Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/329778708

Enhanced combination of strength and ductility in ultrafine-grained


aluminum composites reinforced with high content intragranular
nanoparticles

Article  in  Materials Science and Engineering A · December 2018


DOI: 10.1016/j.msea.2018.12.090

CITATIONS READS

0 178

10 authors, including:

Ai-bin Li Guisong Wang


Harbin Institute of Technology Harbin Institute of Technology
38 PUBLICATIONS   783 CITATIONS    59 PUBLICATIONS   673 CITATIONS   

SEE PROFILE SEE PROFILE

Xuexi Zhang Mingfang Qian


Harbin Institute of Technology Harbin Institute of Technology
85 PUBLICATIONS   1,269 CITATIONS    39 PUBLICATIONS   177 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

I am working on shape memory alloys with functional properties such as shape memory, superelasticity, magnetocaloric effect (MCE), elastocaloric effects (eCE) and
magnetic-field-induced strain (MFIS) View project

genglin View project

All content following this page was uploaded by Mingfang Qian on 03 January 2019.

The user has requested enhancement of the downloaded file.


Materials Science & Engineering A 745 (2019) 10–19

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Enhanced combination of strength and ductility in ultrafine-grained T


aluminum composites reinforced with high content intragranular
nanoparticles
⁎ ⁎
A.B. Li, G.S. Wang, X.X. Zhang , Y.Q. Li, X. Gao, H. Sun, M.F. Qian, X.P. Cui, L. Geng, G.H. Fan
National Key Laboratory of Materials Behavior and Evaluation Technology in Space Environment, School of Materials Science and Engineering, Harbin Institute of
Technology, Harbin 150001, China

A R T I C LE I N FO A B S T R A C T

Keywords: Simultaneous enhancement of the strength and ductility in metal-matrix nanocomposites with high content
Metal-matrix nanocomposites (MMNCs) nano-reinforcements is crucial for the development of novel composite materials. Here SiC/Al nanocomposites
Mechanical alloying with ultrafine Al grains and intragranular 10 vol% SiC nanoparticle were prepared by mechanical alloying and
Hot extrusion hot pressure sintering followed by hot extrusion. The results show that nanocrystalline Al grains and uniformly
Mechanical properties
distributed SiC nanoparticles (SiCnp) in the composite powders are obtained after ball-milling. The prior particle
Strengthening mechanisms
boundary (PPB) in the composite powders cannot be fully removed by sintering at 580 °C under 20 MPa, but can
be eliminated by subsequent hot extrusion at 450 °C with a ratio 25:1. The extruded nanocomposites possess
ultrafine Al grains with diameter 320 nm and intragranular SiCnp created by the dynamic recrystallizations
(DRX). Such optimal microstructure favors a significant improvement of 284% in yield strength (YS) and 259%
in ultimate tensile strength (UTS) compared with monolithic Al, and more interestingly, maintains an elongation
up to 12% due to the homogeneous nanoparticles in fine Al grains. Theoretical analysis shows that the dominant
strengthening factor is Orowan strengthening related to the high content intragranular SiCnp, with second factor
grain boundary strengthening due to ultrafine Al grains.

1. Introduction enhanced from 267 MPa to 302 MPa, and elongation was improved
from 1.61% to 9.2% [6]. For intragranular [6] and intergranular [7–9]
Metal-matrix nanocomposites (MMNCs) reinforced with nano-re- nanocomposites, the addition of a small amount of nanoparticles can
inforcements have attracted considerable attention due to their unique retain good ductility besides enhanced strength. With further increasing
mechanical and physical properties [1,2]. Generally, basic nano- of the nanoparticle content, the strength was enhanced, but ductility
composite microstructures are classified into intra-type, inter-type and was reduced [2,10–12]. Especially, when high content nanoparticles up
nano/nano type. Among them, the most important structure is the to 10 vol% were added, the high strength was usually accompanied by
intra-type nanostructure, where the dispersed nanoparticles are em- serious ductility loss [11–13]. For example, in ultrafine-grained SiC/
bedded within the matrix grains [3]. However, for such intragranular 6063 nanocomposites, when SiC content was increased from 1 vol% to
nanoparticles, the allowable nanoparticle content is relatively low, ty- 10 vol%, UTS was increased from 343 MPa to 603 MPa, but elongation
pically about 1–5 vol% [4]. Therefore, how to uniformly disperse high was reduced from 10% to 2.3% [12]. So far, little work has been re-
content nanoparticles into matrix alloy grains is a tough challenge in ported on an excellent combination of strength and ductility in MMNCs
the preparation of structural nanomaterials. reinforced with high content nanoparticles.
High strength and ductility are essential for structural material ap- Besides the effective strengthening capacity of nano-reinforcements
plications [1]. It is very interesting to use nanoparticles instead of mi- [5,14,15], decreasing the matrix grain size from micrometer to sub-
croparticles to enhance the mechanical performances [5]. Nie et al. micrometer or nanometer also favors the mechanical properties
fabricated AZ91/SiC nanocomposites and found that when 1 vol% SiC [1,12,16]. However, high strength nanocrystalline (< 100 nm) and
nanoparticle (SiCnp) with diameter 60 nm replaced 10 vol% SiC mi- ultrafine-grained (e.g. 100–1000 nm) metals often show seriously re-
croparticles with 10 µm, the ultimate tensile strength (UTS) was duced room temperature tensile ductility compared with their coarse


Corresponding authors.
E-mail addresses: xxzhang@hit.edu.cn (X.X. Zhang), ghfan@hit.edu.cn (G.H. Fan).

https://doi.org/10.1016/j.msea.2018.12.090
Received 1 September 2018; Received in revised form 17 December 2018; Accepted 18 December 2018
Available online 19 December 2018
0921-5093/ © 2018 Elsevier B.V. All rights reserved.
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

grained counterparts [17,18]. Recently, ultrafine-grained nanocompo-


sites have been fabricated through the introduction of the nanoparticles
into ultrafine-grained alloys [12,16,18–20], which promoted the strain
hardening ability and ductility [12,18,19]. However, the ultrafine-
grained nanocomposites exhibited poor ductility when the nanoparticle
content reached 10 vol% [12]. To date, the coupled effects of high
content nanoparticles and ultrafine grains on strength and ductility
have been rarely reported. Here, we have attempted to prepare 10 vol%
SiC/Al nanocomposites with ultrafine grains and intragranular nano-
particles using mechanical alloying and sintering followed by hot ex-
trusion. Their microstructure evolution, tensile properties and
strengthening mechanisms were investigated. A simultaneous en-
hancement of strength and ductility was demonstrated in the extruded
10 vol% SiC/Al nanocomposites.

