) 176–188

Contents lists available at ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

Constitutive modeling of elastoplasticity in spark-plasma sintered metal- MARK


matrix nanocomposites

M.U. Siddiqui, A.F.M. Arif , Saheb Nouari, M. Shahzeb Khan
Mechanical Engineering Department, King Fahd University of Petroleum & Minerals, Dhahran, Saudi Arabia

A R T I C L E I N F O A BS T RAC T

Keywords: A methodology to model the elastoplastic constitutive response of spark plasma sintered metal matrix
Composites particulate nanocomposites is presented. The formulated methodology incorporates the effect of Hall-Petch
Sintering strengthening, Orowan strengthening, dislocation density strengthening, load transfer strengthening and
Grain growth porosity on the elastoplastic response of metal matrix nanocomposite using a mean-field homogenization
Plasticity
approach used in conjunction with a plasticity hardening function for metal matrix that is dependent on the
metal matrix crystallite size, inclusion particle size and inclusion volume fraction. The effect of sintering process
parameters is incorporated into the model using a grain growth model that can estimate the average matrix
crystallite size as a function of sintering time, sintering temperature, inclusion size and inclusion volume
fraction. The formulated methodology has been validated against experimentally measured crystallite sizes and
stress-strain responses of ball-milled and spark plasma sintered Al-Al2O3 nanocomposite samples synthesized
during the current work. Additional validation of the methodology against the experimental stress-strain
response of nanostructured aluminum reported in the literature has also been carried out.

1. Introduction reported by Hall [18] and Petch [19] in the early 1950s. Orowan
strengthening [7,17,20] occurs due to the interaction of the reinforce-
Metal matrix nanocomposites (MMNCs) are very promising mate- ment particles with dislocations causing dislocation bowing leading to
rials that are seeing significant interest from investigators worldwide Orowan loops. Finally, the property mismatch [7,9,17,20] causes the
due to their suitability to a wide range of applications. MMNCs show generation of geometrically necessary dislocations (GNDs) during
higher strength and stiffness compared to conventional composites straining. The overall strengthening of the nanocomposites is a
while maintaining the ductility of the metals. The use of a variety of combination of the above-mentioned strengthening mechanisms.
matrix metals having several types of nanoscale inclusions has been Several models to estimate the strength of nanocomposites and
reported in the literature. The most commonly reported metals nanostructured metals, in general, have been reported in the literature
matrices include Aluminum [1–6], Magnesium [7–9], Copper [10] [7,9,20–24]. Dunstan and Bushby [21] have presented a compressive
and their alloys. For inclusions, the use of ceramic compounds such as analysis of various equations that can be used to model Hall-Petch
SiC [2,11,12], Al2O3 [4–6,10,13], and CNTs [8,14–16] has been effect using several data sets from literature. Khan et al. [22], using
extensively reported. experimental results for Al and Fe, showed that as crystallite size is
MMNCs show significant enhancement of mechanical properties at reduced, the parameters of Hall-Petch equation change. They also
very low volume fractions of reinforcements [17]. This is because formulated a crystallite size, strain rate, and temperature dependent
several new strengthening mechanisms come in play as the reinforce- hardening law for nano-crystalline metals. An improved version of the
ment size is reduced to the nanometer scale. These include the load model was presented by Farrokh and Khan [24]. For estimating the
transfer effect, Hall-Petch strengthening, Orowan strengthening and yield strength of metal matrix nanocomposites, Zhang and Chen [9]
property mismatch. Because of the load transfer effect, the strength of formulated a model that incorporated the effects of load transfer
the soft matrix material is enhanced when the load is transferred to the strengthening, Orowan strengthening and dislocation density strength-
stiffer and harder reinforcement. Hall-Petch strengthening occurs due ening mechanisms. Sanaty-Zadeh [20] compared Zhang and Chen's
to the interaction of dislocations with the grain boundaries. The fact model with a modified Clyne model and found out that Zhang and
that polycrystalline materials are stronger than single crystals was first Chen's model under-predicted the yield strength of nanocomposites


Correspondence to: 1467 KFUPM, Dhahran 31261, Saudi Arabia.
E-mail address: afmarif@kfupm.edu.sa (A.F.M. Arif).

http://dx.doi.org/10.1016/j.msea.2017.02.051
Received 14 November 2016; Received in revised form 14 February 2017; Accepted 15 February 2017
Available online 17 February 2017
0921-5093/ © 2017 Elsevier B.V. All rights reserved.
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

because it did not take into account the Hall-Petch effect. Mirza and (1.5 wt%). Spark plasma sintering machine (FCT system, Germany),
Chen [7] presented an improved model that also took into account the model HP D 5, was used to sinter the as-received aluminum powder
Hall-Petch effect. and the mechanically alloyed pure aluminum and the Al-10 vol% Al2O3
In the current work, a methodology to model the elastoplastic powders. Disc-shaped specimens were prepared using a graphite die of
response of spark plasma sintered metal matrix particulate nanocom- 20 mm diameter. A compaction pressure of 50 MPa and a heating rate
posites is presented. The methodology comprises of two main parts: a of 200 °C/min were used in all sintering experiments. Aluminum as-
model for estimation of average matrix crystallite size in spark plasma received and milled for 24 h, and Al-10 vol% Al2O3 milled for 24 h were
sintered metal matrix composites and elastoplasticity constitutive sintered at a temperature of 550 °C for holding time of 20 min. More
model for the composite based on matrix crystallite size and inclusion details on the synthesis of Al-Al2O3 nanocomposites using mechanical
particle size. The elastoplasticity constitutive model can take into alloying and spark plasma sintering were reported elsewhere [25].
account the effects of matrix crystallite size, inclusion particle size
and inclusion and porosity volume fractions on the elastoplastic stress- 2.3. Microstructure characterization
strain response of metal matrix nanocomposites. The grain growth
model can be used to calculate the average crystallite size of the metal- Microstructural characterization of the ball-milled powder and
matrix as a function of sintering time, sintering temperature, inclusion spark plasma sintered samples was carried out using a field emission
volume fraction and inclusion size. The formulated methodology has scanning electron microscope (FE-SEM) equipped with energy dis-
been validated against experimentally measured crystallite sizes and persive spectroscopy (EDS). X-ray diffraction analysis with Cu radia-
stress-strain responses of ball-milled and spark plasma sintered Al- tion (wavelength λ = 0.15405 nm) was carried out to characterize the
Al2O3 nanocomposite samples synthesized during the current work. crystallite size of the aluminum matrix. The bulk density of the
Additional validation of the methodology against the experimental consolidated samples was measured according to the Archimedes
stress-strain response of nanostructured aluminum reported in the principle using Metler Toledo balance density determination KIT
literature has also been carried out. model AG285.

2. Experimental work 2.4. Mechanical characterization

2.1. Materials Digital microhardness tester (Buehler, USA) was used to measure
the microhardness of the prepared materials. The obtained hardness
Aluminum powder, 99.88% purity supplied by Aluminum Powder values were the average of 12 readings. Conditions of a load of 100 gf
CO. LTD, and α-Al2O3 (with an average particle size of 150 nm), and a time of 12 s were maintained in all measurements. Compression
99.85% purity, supplied by ChemPUR Germany, were used in this tests were performed according to ASTM E9-89a standard using
investigation. Fig. 1 shows the micrographs of Al and Al2O3 powders. Instron universal testing machine model 3367. Multiple standard
More information on the chemical composition and particle size cylindrical specimens were machined from each sintered sample
distribution of the aluminum raw powder was reported elsewhere [25]. (25 mm diameter and 12 mm thickness) using wire EDM to final
dimensions of diameter = 6 mm ( ± 0.02 mm) and length =12 mm
2.2. Methodology and process parameters ( ± 0.02 mm). Compression tests were carried out using a compression
rate of 0.5 mm/min, and the compression load was applied parallel to
Mechanical alloying was used to synthesize Al-10 vol% Al2O3 the direction of the uniaxial spark plasma sintering pressure.
nanocomposite powder. Pure aluminum, as a reference material, was
milled for 24 h. A planetary ball mill (Fritsch Pulverisette, P5, Idar- 3. Modeling methodology
Oberstein, Germany) was used to perform the milling experiments. The
powders were milled in argon inert gas to avoid oxidation. In all The elastoplastic response of metal matrix nanocomposites depends
experiments, the ball-to-powder weight ratio was maintained at 10:1 on the intrinsic properties of the matrix and inclusions and on the
and the speed was maintained at 200 rpm. Stainless steel vials (250 mL interactions between the metal matrix and inclusions particles. The
in volume) and balls (10 mm in diameter) were used. The sticking of constitutive behavior of the metal matrix itself is a function of average
the powder to milling balls and vials was minimized using stearic acid crystallite size of the matrix and the size and properties of the inclusion

Fig. 1. (a) FE-SEM micrograph of Al powder (b) TEM micrograph of Al2O3 powder.

177
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

particles. In general, any modeling approach for predicting the Amat = [(1 − φinc ) I4 + φinc Ba]−1
elastoplastic response of nanocomposites must relate the microstruc- Ainc = Ba : Amat
tural features to the bulk properties of the composite. −1
Ba = [Ι4 + S : Cmat
−1
, alg (Cinc, alg − Cmat , alg )] (2)
In the current work, the overall elastoplastic response of the
nanocomposite was estimated using mean-field homogenization ap- where I4 is the 4th order tensor identity, S is the Eshelby tensor [27]
proach presented in Section 3.1. The effect of composite microstructure and Cmat, alg and Cinc, alg are the algorithmic tangent moduli of the
on the constitutive behavior of the metal matrix is presented in Section matrix and inclusion for the current strain increment.
3.2. In Section 3.3, models to estimate the effect of ball milling and Using the matrix and inclusion strain increments, the stresses in
sintering parameters on the microstructure of composite are presented. the matrix and inclusion and the corresponding tangent moduli were
estimated using the constitutive models of the matrix and inclusion.
Finally, the effective stress and tangent modulus of the pseudo-grain
3.1. Elastoplastic response of metal-matrix nanocomposites were estimated using Eqs. (3) and (4) respectively.
σeff , i = φinc σinc + (1 − φinc ) σmat (3)
In the current work, the incremental mean-field modeling approach
of Doghri and Tinel [26] who proposed a two-step mean-field homo- Ceff , alg, i = φinc Cinc, alg: Ainc + (1 − φinc ) Cmat , alg: Amat (4)
genization approach for modeling materials with multiple heterogene-
ities approach was used. For the estimation of elastoplastic response, Since the strain increments and the algorithmic tangent moduli of
the mean-field homogenization methodology was incrementally ap- the matrix and inclusion are implicitly related, an iterative scheme was
plied to the nanocomposite during a numerical uniaxial compression used to solve for the stresses and strains in the pseudo grains. At the
test in order to determine the stresses and the tangent modulus of the beginning of the iterations, the inclusion strain increment was assumed
composite at each strain increment. to be equal to the macroscale strain increment. Using this value, the
In order to model composites with multiple inclusions (or porosity), entire process was completed once. At the end of the step, the inclusion
the first step of the homogenization approach was to divide the strain increment was recalculated using the updated parameters. The
microstructure into pseudo grains each containing only one type of iterative process was stopped once the difference between the inclusion
inclusion (or porosity). Mean-field homogenization was carried for all strain increment at the beginning and the end of an iteration was
pseudo grains to determine the effective stresses and tangent modulus within tolerance. The iterative scheme is shown in Fig. 3.
in each grain. Finally, homogenization was carried out across all grains Once the effective stresses and tangent moduli of all pseudo grains
to determine the overall effective stresses and tangent modulus of the had been calculated, the macroscale stresses and tangent modulus were
heterogeneous material. The two-step homogenization approach is calculated using Voigt homogenization across all the pseudo grains
shown graphically in Fig. 2. using Eqs. (5) and (6).
For a microstructure containing N types of inclusions i having N
φi
volume fractions φi , the microstructure was divided into N pseudo σM = ∑ σeff , i
1 − φ0 (5)
grains with each pseudo grain having only one type of inclusion. Within i =1

each pseudo grain, the inclusion volume fraction φinc was set equal to N
φi
(1 − φ0 ), where φ0 is the volume fraction of the matrix. Ca lg , M = ∑ Ceff , alg, i
Macroscale strain increments were applied to each pseudo-grain as i =1
1 − φ0 (6)
the boundary condition. Using the mean-field strain localization
relations, the strain increments in the inclusion and matrix were
3.2. Constitutive modeling of metal matrix
estimated using,

Δεinc = Ainc ΔεM The properties of the metal matrix are a function of the average
Δεmat = (ΔεM − φinc Δεinc )/(1 − φinc ) (1) crystallite size of the matrix as well as the size of the inclusion particles
embedded in it. The strengthening of the metal matrix with a reduction
where φinc is the volume fraction of inclusion phase, Amat and Ainc in its grain size is referred to as Hall-Petch effect. The strengthening
are the matrix and inclusion strain localization tensors defined by, due to hard nanometer-sized inclusions embedded in the matrix occurs

Fig. 2. Two-step mean-field homogenization approach.