2. Experimental details

Fig. 1. X-ray diffraction (XRD) patterns of 10 vol% SiC-Al nanocomposite


2.1. Fabrication of 10 vol% SiC/Al nanocomposites
powders ball-milled for different time durations 6–21 h.

Aluminum powders of 99% purity with average diameter 50 µm and


SiCnp of 99% purity with mean diameter 40 nm were selected as the
matrix alloy and reinforcement, respectively. 10 vol% SiC/Al nano-
composites were fabricated by mechanical alloying and hot pressure
sintering followed by hot extrusion. Firstly, under a small amount of
process control agent, Al powders with 10 vol% SiC nanoparticles were
mechanical alloyed with a ball-to-material ratio of 15:1 at 200 rpm for
various time durations (6, 9, 12, 15, 18 and 21 h) in a planetary ball
mill, respectively, from which an optimum milling time was de-
termined. Then the milled powders were vacuum sintered at 580 °C for
1 h under a pressure of 20 MPa. Finally, the sintered billets were ex-
truded at 450 °C with an extrusion ratio of 25:1.

2.2. Microstructure and tensile properties

The microstructure evolution of 10 vol% SiC/Al nanocomposites


during the whole fabrication process and the tensile fracture
morphologies were characterized using a Hitachi S-3000N scanning
electron microscope (SEM). SiCnp distribution, Al grains and SiC-Al
interfacial integrity were observed on a transmission electron micro- Fig. 2. Variation of Al grain size with milling time in 10 vol% SiC-Al nano-
scope (TEM) and high-resolution TEM (HRTEM). The constituent composite powders.
phases of the mechanical alloyed nanocomposite powders were de-
termined by Philips X′pert X-ray diffraction (XRD) using Cu Kα radia- the addition of high content SiCnp. Meanwhile, Al peaks are found to be
tion (λ = 0.154051 nm). Al grain size dXSR during mechanical alloying gradually broadened due to grain refinement and lattice strain accu-
process was estimated using the following Scherrer equation [21]: mulation.
0.9λ Fig. 2 shows the Al grain size determined by the full width at half
dXSR =
B cos θ (1) maximum (FWHM) of the main diffraction peaks in Fig. 1. With in-
creasing milling time, the Al grain size decreases rapidly, until a steady
where B is full width at half maximum (FWHM) of the diffraction peaks
value about 55 nm was reached after 15 h. This suggests that ball mil-
after removing the instrumental broadening, λ is wavelength of in-
ling reaches a steady state after 15 h, which produces steady nano-
cident X-ray radiation and θ is Bragg's diffraction angle. For extruded
crystalline Al in the nanocomposite powders. The grain refinement to
nanocomposites, the grain size distribution of the Al matrix alloy was
nanometer scale by mechanical alloying is attributed to the high dis-
determined from TEM images using Image-Pro Plus 6.0 software. The
location generated by severe plastic deformation [22]. Furthermore, the
relative densities of the sintered and extruded nanocomposites were
addition of nanoparticles accelerates the plastic deformation of the
measured by Archimedes method. Tensile stress-strain curves at room
matrix alloy and formation of nanocrystal grains [23,24] because the
temperature for Al alloy and SiC/Al nanocomposite were tested on an
interaction between nanoparticles and dislocations can induce the for-
Instron-5500 universal testing machine with a rate of 0.5 mm/min.
mation of high density dislocations. Therefore, the grain size decreases
quickly at the beginning until a balance between the grain refinement
3. Results and discussion
induced by severe impact deformation and grain coarsening by thermal
recovery was reached, after which the grain size keeps intact [25].
3.1. Microstructural evolution during mechanical alloying

3.1.1. Nanocrystalline Al in nanocomposite powders 3.1.2. Nanoparticle distribution in the nanocomposite powders
In order to analyze the effect of milling time on the Al grain size It is a challenging task to realize the uniform distribution of high
during mechanical alloying, XRD patterns of 10 vol% SiC-Al nano- content nanoparticles in the nanocomposites. Mechanical alloying is
composite powders were tested, as shown in Fig. 1. It can be seen in effective in achieving a homogeneous distribution of nano-[23–27] and
Fig. 1 that besides Al peaks, obvious SiC peaks are observed because of micro-[14,15] particles in matrix alloys. Fig. 3 displays SEM

11
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

Fig. 3. Representative morphologies of 10 vol% SiC-Al mixed powders after ball-milling for various time durations of (a) 9 h and (b) 15 h. (c) is an enlarged
micrographs of (b).

micrographs of 10 vol% SiC/Al powders milled for 9 h and 15 h, which nanocomposites. In conventional extrusion process, a high enough ex-
represent two typical morphologies before and after steady state, re- trusion ratio is often required to completely break down PPB [30]. The
spectively. Obviously, the nanocomposite powders are flat-shaped after deformation standard of breaking down PPB during hot extrusion is
milling for 9 h (Fig. 3(a)) and equiaxed for 15 h (Fig. 3(b)). The re- given as follows [28]:
peated welding-fracturing processes contribute to the change from
flattened to equiaxed particles, during which homogenous dispersion of ER 1 ⎞
Rs = 0.4 ⎜⎛ ER × AR + + ⎟
nanoparticles was realized [26,27]. Moreover, the addition of high ⎝ AR ER ⎠ (2)
content nanoparticles in Al powders helps in enhancing the Al work
where Rs is the PPB area ratio between the extruded and sintered ma-
hardening rate and nanocomposite powder fracture tendency, thus ac-
terial, ER is extrusion ratio, and AR is the aspect ratio of extruded bar
celerating the SiCnp dispersion process [25]. Ultimately, SiCnp are uni-
cross-sections (AR = 1 for extruded round rod). When Rs > 4, the ex-
formly distributed in the nanocomposite powders after 15 h milling, as
trusion deformation can fully eliminate PPB. In the present work, a
shown in Fig. 3(c).
large extrusion ratio of 25:1, corresponding to Rs = 4.0 using Eq. (2), is
selected to extrude the as-sintered nanocomposites into the round rod.
3.2. Microstructure of sintered nanocomposites Fig. 5 shows microstructure of the extruded 10 vol% SiC/Al nano-
composites with Rs = 4.0. As shown in Fig. 5(a), no PPBs are found in
Mechanical alloying is usually used to fabricate uniformly dis- the extruded nanocomposites, which is consistent with the result pre-
tributed nanoparticle reinforced Al matrix composites. However, the dicted by Eq. (2). A bright field TEM image in Fig. 5(b) further confirms
surfaces of milled aluminum powders are often covered by nanometer- the homogeneous distribution of the SiCnp in the extruded nano-
thick oxide films [28]. Therefore, it is necessary to break and shear off composites. Meanwhile, relative density up to 99.4% was realized after
these oxide films during the bulk material densification process [29]. extrusion. These facts confirm that hot extrusion with Rs = 4.0 may
Usually, hot pressure sintering and hot extrusion are used to consolidate effectively eliminate PPB with oxide films, the particle clusters and
the milled powders into the bulk materials. The microstructure of the micro-pores. So, completely consolidated composites was obtained
as-sintered 10 vol% SiC/Al nanocomposite is characterized in Fig. 4. [31].
Some prior particle boundary (PPB), a few nanoparticle clusters/pores
and debonding along PPB on the fracture surface are found in Fig. 4(a), 3.3.2. Ultrafine grains and intragranular nanoparticles formed during
(b) and (c), respectively, suggesting that some oxide films and weak dynamic recrystallization (DRX)
bonding between the powder particles exist in as-sintered nano- Besides eliminating PPB and densifying nanocomposites, hot ex-
composites. The relative density of 94.2% further proves that the sin- trusion also works in obtaining new microstructures for high strength
tered nanocomposites are incompletely compacted. In addition, the and ductility via DRX. TEM micrographs in Fig. 6 show morphology of
high magnification fractograph in Fig. 4(d) reveals that most SiCnp are the Al grains in the extruded nanocomposite. Equiaxed Al grains formed
homogeneously distributed in Al matrix alloys. During conventional hot by recrystallization are observed in Fig. 6(a), suggesting that dynamic
pressure sintering, the pressure can hardly provide high enough de- recrystallization (DRX) happens during hot extrusion of 10 vol% SiC/Al
formation to completely break up the oxide films along PPB of Al alloys nanocomposites. During DRX of the metallic materials, new grains
[28]. Such residual PPB with the oxide films deteriorates the bonding nucleate and grow under the driving force provided by the storage
strength between the milled Al particles, thereby inhibiting the material dislocation energy. It is noteworthy that the addition of second phase
densification. Therefore, hot pressure sintering is difficult to completely particles in metals may accelerate recrystallization due to the particle
eliminate PPB and to fully densify powders. stimulated nucleation or retard recrystallization by pinning of the grain
boundaries [32]. The transition from accelerated to retarded re-
3.3. Microstructural characterizations of extruded nanocomposites crystallization typically occurs when f/dp > 0.1–0.2 µm−1 [33], where
f is particle volume fraction and dp is particle diameter. In the present
3.3.1. PPB elimination and nanocomposite densification work, the calculated f/dp is 2.5 µm−1, revealing that the addition of
In order to further eliminate PPB and fully densify the materials, it is 10 vol% SiCnp with diameter 40 nm may retard DRX of the nano-
necessary to perform the severe deformation for the sintered composites during hot extrusion. The nanoparticles may pin the grain

12
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

Fig. 4. Optical (a, b) and SEM (c, d) micrographs showing the microstructure of the as-sintered 10 vol% SiC/Al nanocomposite. (a) Prior particle boundaries (PPB)
morphology; (b) Nanoparticle clusters and pores; (c) Debonding along PPB on the fracture surface; (d) Homogeneous distributed nanoparticles in Al matrix alloys.

boundaries (namely Zener pinning) during grain growth. However, in the extruded nanocomposite. Obviously, Al grain size is mainly dis-
during hot extrusion with high extrusion ratio, the high dislocation tributed in the sub-micron range, with average diameter ~320 nm.
density induced by the intense shear deformation may provide a large During DRX, such ultrafine grains are formed through the gradual
driving force to overcome Zener pinning. Therefore, new grains can transformation of the dislocation sub-boundaries produced at low
nucleate and grow in spite of the nanoparticle pinning on the grain strains into ultrafine grains with high angle boundaries produced at
boundaries. Similar recrystallized structures have also been found in large strains [36]. For high volume fractions of particles f ≥ 0.1, the
composites with small secondary phases processed by severe plastic grain diameter D can be estimated based on the effect of second phase
deformation [18,34]. particles on the grain growth using Zener-pinning relation [37]:
It can clearly be seen from Fig. 6(b) that most SiCnp are uniformly
dispersed within the recrystallized Al grains, forming an intragranular dp
D = 0.728
structure. During grain growth of DRX, when the large driving force f 1.02 (3)
provided by severe plastic deformation is larger than Zener pinning
force, the grain boundaries will migrate and pass through the nano- where D is grain diameter, while dp and f are particle diameter and
particles until the driving force is equal to Zener pinning force. Con- volume fraction, respectively. The calculated grain diameter D
sequently, most nanoparticles are left inside the recrystallized Al grains = 306 nm is close to the experimentally measured value of 320 nm,
due to DRX. Likewise, such nanoparticles distributed within the re- further demonstrating that the addition of 10 vol% SiC nanoparticles
crystallized grains have also been observed in SiCnp/AZ91 nano- with diameter 40 nm can indeed result in the formation of ultrafine-
composite processed by forging [6]. Moreover, the effect of particle grained Al alloy due to the nanoparticle pinning effect on the grain
pinning on grain growth has been analyzed by three-dimensional phase- growth during DRX. Meanwhile, the ultrafine grains exhibit an ex-
field simulations [35]. tensive size range, as shown in Fig. 7, which has also been attained
Fig. 7 shows the size distribution histogram of the ultrafine Al grains through DRX in other composite materials [18].

Fig. 5. (a) Optical microscope and (b) Bright field TEM micrographs showing the microstructure of the extruded 10 vol% SiC/Al nanocomposite.

13
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

Fig. 6. Bright field TEM micrographs showing morphologies of the Al grains in the extruded 10 vol% SiC/Al nanocomposites. (a) Recrystallized Al grains, (b)
Intragranular SiCnp.

orientation relationship in some interfaces between Al and SiC, e.g. (11-


1)Al//(11-1)SiC and < 011 > Al// < 011 > SiC. Generally, the grain nu-
cleation and growth processes during DRX obey the law of the lowest
free energy. Moreover, the nanoparticles can act as the nucleation sides
during DRX [39]. Along the {111}Al plane and < 110 > Al direction,
the highest growth rate of Al on the surface of SiC can be achieved to
decrease the free energy [40]. At the same time, there is a small lattice
match between {111}Al and {111}SiC due to their identical face cen-
tered cubic structure and similar inter-planar spacing. Therefore, pre-
ferred orientation relationships, e.g. (11-1)Al//(11-1)SiC and < 011 >
Al// < 011 > SiC, exist in the nanocomposite which helps in the crea-
tion of a strong interfacial bonding strength between Al and SiC. So an
optimal microstructures for 10 vol% SiC/Al nanocomposites with uni-
formly distributed intragranular nanoparticles and ultrafine-grained
matrix as well as the strong bonding interfaces have been realized using
mechanical alloying, hot sintering and hot extrusion processes.