178
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Fig. 3. Effective stress and tangent modulus update algorithm for mean-field homogenization.

through the Orowan strengthening and dislocation density strengthen- Bushby [28] have also argued that Eq. (15) are more closely related to
ing mechanisms [7,17,23,28]. the physical mechanisms acting to increase the strength of the material.
The elastoplastic constitutive behavior of the metal matrix was
σy = σy 0 + k σ / d x
defined by Eqs. (7)–(9).
ε y = ε y 0 + kε / d x (14)
σ = C: εel (7)
ln(d / a 0 )
εel = ε − εpl (8) σy = σy 0 + k σ
d /a 0
1 ln(d / a 0 )
ε= [(∇u )T + ∇u + (∇u )T ∇u] εy = εy 0 + kε
2 (9) d /a 0 (15)

where σ is the true stress, C is the elasticity tensor and ε, εel and εpl are where d is the average grain size, a0 is lattice parameter of the matrix
the total, elastic and plastic strain tensors. metal, σy and εy are the yield stress and yield strain and σy0, εy0, kσ,
The rate-independent plasticity of the matrix material was modeled kε and x are material parameters.
using a voce-type hardening model. The model is defined by Eqs. (10)– It has been previously reported that as the grain size is reduced, the
(13). simple relationship of Eq. (14) does not hold [22,29,30]. Khan et al.
[22] showed that below a threshold value, the parameters of the Hall-
∂F
εṗ = εṗ , eff Petch equation change. In Fig. 4, Eqs. (14) and (15) have been fitted to
∂S (10)
two different data sets from literature [31,32]. The fitted relationships
F = σmises − σy (11) are also shown in the figure. The data in Fig. 4 represents average grain
sizes from 330 nm to 41 µm. The bilinear relationship is clearly visible
σy = σ0 + (σA − σ0 )(1 − exp(−εp, eff /ε0 )) + mεp, eff (12) in Fig. 4(a) in which the parameters for Eq. (14) change at a grain size
of 1.6 µm and 4.4 µm for datasets 1 and 2 respectively. Eq. (15), on the
εṗ , eff ≥ 0, F (σ , σys ) ≤ 0, εṗ , eff F = 0 (13) other hand, is able to capture the entire range of crystallite sizes using
where εp is the plastic strain tensor, εp, eff is the von Mises effective one set of parameters.
plastic strain, σmises is the von Mises stress, σy0 and σy are the initial and
current yield stress and σA , m and ε0 are parameters used in the 3.2.2. Effect of nanometer-sized inclusions
hardening law. Orowan strengthening of the metal matrix due to the presence of
nanometer-sized inclusions was modeled using Eq. (16) [7,17,20].
3.2.1. Effect of matrix crystallite size 0.13bmat Gmat ⎛ d ⎞
The Hall-Petch effect is most commonly represented using either Δσy, OR = ln ⎜ inc ⎟
dinc ( 3 1/2φinc − 1) ⎝ 2bmat ⎠ (16)
the widely used Eq. (14) or Eq. (15) [21,28]. In Eq. (14), the parameter
x can range from 0 to 1. Dunstan and Bushby [28] have suggested that where Δσy, OR is the increment in yield strength of the metal matrix due
x=1 provides better estimates. On the other hand, Khan et al. [22] and to Orowan strengthening, bmat is the burger's length of the matrix,
Farrokh and Khan [24] have suggested using x=0.5. Dunstan and Gmat is the shear modulus of the matrix and dinc is the inclusion size.

179
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Fig. 4. Modeling Hall-Petch effect using (a) Eq. (14) and (b) Eq. (15) for Aluminum using data from Tsuji et al. [31] (Dataset 1) and Yu et al. [32] (Dataset 2).

The enhancement due to elastic modulus (EM) and coefficient of voce hardening law of metals. The Voce hardening law, shown in Eq.
thermal expansion (CTE) mismatch was modeled using Eq. (17) (12), has four parameters, σ0, σA, m and ε0 that control the shape of
[7,9,17,20]. the hardening curve in which the parameter σ0 is the initial yield stress
⎛ ⎞ of the material at the onset of yielding.
Δσy, X = MβGmat bmat ⎜ ρ X ⎟ In the current work, the experimental data of Khan et al. [22] and
⎝ ⎠ Farrokh and Khan [24] was used to study the effect of matrix crystallite
CTE
AΔαΔTφinc size on the hardening law parameters. For both these datasets, it was
ρ =
bmat dinc (1 − φinc ) reported that the strength of the material shows a piecewise linear
6φinc relationship with d−0.5 where d is the average crystallite size of the
EM
ρ = ε
πdinc (17) material [22,24]. It is important to note here that Khan et al. [22]
reported crystallite sizes of ball-milled aluminum powder and not the
Where X can be either EM or CTE, ρCTE and ρ EM are the dislocation sintered samples. Therefore, their results cannot be directly correlated
densities generated due to CTE and EM mismatch respectively, Δα is with other data sets. The variation of hardening model parameters is
the CTE mismatch between matrix and inclusion and ΔT is the shown in Figs. 5 and 6 for Khan et al.’s data and Farrokh and Khan's
difference between processing and testing temperatures, M is the data respectively. The results of both datasets revealed that three of the
Taylor's factor (≈ 3 ), β(=1.15) is a constant and A is a geometric four parameters, σ0, σA and m, showed bilinear variation with d−0.5.
factor ranging from 10 to 12 depending on the shape of inclusion The fourth parameter, ε0, showed deviation from the bilinear behavior.
particles. Its value is 12 for spherical inclusions. But, as will be shown in the results, a bilinear approximation of ε0 still
provides good agreement with experimental data.
3.2.3. Overall strength of the metal matrix Based on the fact that all parameters in the hardening function
The overall strengthening of the metal matrix is a combination of all show a dependence of matrix crystallite size, the hardening function of
of the above mechanisms. Several ways for the determination of the the matrix was modified as,
overall strengthening effect have been reported in the literature
[7,9,17,20]. These range from models that consider the strengthening σ (εp, eff ) = σ0 (d ) + {σA (d ) − σ0 (d )}[1 − exp{−εp, eff /ε0 (d )}] + m (d ) εp, eff
mechanisms to be additive to models which consider the synergic (19)
effects of the various mechanisms. Sanaty-Zadeh [20] carried out a where the crystallite size dependence of the model parameters is
comparison of these various approached and found that the Clyne defined by Hall-Petch type Eq. (20).
model [20,33], shown in Eq. (18), provided the most accurate predic-
tions. Therefore, it was used in the current work to determine the σ0 (d ) = σ0* + kσ 0 / d
overall strengthening of the matrix. It should be noted that load σA (d ) = σA* + kσA / d
transfer strengthening and porosity do not affect the constitutive
m (d ) = m 0* + k m / d
behavior of the metal matrix and are directly taken care of during
mean-field homogenization. ε0 (d ) = ε0* + kε0 / d (20)