Fig. 7. The grain size distribution histograms of the extruded 10 vol% SiC/Al
nanocomposites. 3.4. Tensile properties of extruded 10 vol% SiC/Al nanocomposite

3.3.3. Interfacial integrity between SiC nanoparticle and Al alloy Fig. 9 illustrates the tensile engineering and true stress-strain curves
Fig. 8 shows the interfacial microstructure of the extruded SiC/Al and corresponding fractograph of the extruded pure Al and 10 vol%
nanocomposite characterized by TEM and HRTEM. The clean particle- SiC/Al nanocomposite. From the engineering stress-strain plots in
matrix interface without any reaction phases is displayed in Fig. 8(a). Fig. 9(a), the yield strength (YS) 280 ± 6 MPa and UTS 395 ± 7 MPa
The absence of reaction between SiC and Al may be attributed to the all was determined in the nanocomposite, which increase by 284% and
solid state during ball milling, sintering and hot extrusion, thus the 259% compared to those of the Al alloy with YS 73 ± 3 MPa and UTS
prevention of the interfacial reaction [38]. 110 ± 4 MPa, respectively. More interestingly, the nanocomposite
HRTEM image in Fig. 8(b) shows the presence of some specific exhibits a high elongation up to 12 ± 1%. Similar with the pure Al
(Fig. 9(b)), the fractograph of the nanocomposite shown in Fig. 9(c)

Fig. 8. (a) TEM and (b) HRTEM images showing the SiC-Al interfacial microstructure in the extruded 10 vol% SiC/Al nanocomposites.

14
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

Fig. 9. (a) Tensile engineering and true stress-strain curves of the Al alloy and 10 vol% SiCnp/Al nanocomposite, (b) Fractograph of Al alloy, (c) Fractograph of
nanocomposite.

contains high-density ductile dimples, which is consistent to the high decreasing from 17% in the latter [41] to 12% in the former composite.
elongation of the nanocomposite. Moreover, such ductile dimples in Meanwhile, the tensile properties of the present composite are also
nanocomposite are much smaller and shallower than those in pure Al, favorably comparable with those of the Al-6Cu-0.4Mn composite re-
also verifying that Al matrix grains are very fine in the nanocomposite. inforced by 10 vol% SiC microparticles [42]. Therefore, the present
In order to better evaluate the mechanical properties of nano- nanocomposite with homogeneously dispersed high content in-
composites with high content intragranular SiC nanoparticles and ul- tragranular SiCnp and ultrafine Al grains possesses excellent combina-
trafine-grained Al matrix, some tensile properties of Al matrix compo- tion of strength and ductility.
sites reinforced by micro- and nano-sized SiC particles are summarized
in Fig. 10. Obviously, the addition of both micro- and nano-particles 3.5. Strengthening mechanisms of the nanocomposites
enhances the tensile strength, but reduces the ductility [2,11,12,41]. It
is worth noting that the introduction of small amount of nanoparticles The present nanocomposite with high content intragranular SiCnp
retains good ductility [2,11,12], but a high content nanoparticles up to exhibits the high YS up to 280 MPa, which is 284% higher compared
10 vol% often results in the poor ductility [11,12]. Different from the with Al alloy. Therefore, it is necessary to identify the contributions of
above-mentioned 10 vol% SiC/Al nanocomposite [11], the nano- the various strengthening effects on the nanocomposite strength.
composite synthesized in the present work exhibits high strength as For the microstructure aspects of MMNCs, the matrix may be na-
well as good ductility. Moreover, UTS of the present nanocomposite nocrystalline, ultrafine or coarse grained metals, and may contain in-
(395 MPa) is much higher than that of the 10 vol% SiC microparticle tragranular or intergranular nanoparticles. Such complex structures
reinforced Al composite (215 MPa) [41], with elongation slightly will play different contributions to MMNCs strength. In general, the
strengthening mechanisms of MMNCs include the grain boundary
strengthening (namely Hall-Petch strengthening ΔσHP), nanoparticle
strengthening (namely Orowan strengthening ΔσOR), thermal mismatch
strengthening (ΔσCTE) and load transfer from matrix to reinforcements
(ΔσLT) [43,44].

3.5.1. Orowan strengthening associated with intragranular nanoparticles


As is known, Orowan strengthening is induced by the interaction of
dislocations with intragranular nanoparticles. Therefore, such me-
chanism is effective only when the second phase particles are located
within the matrix alloy grains. The contribution of Orowan strength-
ening ΔσOR can be determined according to [43,44]:
0.13Gb dp
ΔσOR = ln
dp [(1/2f )1/3 − 1] 2b (4)

where G is the shear modulus of the matrix alloy (26 GPa for Al), b is the
Burger's vector (0.286 nm for Al) [45,46], dp is the particle diameter
and f is the particle volume fraction. As expected for the present na-
nocomposites, the estimated contribution of Orowan strengthening to
Fig. 10. The ultimate tensile strength (UTS) and elongation of the present YS can reach up to 145 MPa. This is because that the addition of high
10 vol% SiC/Al nanocomposite and some other Al matrix nanocomposites re- content nanoparticles up to 10 vol% can significantly reduce the par-
inforced by micro- and nano-sized particles with different volume fractions. ticle inter-spacing and create a strong interaction between dislocations