Δσy = 2
ΔσOR 2
+ ΔσCTE 2
+ ΔσEM 2
+ ΔσHall To include the effect of nanometer-sized inclusions into the hard-
− Petch (18)
ening function, the Clyne model, defined by Eq. (18), was applied to
parameters σ0 and σA of the hardening function since these two
3.2.4. Matrix crystallite size and inclusion particle size dependent parameters represent stresses. Eq. (21) define the two parameters as a
hardening of metal matrix function of matrix crystallite size, inclusion particle size and inclusion
The hardening behavior in metals can be modeled using several volume fraction.
types of models. Widely used models include bilinear hardening,
σ0 (d , dinc, φinc ) = σ0, large grains + Δσ0 (d , dinc, φinc )
power-law hardening, and Voce hardening law. In the current work,
an approach was developed to incorporate the grain size effect into the σA (d , dinc, φinc ) = σA, large grains + ΔσA (d , dinc, φinc ) (21)

180
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Fig. 5. Crystallite size dependence of Voce hardening law parameters for Aluminum using data from Khan et al. [22].

Fig. 6. Crystallite size dependence of Voce hardening law parameters for Aluminum
using data from Farrokh and Khan [24].
Fig. 7. Grain size reduction of pure Al during ball milling.

d milled = a1 exp(a2 t ) + a3 exp(a4 t ) (22)


3.3. Effect of process parameters on matrix crystallite size
The grain growth during the sintering process was modeled using
The properties of particulate nanocomposites are a strong function the proposed Eq. (23). In Eq. (23), the first term in the first parenthesis
of the crystallite size of the matrix material. This crystallite size is, in takes into account the effect of grain size on the grain growth rate and
turn, a function of the process parameters used in the synthesis of the the second term in the first parenthesis takes into account the effect of
nanocomposite. A commonly used approach in the fabrication of Zener pinning effect [3,34,35] due to nanoscale inclusion within the
particulate metal matrix nanocomposites involves the step of ball- metal matrix. This effect is directly proportional to the inclusion
milling the matrix and inclusion powder together followed by the step volume fraction and inversely proportional to the inclusion size [36].
of high-temperature consolidation such as hot pressing, hot extrusion Eq. (23) is similar to the differential form of grain growth models used
or spark plasma sintering. The process parameters of both these steps by Zhou et al. [37] and Ye et al. [38] with the additional effect of Zener
determine the final crystallite size of the matrix. pinning.
The grain size reduction during ball milling can be easily captured
by a double exponential function as shown in Eq. (22). Data of grain dG ⎛k φ ⎞ ⎛ −QG ⎞
= ⎜ 0n − kinc inc ⎟ exp ⎜ ⎟ , G (0) = d milled
size reduction of Aluminum powder during ball milling from two dt ⎝G dinc ⎠ ⎝ RT ⎠ (23)
different sources was used in the current work. Khan et al. [22] carried
out high energy ball milling at 100 rpm with a ball to powder ratio of where G is the average grain size of the matrix during sintering, dmilled
20:1 to make nanostructured Al. In the experiments during current is the initial grain size, T is the sintering temperature, R is the universal
work, ball milling of Al powder was carried out at 200 rpm with a ball gas constant and k0, kinc, n and QG are parameters describing the
to powder ratio of 10:1. As shown in Fig. 7, Eq. (22) can capture a grain growth behavior of the material.
reduction as well as an increase in grain size. The parameters a1, a2, a3 Eqs. (22) and (23) can be used to determine the average crystallite
and a4 will depend on the ball-milling process parameters as well as size of ball milled and sintered samples. The transition of matrix metal
the properties of the powder being milled. crystallite size from as-received powder to sintered sample is shown

181
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Fig. 8. Transition of matrix metal crystallite size from as-received powder to sintered sample.

Table 1
Summary of mechanical properties of sintered samples.

Sample Porosity Crystallite size Crystallite size Vickers Yield strength (0.2% Compressive Strength
fraction (%) (milled powder) (nm) (sintered sample) hardness (MPa) offset) (MPa) (MPa)
(nm)

Al 0.20 298 366 326.3 74.33 ± 20.97 204.43 ± 15.86


(as received)

Al 2.98 62 108 421.0 166.90–224.01 371.69 ± 69.15


(24 h milled)

Al-10%Al2O3 (24 h 3.43 33 53 1309.7 311.40 ± 44.24 432.87 ± 91.01


milled)

graphically in Fig. 8. yield strength to 166.90–224.01 MPa. The addition of 10 vol% Al2O3
and milling for 24 h increased the yield strength to 311.40 ±
44.24 MPa. The compressive strength of the monolithic un-milled
4. Results and discussion aluminum was 204.43 ± 15.86 MPa. Milling for 24 h increased the
strength to 371.69 ± 69.15 MPa. The addition of 10 vol% Al2O3 and
4.1. Measured properties milling for 24 h increased the strength to 432.87 ± 91.01 MPa. The
variation in the yield and compressive strengths of the sintered samples
A summary of properties of milled powders and sintered samples is highlight the distribution of properties within the sintered samples.
provided in Table 1. Mechanical alloying of pure aluminum for 24 h This variation was determined by testing multiple cylindrical specimen
decreased its crystallite size from 298 to 62.24 nm. As for the cut from each sintered sample. The stress-strain response of the
composite, milling for 24 h and the addition of 10 vol% of Al2O3 sintered samples under compression is presented in Fig. 10 in which
nanoparticles reduced the crystallite size of the α-aluminum phase to the average stress-strain response of each sintered sample (calculated
32.69 nm. Sintering of the as-received aluminum, the mechanically using compression tests on multiple cylindrical specimens from the
alloyed aluminum, and the mechanically alloyed Al- 10 vol% of Al2O3 same sample) is presented along with variation in the stress-strain
nanocomposite increased the crystallite size of the α-aluminum phase response. The increase in hardness, yield strength and compressive
to 366, 108.18, and 53.4 nm, respectively. This indicates that Al2O3 strength can be attributed to the small grain size of the matrix (Hall-
nanoparticles, on the one hand, acted as grinding medium and Petch theory), the presence of nanoparticles (Orowan strengthening),
enhanced the milling effect, and on the other hand, acted as grain an increase in dislocations’ density, load transfer from the matrix to the
growth inhibitors during sintering. reinforcement, and strain gradient [17,39,40].
The as-received pure aluminum and pure aluminum milled for 24 h
had relative density values of 98.8% and 97.02%, respectively. Milling
for 24 h and addition of 10 vol% Al2O3 reduced the relative density of 4.2. Modeling of pure nanostructured aluminum
96.57%. This could be due to the fact that sintering of nano-composite
powders become more difficult at higher fractions of ceramic nano- A comparison of the 0.2% yield stress of pure aluminum samples
particles. Fig. 9 shows optical micrographs of pure Al samples sintered sintered in the current work with predicted values using Hall-Petch
using as-received and 24 h milled Al powders. relationships (14) and (15) is presented in Table 2. The predicted yield
The sintered as-received aluminum and the 24 h mechanically strains are multiplied with the experimentally determined elastic
alloyed aluminum had Vickers hardness values of 326.3 and moduli of the sintered samples to calculate yield stresses. In the work
421 MPa, respectively. The Al-10 vol% Al2O3 composite had a Vickers of Farrokh and Khan [24], the sintered samples had a different elastic
hardness of 1309.7 MPa. The monolithic un-milled aluminum had a modulus than current work. Therefore, the Hall-Petch relationship (14)
yield strength of 74.33 ± 20.97 MPa. Milling for 24 h increased the using their data was scaled to the modulus of elasticity of samples in