15
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

and nanoparticles. Therefore, intragranular nanoparticle strengthening composites reinforced with particles less than 1 µm [51]. Orowan
makes a dominant contribution to the present nanocomposite. strengthening mechanism is known as an important strengthening
mechanism in MMNCs with grain size larger than particle size. How-
3.5.2. Grain boundary strengthening associated with grain size ever, in some nanocomposites, particles mainly located at grain
The above-mentioned experimental and theoretical results all prove boundaries, thus the strengthening effect due to Orowan mechanism
that high content SiCnp can create ultrafine Al grains, which was formed was not considered [16].
by pinning of Al grain boundaries by SiCnp during grain growth process In present ultrafine-grained aluminum nanocomposites, the addi-
of DRX. For the ultrafine-grained matrix alloys, the grain boundary tion of 10 vol% SiCnp with diameter 40 nm can pin the grain boundaries
strengthening effect ΔσHP can also be calculated by Hall-Petch equation during the grain growth, forming ultrafine-grained Al matrix with grain
[45–47]: diameter 320 nm. Meanwhile, such ultrafine grains contribute 71 MPa
to YS by Hall-Petch strengthening; thus it is an important strengthening
∆σHP = KD−1/2 (5)
factor. For some ultrafine-grained nanocomposites, the grain boundary
where K is a constant, D is average grain diameter. For ultrafine-grained strengthening still makes the largest contribution to strength due to the
pure aluminum, K = 40 MPa μm1/2 [45,46]. The average grain dia- existence of intergranular nanoparticles [12,16]. During the cooling
meter is 320 nm measured in our study). Using Eq. (5), the strength- process of the nanocomposite fabrication, thermal mismatch between
ening value is estimated to be 71 MPa, suggesting that the grain reinforcement and matrix induces dislocations around the nanoparticle
boundary strengthening plays a relatively important role in the strength interfaces. Meanwhile, high dislocation density exists near the interface
improvement of 10 vol% SiC/Al nanocomposites. between the matrix and reinforcement particles [52]. Therefore, such
high dislocations only strengthen the small punched zone around the
3.5.3. Thermal mismatch strengthening nanoparticle interfaces rather than the entire matrix, consequently
Generally, the mismatch of the coefficient of thermal expansion contributing only 45–56 MPa to YS by thermal mismatch strengthening.
(CTE) between reinforcement and matrix generates the dislocations in Unlike continuously reinforced composite, the load transfer from ma-
the near-interface areas of the reinforcement, and thus strengthens the trix to nanoparticles through interfacial shear stresses is minor [12] due
matrix alloy. It is noteworthy that, for micro-particle reinforced com- to the small aspect ratio and low volume fraction of nanoparticles
posite, only when the particle volume fraction is larger than 20%, the [12,16,44,52]. In a word, Orowan strengthening due to high content
thermal mismatch induced dislocations can strengthen the entire matrix intragranular nanoparticles plays the dominant role in the nano-
alloys [48]. However, for nanoparticle reinforcements, the content is composite strength followed by the grain boundary strengthening due
difficult to reach up to 20 vol%. Consequently the dislocations punched to ultrafine grains.
out from the nanoparticles into the matrix can only strengthen the small
punched zone around the nanoparticles rather than the entire matrix. 3.5.5. Overall strengthening behaviors based on available models
Therefore, the dislocation density ρ in the punched zone due to the From the calculation results of the indirect strengthening (Orowan
thermal mismatch is given by [49]: strengthening, grain refinement, thermal mismatch) and direct
6fΔαΔT k strengthening (load transfer) effects, the overall strengthening beha-
ρ= ⎛ ⎞ ln(f −1/3 )
b2 ⎝ G ⎠ (6) viors can be estimated. Generally, the contributions to the increase in
YS by various strengthening mechanisms could be taken as a simple
where Δα is the CTE difference between matrix αm and reinforcements summation or the root of the sum of squares of the different mechan-
αp, k is matrix shear strength, ΔT is difference between processing and isms [16,43–45,47,53]. In addition, a modified shear lag model con-
room temperatures. For the present SiC/Al composites, αm sidering micromechanics strengthening mechanisms (indirect con-
= 23.6 × 10−6 /K, αp = 4.3 × 10−6 /K [49] and ΔT = 430 K. The tribution of Orowan effect), grain size strengthening and CTE mismatch
corresponding strengthening effect ΔσCTE due to the dislocations gen- strengthening, was often used to calculate YS of the composites
erated by CTE mismatch can be calculated by [48–50]: [16,50–52,54,55]. The above-mentioned theoretical predictions were
ΔσCTE = aGbρ1/2 (7) found to be in good agreement with the experimental data. Akbarpour
et al. has applied these three calculation methods to more accurately
where a is a constant, equal to 1 [49] or 1.25 [48,50]. The contribution estimate the overall strengthening behavior of ultrafine-grained Cu
of thermal mismatch strengthening is predicted to be 45–56 MPa. matrix nanocomposites with 2 vol% SiCnp [16]. Therefore, in our work,
YS of nanocomposites σcy is estimated by three common calculation
3.5.4. Load transfer from matrix to reinforcements methods using the simple linear summation, the root of the sum of
As shown in Fig. 8, the present SiC/Al nanocomposite has a good squares and modified shear lag model. The first model using simple
interfacial bonding between Al matrix and SiC nanoparticles, sug- linear superposition of several strengthening mechanisms can be given
gesting an effective load transfer from matrix to nanoparticles. A widely as follows [16]:
accepted equation of increase in YS due to load transfer ΔσLT can be
given by [16,43,44]: σCy = σ0 + ΔσHP + ΔσOR + ΔσCTE + ΔσLD (9)

ΔσLT = 0.5σm f (8) where σ0 is the intrinsic frictional stress of the material in the absence of
any other strengthening mechanism (20 MPa for pure Al [45]). In Eq.
where σm is YS of unreinforced Al matrix (73 MPa measured in our
(9), σ0+ΔσHP represent the yield strength of the matrix in the composite,
work). The value of ΔσLT is calculated to be 4 MPa, suggesting that
while ΔσOR + ΔσCTE +ΔσLD are the increase of the yield strength due to
strength improvement due to load transfer is small.
the present of nanoparticles. The yield strength σcy of nanocomposites is
For 10 vol% SiC/Al nanocomposite, the contributions of the
calculated to be 285–296 MPa, and it is reasonably close to the ex-
abovementioned various strengthening strategies ΔσOR, ΔσHP, ΔσCTE and
perimental yield stresses of 280 MPa
ΔσLT are 145, 71, 45–56 and 4 MPa, respectively. In the present in-
The second model using square root of the sum of the squares can be
tragranular nanocomposites, SiCnp mainly locate within the Al grains
applied to calculate the total improvement in YS of the composite Δσ ,
through DRX. Such high content intragranular nanoparticles, which
giving the following equation [16,43]:
directly interact with moving dislocations, can give full play to the
Orowan strengthening mechanism. Therefore, nanoparticle strength- Δσ = (Δσ 2HP + Δσ OR
2 2
+ Δσ CTE + Δσ 2LD) (10)
ening provides very high contribution to YS (145 MPa). As reported by
Zhang et al., Orowan strengthening contributes the most to YS of the The total improvement in YS of the composite Δσ can also be

16
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

obtained according to [44]:


2 2
Δσ = ΔσHP +ΔσLD + Δσ OR + Δσ CTE (11)

Thus, YS of the composite σcy is described [16]:


σcy = σm + Δσ (12)

where YS of unreinforced Al matrix σm = 73 MPa. Using Eq. (10) and


Eq. (11), the total improvement in YS of nanocomposite Δσ is predicted
to be 168–171 MPa and 227–230 MPa, and the corresponding YS of
nanocomposites σcy is calculated to be 241–244 MPa and 300–303 MPa.
Compared with the experimental YS of 280 MPa, the prediction value
using the root of the sum of squares of ΔσOR and ΔσCTE is more accurate
than the root of the sum of squares of all strengthening factors, sug-
gesting that the analysis model using the root of the sum of squares of
ΔσOR and ΔσCTE is more suitable to predict the nanocomposite
strengthening behavior. This is because both Orowan and thermal Fig. 11. Strain hardening rate - true strain and true stress - true strain curve and
their intersection point (necking point).
mismatch strengthening are associated with the particle-dislocation
interaction.
The third model of modified shear lag theory considering the load Considère criterion (Eq. (16)) governing the onset of localized de-
transfer mechanism, which is used to predict YS for discontinuously formation for plastic instability [18,56]:
reinforced composites, can be expressed by the following equation

[51,55]: σ≫⎛⎜ ⎞⎟
⎝ dε ⎠ε̇ (16)
σcy = σmy [f (S+2)/2 + (1−f )] (13)
where σ and ε are true stress and true strain in Fig. 9(a), respectively, so
where σcy and σmy are YS of the composite and matrix materials, S is the that dσ is the strain-hardening rate.
aspect ratio of the reinforcement (S = 1 for equiaxed particles). dε
Fig. 11 shows the strain hardening rate - true strain and true stress-
Therefore, YS of particle reinforced composites σcy can be simplified as
true strain curves. Obviously, ultrafine-grained aluminum nano-
follows [16,50]:
composites exhibit higher strain-hardening rate due to homogeneously
σcy = σmy (1 + 0.5f ) (14) dispersed nanoparticles within the ultrafine-grained alloys [56]. It is
well known that high work-hardening rate helps in delaying the loca-
where σmy represents the YS of matrix with the presence of nano- lized deformation (necking) under tensile stress, and thus is favorable
particles. The addition of nanoparticles in the matrix may change the for the uniform plastic deformation. As expected, the necking point
matrix microstructure (i.e. grain size, dislocation density formed by occurs at a true strain of 5.8%, which is in good agreement with 6%
Orowan strengthening and CTE mismatch) and thus indirectly en- uniform elongation measured from the engineering stress–strain curves
hanced the matrix alloy strength. Therefore, considering the indirect in Fig. 9(a). This confirms the high uniform elongation in 10 vol% SiC
contribution of Orowan strengthening ( ΔσOR ), grain size strengthening reinforced ultrafine-grained aluminum nanocomposites, which is im-
( ΔσHP ), and CTE mismatch strengthening ( ΔσCTE ), σmy can be written as portant for various structural applications. For the intragranular na-
[50,51]: nocomposites, introducing high content intragranular nanoparticles can
σmy = σ0 + ΔσOR + ΔσHP + ΔσCTE (15) induce high density dislocation in ultrafine-grained aluminum matrix
due to Orowan strengthening and thermal mismatch, providing a sig-
According to the modified shear lag model combining indirect and nificant strain hardening effect. Therefore, large uniform elongation
direct strengthening effects, YS of nanocomposites σcy is determined to accompanies by high tensile strength. Similar results has also been re-
be 295–307 MPa, which is similar to the yield stress using the root of ported by Takata et al., in which nanoparticles dispersed in the ultra-
the sum of squares of the different mechanisms. fine-grained matrix significantly enhanced strain hardening, resulting
Using the simple linear summation and the root of the sum of in the simultaneous improvement of uniform elongation and tensile
squares of the different mechanisms as well as a modified shear lag strength [56].
model, the yield strength σcy is estimated to be 285–296 MPa, On the other hand, the composite plasticity is very sensitive to na-
300–303 MPa and 295–307 MPa, respectively. These prediction values noparticle content and distribution. During the deformation of the in-
are comparable to the experimental YS 280 MPa, showing that these tergranular nanocomposites, high content intergranular nanoparticles
theoretical methods can be applied to predict the overall strengthening form a continuous network along the grain boundaries and conse-
behavior of ultrafine-grained aluminum composites reinforced with quently induce serious grain boundary embrittlement and catastrophic
high content intragranular nanoparticles, and to determine the con- macro-brittle fracture at a small strain [7]. Compared to intergranular
tributions of each strengthening mechanism. nanocomposites, high content intragranular nanoparticles uniformly
dispersed within the plastic matrix grains do not deteriorate the con-
3.6. Ductility analysis of the nanocomposites tinuity of the matrix alloy, and therefore are favorable for the plastic
deformation of the matrix. So a good combination of ductility and
Besides the high strength, ultrafine-grained aluminum composites strength is achieved in the ultrafine-grained Al matrix composites re-
reinforced with high content intragranular nanoparticles simulta- inforced with high content intragranular nanoparticles.
neously exhibit good ductility up to 12% total elongation to failure, Besides the introduction of second phase particles, the appropriate
which is quite different from the limited uniform elongation in nano- grain size distribution favors the strain hardening and thus the ductility
crystalline and ultrafine-grained metals due to the high YS and poor of ultrafine-grained metals [58]. In the present work, the wide grain
strain hardening [56,57]. Therefore, a correlation between the plastic size distribution in 50–900 nm range (in Fig. 7) and large amount of
instability and necking failure tendency was introduced to explain the intragranular nanoparticles (in Fig. 6) prevent the localized deforma-
obtained ductility of ultrafine-grained aluminum nanocomposites. The tion and premature fracture, and thus induce significant strain hard-
uniform elongation on the curves in Fig. 9(a) was determined by the ening.