Fig. 9. Optical micrographs of Al samples sintered from (a) as-received Al powder (b) 24 h milled Al powder.

182
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

agreement with experimental data.

4.3. Modeling of elastoplastic response of aluminum matrix-alumina


inclusion nanocomposites

Using the stress-strain curves of sintered aluminum samples made


using as-received aluminum powder and 24 h milled aluminum
powder, the crystallite size dependence of hardening law parameters
was determined. The variation of the hardening law parameters with
the inverse of the square root of crystallite size is shown in Fig. 12 in
which the fitted Hall-Petch type relationships for all parameters is also
shown. A comparison of the modeled hardening behavior of two pure
Al sintered samples is plotted in Fig. 13 along with the experimental
hardening behavior of the samples.
In order to model the stress-strain response of the Al-10% Al2O3
sample, the parameters m and ε0 were scaled using the Hall-Petch type
relationships shown in Fig. 12 for the average crystallite of Al matrix in
the sintered composite (53.4 nm). For determining parameters so and
sA, the Clyne model represented by Eq. (18) was used. Three
strengthening mechanisms were incorporated into the parameters:
Hall-Petch effect, Orowan strengthening and dislocation density effect
Fig. 10. Stress-strain response of sintered samples under compression. due to the coefficient of thermal expansion mismatch. The Clyne model
calculates the overall change in the two parameters so and sA. This
Table 2 difference needs to be added to the parameters of the coarse-grained
0.2% Yield stress prediction for spark plasma sintered samples.
material in order to determine the new parameters. Since no coarse-
Crystallite size grained samples were made in the current work, the equations for so
and sA were extrapolated to determine the parameter values from
366 nm 106 nm which the change was calculated. The parameter values at an average
crystallite size of 549.1 nm were used since at this value, so and sA have
Method Yield Yield stress Yield Yield
strain [1] [MPa] strain [1] stress
an equal value of 46.99 MPa. At higher crystallite sizes, so comes out
[MPa] greater than sA which is physically not possible.
For an average inclusion particle size of 150 nm, inclusion volume
Experimental – 2.29×10−3 74.33 ± 20.97 6.17×10−3 166.90– fraction equal to 10% and a shear modulus of 10.74 GPa (calculated
Current work 224.01
−3 −3 using E = 29.04 GPa and ν = 0.35), the contributions of Orowan
Eq. (14) - Dataset 3.50×10 82.46 7.06×10 205.02
1 [31] strengthening and dislocation density strengthening were found out
Eq. (14) - Dataset 4.00×10−3 94.24 8.44×10−3 245.10 to be equal to 9.33 MPa and 103.98 MPa. The contribution of Hall-
2 [32] Petch effect to so and sA for an average crystallite size of 53.4 nm were
Eq. (15) - Dataset 3.65×10−3 85.99 9.64×10−3 279.82
163.26 MPa and 341.23 MPa respectively. By applying the Clyne
1 [31]
Eq. (15) - Dataset 3.88×10 −3
91.38 9.78×10 −3
283.97
model, parameters so and sA were found out be equal to 240.26 MPa
2 [32] and 403.55 MPa respectively. Parameter m and εo were determined to
Eq. (14) - Farrokh 2.43×10−3 56.32 4.23×10−3 120.83 be equal to 2189.32 MPa and 1.68×10−3 respectively. The modeled
& Khan's data stress-strain responses of all sintered samples are presented in Fig. 14.
In order to analyze the relative contributions of Hall-Petch effect
and Orowan and dislocation density strengthening, parametric studies
current work. It can be seen from the results that Eqs. (14) and (15)
were carried out to study the effects of inclusion volume fraction,
using datasets 1 and 2 both provided excellent predictions for spark
inclusion particle size and matrix crystallite size on the stress-strain
plasma sintered aluminum samples of as-received Al powder
response of the Al-Al2O3 nanocomposite. Varying inclusion volume
(d=366 nm). For the 24 h ball-milled and spark plasma sintered
fraction and particle size affects the contributions of Orowan strength-
aluminum sample (d=108 nm), Eq. (14) provided better estimates
ening and dislocation density strengthening while varying matrix
than Eq. (15). This is because a crystallite size of 106 nm is smaller
crystallite size affects the contribution of Hall-Petch effect.
than those in datasets 1 and 2. The slope of Eq. (15) increases as d/a0
The results of the parametric studies are shown in Figs. 15 and 16.
increases which result in an overestimation of yield stress. Eq. (14)
The results of the parametric studies show that the biggest contributing
using Farrokh and Khan's data underestimated the yield stresses.
factor to the enhancement of yield strength and flow stress of Al-Al2O3
The complete elastoplastic response of pure Al samples under
nanocomposite is the Hall-Petch effect. As the crystallite size was
compression was modeled using the constitutive model defined in
reduced from 300 nm to 50 nm, the 0.2% yield strength of the
Section 3.2 for the pure Al samples sintered by Khan et al. [22] and
composite increased from 141 MPa to 373 MPa for an inclusion
Farrokh and Khan [24]. The parameters of the crystallite size depen-
volume fraction of 5% and inclusion particle size equal to 150 nm.
dent hardening law were taken from Figs. 5 and 6 for Khan et al.’s data
The contribution of Hall-Petch effect becomes especially high for ball-
and Farrokh and Khan's data respectively. The modeling results are
milled composites since ceramic inclusion particles act as an additional
presented in Fig. 11 along the Khan et al.’s and Farrokh and Khan's
grinding medium for the metal matrix powder reducing its crystallite
experimental results. The modeling results that the crystallite size
size.
dependent hardening law presented in the current work is able to
For inclusion, there are two parameters, volume fraction and
accurately capture the effect of crystallite size on the elastoplastic
particle size, which affect the strength of the nanocomposite.
response of nanostructured metals. The assumption of linear depen-
Increasing the inclusion volume fraction provided a relatively smaller
dence of parameter ε0 on the inverse of square root of crystallite size
improvement compared to the Hall-Petch effect. Adding 10% inclusion
did not adversely affect the model results and the results had good
improved the 0.2% yield strength from 205 MPa to 256 MPa.