17
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

4. Conclusion mechanical analysis and optimal selection of 7075 aluminum alloy based composite
reinforced with alumina nanoparticles, Mater. Chem. Phys. 178 (2016) 119–127.
[11] L. Kollo, C.R. Bradbury, R. Veinthal, C. Jaggi, E. Carreno-Morelli, M. Leparoux,
10 vol% SiC/Al nanocomposites with intragranular nanoparticles Nano-silicon carbide reinforced aluminium produced by high-energy milling and
and ultrafine-grained matrix were successfully prepared by mechanical hot consolidation, Mater. Sci. Eng. A 528 (2011) 6606–6615.
alloying and hot pressure sintering followed by hot extrusion. The main [12] X. Yao, Z. Zhang, Y.F. Zheng, C. Kong, M.Z. Quadir, J.M. Liang, D.L. Zhang, Effects
of SiC nanoparticle content on the microstructure and tensile mechanical properties
conclusions can be drawn as follows: of ultrafine grained AA6063-SiCnp nanocomposites fabricated by powder me-
tallurgy, J. Mater. Sci. Technol. 33 (2017) 1023–1030.
(1) Mechanical alloying process of 10 vol% SiC-Al nanocomposite [13] C.J. Lee, J.C. Huang, P.J. Hsieh, Mg based nano-composites fabricated by friction
stir processing, Scr. Mater. 54 (2006) 1415–1420.
powders reached a steady state after 15 h. Nanocomposite powders [14] O. El-Kady, A. Fathy, Effect of SiC particle size on the physical and mechanical
with nanocrystalline Al grains and high content SiC nanoparticles properties of extruded Al matrix nanocomposites, Mater. Des. 54 (2014) 348–353.
homogeneously embedded in the Al powders were obtained at a [15] M. Keneshloo, M. Paidar, M. Taheri, Role of SiC ceramic particles on the physical
and mechanical properties of Al-4%Cu metal matrix composite fabricated via me-
milling time of 15 h. The composite billet prepared by sintering at
chanical alloying, J. Compos. Mater. 51 (2017) 1285–1298.
580 °C under a pressure of 20 MPa exhibited obvious PPB, a few [16] M.R. Akbarpour, E. Salahi, F.A. Hesari, H.S. Kim, A. Simchi, Effect of nanoparticle
nanoparticle clusters and pores. content on the microstructural and mechanical properties of nano-SiC dispersed
(2) Hot extrusion at 450 °C with an extrusion ratio 25:1 not only bulk ultrafine-grained Cu matrix composites, Mater. Des. 52 (2013) 881–887.
[17] E. Ma, Instabilities and ductility of nanocrystalline and ultrafine-grained metals,
eliminated PPB retained in the as-sintered nanocomposites, but also Scr. Mater. 49 (2003) 663–668.
produced submicron Al matrix grains via dynamic recrystallization [18] C.M. Hu, C.M. Lai, X.H. Du, N.J. Ho, J.C. Huang, Enhanced tensile plasticity in
(DRX). In addition, extruded nanocomposites exhibited minimal ultrafine-grained metallic composite fabricated by friction stir process, Scr. Mater.
59 (2008) 1163–1166.
porosity, good interfacial bonding between SiC and Al, and in- [19] Z.H. Zhang, T. Topping, Y. Li, R. Vogt, Y.Z. Hou, C. Haines, Mechanical behavior of
tragranular particle distribution state formed by DRX process. ultrafine-grained Al composites reinforced with B4C nanoparticles, Scr. Mater. 65
(3) The introduction of intragranular nanoparticles and formation of (2011) 652–655.
[20] M. Rezayat, A. Akbarzadeh, A.O. whadi, Fabrication of high-strength Al/SiCp na-
ultrafine matrix grains in pure aluminum remarkably increased the nocomposite sheets by accumulative roll bonding, Metall. Mater. Trans. A 43
ultimate tensile strength (UTS) from 110 MPa to 395 MPa (increase (2012) 2085–2093.
up to 259%) and yield strength (YS) from 73 MPa to 280 MPa (up to [21] S. Chakravarty, K. Sikdar, S.S. Singh, D. Roy, C.C. Koch, Grain size stabilization and
strengthening of cryomilled nanostructured Cu 12at% Al alloy, J. Alloy. Compd.
284%), while retaining a good fracture elongation up to 12%. 716 (2017) 197–203.
(4) For 10 vol% SiC/Al nanocomposites, Orowan strengthening attrib- [22] C. Suryanarayana, Mechanical alloying and milling, Prog. Mater. Sci. 46 (2001)
uted to high content intragranular SiC nanoparticles was the 1–184.
[23] B. Li, F. Sun, Q.Z. Cai, J.F. Cheng, B.Y. Zhao, Effect of TiN nanoparticles on mi-
dominant mechanism for the enhancement of strength, followed by
crostructure and properties of Al2024-TiN nanocomposite by high energy milling
grain boundary strengthening due to ultrafine Al grains. and spark plasma sintering, J. Alloy. Compd. 726 (2017) 638–650.
[24] M. Cabeza, I. Feijoo, P. Merino, G. Pena, M.C. Pérez, S. Cruz, P. Rey, Effect of high
Acknowledgements energy ball milling on the morphology, microstructure and properties of nano-sized
TiC particle-reinforced 6005A aluminium alloy matrix composite, Powder Technol.
321 (2017) 31–43.
We gratefully acknowledge the financial supports from (a) National [25] M. Khakbiz, F. Akhlaghi, Synthesis and structural characterization of Al-B4C na-
Key Research and Development Plan, No: 2017YFB0703103), (b) Fund nocomposite powders by mechanical alloying, J. Alloy. Compd. 479 (2009)
334–341.
of National Key Laboratory of Materials Behavior and Evaluation [26] A. Wagih, Effect of milling time on morphology and microstructure of Al-Mg/Al2O3
Technology in Space Environment and (c) National Natural Science nanocomposite powder produced by mechanical alloying, Int. J. Adv. Eng. Sci. 4
Foundation of China (Grant Nos. U1537201, 51401068, 51101041, (2014) 1–7.
[27] Z.R. Hesabi, A. Simchi, S.M.S. Reihani, Structural evolution during mechanical
51201047 and 51771064). milling of nanometric and micrometric Al2O3 reinforced Al matrix composites,
Mater. Sci. Eng. A 428 (2006) 159–168.
Data availability statement [28] Y.W. Kim, W.M. Griffith, F.H. Froes, Surface oxides in P/M aluminum alloys, JOM 8
(1985) 27–32.
[29] H. Abdoli, H. Asgharzadeh, E. Salahi, Sintering behavior of Al-AlN nanostructured
The raw data required to reproduce these findings are available composite powder synthesized by high-energy ball milling, J. Alloy. Compd. 473
from the corresponding author upon request. (2009) 116–122.
[30] P.B. Berbon, W.H. Bingel, R.S. Mishra, Friction stir processing: a tool to homogenize
nanocomposite aluminum alloys, Scr. Mater. 44 (2001) 61–66.
References [31] J.B. Fogagnolo, M.H. Robert, J.M. Torralba, Mechanically alloyed AlN particle re-
inforced Al-6061 matrix composites: powder processing, consolidation and me-
[1] S.C. Tjong, Novel nanoparticle-reinforced metal matrix composites with enhanced chanical strength and hardness of the as-extruded materials, Mater. Sci. Eng. 426
mechanical properties, Adv. Eng. Mater. 9 (2007) 639–652. (2006) 85–94.
[2] M.P. Reddy, R.A. Shakoor, G. Parande, V. Manakari, F. Ubaid, A.M.A. Mohamed, [32] K.S. Tun, M. Gupta, Effect of extrusion ratio on microstructure and mechanical
M. Gupta, Enhanced performance of nano-sized SiC reinforced Al metal matrix properties of microwave-sintered magnesium and Mg/Y2O3 nanocomposite, J.
nanocomposites synthesized through microwave sintering and hot extrusion tech- Mater. Sci. 43 (2008) 4503–4511.
niques, Prog. Nat. Sci. Mater. 27 (2017) 606–614. [33] M. Ferry, P.R. Munroe, Recrystallization kinetics and final grain size in a cold rolled
[3] S.M. Choi, H. Awaji, Nanocomposites–a new material design concept, Sci. Technol. particulate reinforced Al-based MMC, Compos. Part A 35 (2004) 1017–1025.
Adv. Mater. 6 (2005) 2–10. [34] R. Kaibyshev, S. Malopheyev, Mechanisms of dynamic recrystallization in alu-
[4] Z. Zhang, D.L. Chen, Contribution of Orowan strengthening effect in particulate minum alloys, Mater. Sci. Forum 794–796 (2014) 784–789.
reinforced metal matrix nanocomposites, Mater. Sci. Eng. A 483–484 (2008) [35] L. Vanherpe, N. Moelans, B. Blanpain, S. Vandewalle, Pinning effect of spheroid
148–152. second-phase particles on grain growth studied by three-dimensional phase-field
[5] M.K. Akbari, H.R. Baharvandi, K. Shirvanimoghaddam, Tensile and fracture beha- simulations, Compos. Mater. Sci. 49 (2010) 340–350.
vior of nano/micro TiB2 particle reinforced casting A356 aluminum alloy compo- [36] T. Sakai, A. Belyakov, R. Kaibyshev, H. Miura, J.J. Jonas, Dynamic and post-dy-
sites, Mater. Des. 66 (2015) 150–161. namic recrystallization under hot, cold and severe plastic deformation conditions,
[6] K.B. Nie, K.K. Deng, X.J. Wang, T. Wang, K. Wu, Influence of SiC nanoparticles Prog. Mater. Sci. 60 (2014) 130–207.
addition on the microstructural evolution and mechanical properties of AZ91 alloy [37] M. Miodownik, E.A. Holm, G.N. Hassold, Highly parallel computer simulations of
during isothermal multidirectional forging, Mater. Charact. 124 (2017) 14–24. particle pinning, Scr. Mater. 42 (2000) 1173–1177.
[7] Y.C. Kang, S.L. Chan, Tensile properties of nanometric Al2O3 particulate reinforced [38] Y. Saberi, S.M. Zebarjad, G.H. Akbari, On the role of nano-size SiC on lattice strain
aluminum matrix composites, Mater. Chem. Phys. 85 (2004) 438–443. and grain size of Al/SiC nanocomposite, J. Alloy. Compd. 484 (2009) 637–640.
[8] A. Mazahery, H. Abdizadeh, H.R. Baharvandi, Development of high-performance [39] S.F. Hassan, M. Gupta, Development of nano-Y2O3 containing magnesium nano-
A356/nano-Al2O3 composites, Mater. Sci. Eng. A 518 (2009) 61–64. composites using solidification processing, J. Alloy. Compd. 429 (2007) 176–183.
[9] C.L. He, J.M. Wang, W.X. Yu, Q.K. Cai, F.M. He, X.D. Sun, Microstructure and [40] Z.R. Liu, D.Z. Wang, C.K. Yao, Interface characterization of a SiC whisker Al com-
tensile behavior of aluminum matrix composites reinforced with SiC nanoparticles, posite, J. Mater. Sci. 31 (1996) 6403–6407.
Rare Metal. Mater. Eng. 126 (2005) 554–557. [41] C. He, J. Wang, W. Yu, Q. Cai, F. He, X. Sun, Microstructure and tensile behavior of
[10] H.R. Ezatpour, M.T. Parizi, S.A. Sajjadi, G.R. Ebrahimi, A. Chaichi, Microstructure, aluminum matrix composites reinforced with SiC nanoparticles, Rare Met. Mater.
Eng. 35 (2006) 156–160.