183
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Fig. 11. Comparison of modeled elastoplastic response of Aluminum with (a) Khan et al.’s experiments [22] and (b) Farrokh and Khan's experiments [24].

Fig. 12. Crystallite size dependence of Voce hardening law parameters (a) s0 and sA and (b) m and ε0 for spark plasma sintered Aluminum samples.

Fig. 13. Comparison of hardening model with experimentally determined hardening Fig. 14. Comparison of modeled stress-strain response of spark plasma sintered
behavior of spark plasma sintered Aluminum. samples with experimental results.

184
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Table 3
Sample crystallite sizes used in grain growth model parameter estimation.

Sample Crystallite size [nm]

Before sintering After sintering


Pure Al – As received – 20 min sintered 298 366
Pure Al – 24 h milled – 5 min sintered 62 66
Pure Al – 24 h milled – 10 min sintered 62 87
Pure Al – 24 h milled – 15 min sintered 62 93
Pure Al – 24 h milled – 20 min sintered 62 108
Al−2% Al2O3 −24 h milled – 20 min 40 83
sintered
Al−10% Al2O3 −24 h milled – 20 min 33 53
sintered
Al−15% Al2O3 −24 h milled – 20 min 28 42
sintered

Table 4
Grain growth model parameters.

Parameter Value

k0 5 × 10−3 nm0.691/s
Fig. 15. Effect of matrix crystallite size on stress-strain response of Al-Al2O3 nanocom- kinc 14.94 × 10−3 nm2/s
posite. QG 125.12 kJ/mol
n −0.31
Reduction in inclusion particle size also resulted in improvement in
composite strength. For Al-Al2O3 nanocomposite, the improvement
was found to be more significant for particle sizes under 100 nm. When used in the parameter estimation all listed in Table 3.
the inclusion size was reduced from 150 nm to 100 nm the 0.2% yield The grain growth model was implemented using an explicit Range-
strength increases from 230 MPa to 240 MPa. For inclusion particle Kutta method [41] in MATLAB. Estimation of the model parameters
sizes 50 nm and 25 nm, the 0.2% yield strength of the composite was was carried out using the Nelder-Mead simplex search algorithm [42]
265 MPa and 306 MPa respectively. by minimizing the total root-mean-squared error in the final crystallite
size prediction for all samples. The estimated model parameters are
listed in Table 4. A comparison of the predicted crystallite sizes with
4.4. Modeling of elastoplastic response of spark plasma sintered experimentally measured sizes is shown in Fig. 17. The figure shows
alumina matrix nanocomposites that the proposed grain growth model is able to accurately model the
effects of sintering time as well as inclusion volume fraction.
In order to determine the three model parameters in Eq. (23) used Using the grain growth model in combination with the elastoplas-
for calculating grain growth, additional pure aluminum and composite ticity model, parametric studies were carried out to study the effect of
samples were made using spark plasma sintering. For 24 h milled pure sintering time and temperature on the elastoplastic response of Al-
aluminum, additional samples were made using sintering times of Al2O3 nanocomposites. The initial crystallite sizes for all modeled cases
5 min, 10 min and 15 min. Additional composite samples were made were taken from Table 3. The inclusion particle size was taken as
with 2% and 15% alumina and a ball-milling time of 24 h. The sintering 150 nm and the porosity fraction for all cases was assumed to be equal
temperature was kept constant at 550 °C for all samples. In total, five to 2%. Same porosity volume fraction was used for all cases since it has
new samples were made in addition to the three samples that have been been previously shown that same amount of porosity is obtained in the
mentioned before. The average crystallite sizes of all sintered samples sintering time range of 5–20 min [25]. The results of the parametric

Fig. 16. Effect of inclusion (a) volume fraction and (b) particle size on stress-strain response of Al-Al2O3 nanocomposite.

185
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Fig. 17. Comparison of grain growth model predictions with experimental results. (a) Effect of sintering time on grain growth in pure Al. (b) Effect of inclusion volume fraction on
matrix grain growth in Al-Al2O3 nanocomposite for a sintering time of 20 min.

Fig. 18. Effect of sintering time on the elastoplastic response of 24 h milled (a) pure Al, (b) Al-2% Al2O3 composite, (c) Al-10% Al2O3 composite and (d) Al-15% Al2O3 composite.
Sintering temperature is 823 K.

186
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

Fig. 19. Effect of sintering temperature on the elastoplastic response of 24 h milled (a) pure Al, (b) Al-2% Al2O3 composite, (c) Al-10% Al2O3 composite and (d) Al-15% Al2O3
composite. Sintering time is 20 min.

studies are shown in Figs. 18 and 19 which show the effects of sintering ening, Orowan strengthening, dislocation density strengthening, load
time and sintering temperature respectively. The results show that transfer strengthening and porosity using a mean-field homogenization
increasing the sintering temperature and time beyond what is required approach along with a plasticity hardening function for the metal
for good densification would result in a decrease in the strength of the matrix that is dependent on the metal matrix crystallite size, inclusion
material. In the current study, when the sintering time was increased particle size and inclusion and porosity volume fractions. The effect of
from 5 min to 20 min while keeping the sintering temperature constant process parameters is incorporated into the model using a grain growth
at 823 K, the 0.2% yield strength reduced from 217 MPa to 189 MPa model that can estimate the average matrix crystallite size as a function
for the 24 h milled pure Al sample, from 263 MPa to 229 MPa for 2% of sintering time, sintering temperature, inclusion size and inclusion
alumina composite, from 387 MPa to 350 MPa for 10% alumina volume fraction.
composite and from 507 MPa to 463 MPa for 15% alumina composite. The formulated methodology was validated against experimentally
An increase in the sintering temperature from 798 K to 873 K with a determined crystallite sizes and elastoplastic response of nanostruc-
sintering time of 20 min reduced the 0.2% yield strength from 174 MPa tured aluminum and aluminum matrix-alumina reinforced nanocom-
to 143 MPa for the 24 h milled pure Al sample, from 215 MPa to posite synthesized using ball milling and spark plasma sintering in the
179 MPa for 2% alumina composite, from 322 MPa to 280 MPa for current work. Additional validations of the proposed hardening func-
10% alumina composite and from 420 MPa to 374 MPa for 15% tion were carried out against the elastoplastic response of nanostruc-
alumina composite. tured aluminum taken from literature. Both validations showed that
the proposed methodology is able to accurately capture the dependence
5. Conclusions of elastoplastic response of nanostructured metals and metal-matrix
nanocomposites on the metal matrix crystallite size, inclusion particle
In the current work, a methodology to model the constitutive size and inclusion and porosity volume fractions.
response of elastoplasticity in spark plasma sintered nanostructured Using the formulated methodology, parametric studies were carried
metals and metal-matrix nanocomposites has been presented. The out by varying matrix crystallite size, inclusion particle size and
formulated methodology incorporates the effect of Hall-Petch strength- inclusion volume fraction to determine the relative contributions of

187
M.U. Siddiqui et al. Materials Science & Engineering A 689 (2017) 176–188

various strengthening mechanisms in a metal matrix nanocomposite. It composite powders, Compos. Part A Appl. Sci. Manuf. 41 (2010) 322–326. http://
dx.doi.org/10.1016/j.compositesa.2009.09.028.
is found out that major contribution to strengthening comes from the [16] H. Choi, J. Shin, B. Min, J. Park, D. Bae, Reinforcing effects of carbon nanotubes in
Hall-Petch strengthening. Reduction in inclusion particle size im- structural aluminum matrix nanocomposites, J. Mater. Res. 24 (2011) 2610–2616.
proved the strength of the nanocomposite as expected but the http://dx.doi.org/10.1557/jmr.2009.0318.
[17] R. Casati, M. Vedani, Metal matrix composites reinforced by nano-particles—a
improvement was relatively more significant below a particle size of review, Metals 4 (2014) 65–83. http://dx.doi.org/10.3390/met4010065.
100 nm. On the other hand, increasing the inclusion volume fraction [18] E.O. Hall, The deformation and ageing of mild steel: III discussion of results, Proc.
increased the strength of the nanocomposite almost linearly. Phys. Soc. Sect. B. 64 (1951) 747–753. http://dx.doi.org/10.1088/0370-1301/64/
9/303.
Parametric studies were also conducted to study the effect of [19] N.J. Petch, The cleavage strength of polycrystals, J. Iron Steel Inst. 174 (1953)
sintering time and temperature on the elastoplastic stress-strain 25–28. http://dx.doi.org/10.1007/BF01972547.
response of metal-matrix nanocomposites. From the results, it was [20] A. Sanaty-Zadeh, Comparison between current models for the strength of parti-
culate-reinforced metal matrix nanocomposites with emphasis on consideration of
concluded that the sintering time and temperature should be set at the
Hall-Petch effect, Mater. Sci. Eng. A. 531 (2012) 112–118. http://dx.doi.org/
minimum required values for good densification. Unnecessarily in- 10.1016/j.msea.2011.10.043.
creasing the sintering time and temperature results in a decrease in the [21] D.J. Dunstan, A.J. Bushby, Grain size dependence of the strength of metals: the
strength of the sintered material. Hall-Petch effect does not scale as the inverse square root of grain size, Int. J. Plast.
53 (2014) 56–65. http://dx.doi.org/10.1016/j.ijplas.2013.07.004.
[22] A.S. Khan, Y.S. Suh, X. Chen, L. Takacs, H. Zhang, Nanocrystalline aluminum and
Acknowledgements iron: mechanical behavior at quasi-static and high strain rates, and constitutive
modeling, Int. J. Plast. 22 (2006) 195–209. http://dx.doi.org/10.1016/j.ij-
plas.2004.07.008.
The authors would like to acknowledge the support provided by [23] A.S. Khan, B. Farrokh, L. Takacs, Effect of grain refinement on mechanical
King Abdulaziz City for Science and Technology (KACST) through the properties of ball-milled bulk aluminum, Mater. Sci. Eng. A. 489 (2008) 77–84.
Science and Technology Unit at King Fahd University of Petroleum & http://dx.doi.org/10.1016/j.msea.2008.01.045.
[24] B. Farrokh, A.S. Khan, Grain size, strain rate, and temperature dependence of flow
Minerals (KFUPM) for funding this work through project 12- stress in ultra-fine grained and nanocrystalline Cu and Al: synthesis, experiment,
NAN2374-04 as part of the National Science, Technology and and constitutive modeling, Int. J. Plast. 25 (2009) 715–732. http://dx.doi.org/
Innovation Plan. 10.1016/j.ijplas.2008.08.001.
[25] N. Saheb, M. Shahzeb Khan, A.S. Hakeem, Effect of processing on mechanically
alloyed and spark plasma sintered Al-Al 2 O 3 nanocomposites, J. Nanomater. 2015
References (2015) 1–13. http://dx.doi.org/10.1155/2015/609824.
[26] I. Doghri, L. Tinel, Micromechanical modeling and computation of elasto-plastic
materials reinforced with distributed-orientation fibers, Int. J. Plast. 21 (2005)
[1] R. Casati, Aluminum Matrix Composites Reinforced with Alumina, Springer,
1919–1940. http://dx.doi.org/10.1016/j.ijplas.2004.09.003.
Milano, Italy, 2016. http://dx.doi.org/10.1007/978-3-319-27732-5.
[27] J.D. Eshelby, The determination of the elastic field of an ellipsoidal inclusion, and
[2] I. Aliyu, N. Saheb, S. Hassan, N. Al-Aqeeli, Microstructure and properties of spark
related problems, Proc. R. Soc. Lond. A. Math. Phys. Sci. 241 (1957) 376–396.
plasma sintered aluminum containing 1 wt.% SiC nanoparticles, Metals 5 (2015)
[28] Y. Li, A.J. Bushby, D.J. Dunstan, The Hall-Petch effect as a manifestation of the
70–83. http://dx.doi.org/10.3390/met5010070.
general size effect, arXiv:1507.01223 [Cond-Mat.mtrl-Sci] 1–33, 2015.
[3] D.C. Van Aken, P.E. Krajewski, G.M. Vyletel, J.E. Allison, J.W. Jones,
[29] H.J. Choi, S.W. Lee, J.S. Park, D.H. Bae, Tensile behavior of bulk nanocrystalline
Recrystallization and grain growth phenomena in a particle-reinforced aluminum
aluminum synthesized by hot extrusion of ball-milled powders, Scr. Mater. 59
composite, Metall. Mater. Trans. A. 26 (1995) 1395–1405. http://dx.doi.org/
(2008) 1123–1126. http://dx.doi.org/10.1016/j.scriptamat.2008.07.030.
10.1007/BF02647590.
[30] N. Kamikawa, X. Huang, N. Tsuji, N. Hansen, Strengthening mechanisms in
[4] Y.C. Kang, S.L.I. Chan, Tensile properties of nanometric Al2O3 particulate-
nanostructured high-purity aluminium deformed to high strain and annealed, Acta
reinforced aluminum matrix composites, Mater. Chem. Phys. 85 (2004) 438–443.
Mater. 57 (2009) 4198–4208. http://dx.doi.org/10.1016/j.actamat.2009.05.017.
http://dx.doi.org/10.1016/j.matchemphys.2004.02.002.
[31] N. Tsuji, Y. Ito, Y. Saito, Y. Minamino, Strength and ductility of ultrafine grained
[5] A. Mazahery, M. Ostadshabani, Investigation on mechanical properties of nano-
aluminum and iron produced by ARB and annealing, Scr. Mater. 47 (2002)
Al2O3-reinforced aluminum matrix composites, J. Compos. Mater. 45 (2011)
893–899. http://dx.doi.org/10.1016/S1359-6462(02)00282-8.
2579–2586. http://dx.doi.org/10.1177/0021998311401111.
[32] C.Y. Yu, P.W. Kao, C.P. Chang, Transition of tensile deformation behaviors in
[6] A. Mazahery, H. Abdizadeh, H.R. Baharvandi, Development of high-performance
ultrafine-grained aluminum, Acta Mater. 53 (2005) 4019–4028. http://dx.doi.org/
A356/nano-Al2O3 composites, Mater. Sci. Eng. A. 518 (2009) 61–64. http://
10.1016/j.actamat.2005.05.005.
dx.doi.org/10.1016/j.msea.2009.04.014.
[33] D. Hull, An introduction to composite materials, Igarss 2014 (2014) 1–5. http://
[7] F. Mirza, D. Chen, A unified model for the prediction of yield strength in
dx.doi.org/10.1007/s13398-014-0173-7.2.
particulate-reinforced metal matrix nanocomposites, Materials 8 (2015)
[34] K.R. Phaneesh, A. Bhat, P. Mukherjee, K.T. Kashyap, On the Zener limit of grain
5138–5153. http://dx.doi.org/10.3390/ma8085138.
growth through 2D Monte Carlo simulation, Comput. Mater. Sci. 58 (2012)
[8] C.S. Goh, J. Wei, L.C. Lee, M. Gupta, Ductility improvement and fatigue studies in
188–191. http://dx.doi.org/10.1016/j.commatsci.2012.02.013.
Mg-CNT nanocomposites, Compos. Sci. Technol. 68 (2008) 1432–1439. http://
[35] G.N. Hassold, E. Holm, Effects of particle size on inhibited grain growth, Scr. Met.
dx.doi.org/10.1016/j.compscitech.2007.10.057.
Mater. 24 (1990) 11–13 〈http://en.scientificcommons.org/
[9] Z. Zhang, D. Chen, Consideration of Orowan strengthening effect in particulate-
854714%5Cnpapers2://publication/uuid/D1AD592A-FB73-4124-869B-
reinforced metal matrix nanocomposites: a model for predicting their yield
8C758E4B0D74〉.
strength, Scr. Mater. 54 (2006) 1321–1326. http://dx.doi.org/10.1016/j.scripta-
[36] L.Kang Suk-Joong, Sintering Densification,Grain Growth, and Microstructure
mat.2005.12.017.
(2005). http://dx.doi.org/10.1016/B978-075066385-4/50009-1.
[10] F. Shehata, A. Fathy, M. Abdelhameed, S.F. Moustafa, Preparation and properties
[37] F. Zhou, J. Lee, S. Dallek, E.J. Lavernia, High grain size stability of nanocrystalline
of Al2O3 nanoparticle reinforced copper matrix composites by in situ processing,
Al prepared by mechanical attrition, J. Mater. Res. 16 (2011) 3451–3458. http://
Mater. Des. 30 (2009) 2756–2762. http://dx.doi.org/10.1016/
dx.doi.org/10.1557/JMR.2001.0474.
j.matdes.2008.10.005.
[38] J. Ye, L. Ajdelsztajn, J.M. Schoenung, Bulk nanocrystalline aluminum 5083 alloy
[11] N. Saheb, I.K. Aliyu, S.F. Hassan, N. Al-Aqeeli, Matrix structure evolution and
fabricated by a novel technique: cryomilling and spark plasma sintering, Metall.
nanoreinforcement distribution in mechanically milled and spark plasma sintered
Mater. Trans. A Phys. Metall. Mater. Sci. 37 (2006) 2569–2579. http://dx.doi.org/
Al-SiC nanocomposites, Materials 6 (2014) 6748–6767. http://dx.doi.org/
10.1007/BF02586229.
10.3390/ma7096748.
[39] N. Chawla, Y.-L. Shen, Mechanical Behavior of Particle Reinforced Metal Matrix
[12] S. Kamrani, R. Riedel, S.M. Seyed Reihani, H.J. Kleebe, Effect of reinforcement
Composites, Adv. Eng. Mater. 3 (2001) 357–370. http://dx.doi.org/10.1002/1527-
volume fraction on the mechanical properties of Al–SiC nanocomposites produced
2648(200106)3:6 < 357::AID-ADEM357 > 3.3.CO;2-9.
by mechanical alloying and consolidation, J. Compos. Mater. 44 (2010) 313–326.
[40] M.A. Muñoz-Morris, C.G. Oca, D.G. Morris, An analysis of strengthening me-
http://dx.doi.org/10.1177/0021998309347570.
chanisms in a mechanically alloyed, oxide dispersion strengthened iron aluminide
[13] M. Rahimian, N. Ehsani, N. Parvin, H.R. Baharvandi, The effect of particle size,
intermetallic, Acta Mater. 50 (2002) 2825–2836. http://dx.doi.org/10.1016/
sintering temperature and sintering time on the properties of Al-Al2O3 composites,
S1359-6454(02)00101-5.
made by powder metallurgy, J. Mater. Process. Technol. 209 (2009) 5387–5393.
[41] G. Forsythe, M. Malcolm, C. Moler, Computer Methods for Mathematical
http://dx.doi.org/10.1016/j.jmatprotec.2009.04.007.
Computations, Prentice-Hall, New Jersey, 1977.
[14] Y. Wu, G.Y. Kim, Carbon nanotube reinforced aluminum composite fabricated by
[42] J.C. Lagarias, J.A. Reeds, M.H. Wright, P.E. Wright, Convergence properties of the
semi-solid powder processing, J. Mater. Process. Technol. 211 (2011) 1341–1347.
Nelder–Mead simplex method in low dimensions, SIAM J. Optim. 9 (1998)
http://dx.doi.org/10.1016/j.jmatprotec.2011.03.007.
112–147. http://dx.doi.org/10.1137/S1052623496303470.
[15] K. Morsi, A.M.K. Esawi, S. Lanka, A. Sayed, M. Taher, Spark plasma extrusion
(SPE) of ball-milled aluminum and carbon nanotube reinforced aluminum

188

You might also like