18
A.B. Li et al. Materials Science & Engineering A 745 (2019) 10–19

[42] A. Slipenyuk, V. Kuprin, Y. Milman, V. Goncharuk, J. Eckert, Properties of P/M matrix composites produced from powder mixtures via friction stir processing,
processed particle reinforced metal matrix composites specified by reinforcement Compos. Sci. Technol. 71 (2011) 693–698.
concentration and matrix-to-reinforcement particle size ratio, Acta Mater. 54 [51] Q. Zhang, B.L. Xiao, W.G. Wang, Z.Y. Ma, Reactive mechanism and mechanical
(2006) 157–166. properties of in situ composites fabricated from an Al-TiO2 system by friction stir
[43] C.S. Goh, J. Wei, L.C. Lee, M. Gupta, Properties and deformation behaviour of Mg- processing, Acta Mater. 60 (2012) 7090–7103.
Y2O3 nanocomposites, Acta Mater. 55 (2007) 5115–5121. [52] Z. Zhang, D.L. Chen, Consideration of Orowan strengthening effect in particulate-
[44] M. Habibnejad-Korayem, R. Mahmudi, W.J. Poole, Enhanced properties of Mg- reinforced metal matrix nanocomposites: a model for predicting their yield
based nano-composites reinforced with Al2O3 nano-particles, Mater. Sci. Eng. A strength, Scr. Mater. 54 (2006) 1321–1326.
519 (2009) 198–203. [53] N. Kumar, R.S. Mishra, Additivity of strengthening mechanisms in ultrafine grained
[45] J. Gubicza, G. Dirras, P. Szommer, B. Bacroix, Microstructure and yield strength of Al–Mg–Sc alloy, Mater. Sci. Eng. A 580 (2013) 175–183.
ultrafine grained aluminum processed by hot isostatic pressing, Mater. Sci. Eng. A [54] Ho.J. Ryu Seung, I. Cha, Soon H. Hong, Generalized shear-lag model for load
458 (2007) 385–390. transfer in SiC/Al metal-matrix composites, J. Mater. Res. 18 (2003) 2851–2858.
[46] Niels Hansen, Hall-Petch relation and boundary strengthening, Scr. Mater. 51 [55] V.C. Nardone, K.M. Prewo, On the strength of discontinuous silicon carbide re-
(2004) 801–806. inforced aluminum composites, Scr. Metall. Mater. 20 (1986) 43–48.
[47] Buyang Cao, Shailendra P. Joshi, K.T. Ramesh, Strengthening mechanisms in [56] Naoki Takata, Yusuke Ohtake, Kazuhisa Kita, Kazuo Kitagawa, Nobuhiro Tsuji,
cryomilled ultrafine-grained aluminum alloy at quasi-static and dynamic rates of Increasing the ductility of ultrafine-grained copper alloy by introducing fine pre-
loading, Scr. Mater. 60 (2009) 619–622. cipitates, Scr. Mater. 60 (2009) 590–593.
[48] Y.S. Suh, S.P. Joshi, K.T. Ramesh, An enhanced continuum model for size-depen- [57] M. Dao, L. Lu, R.J. Asaro, J.T.M.D. Hosson, E. Ma, Toward a quantitative under-
dent strengthening and failure of particle-reinforced composites, Acta Mater. 57 standing of mechanical behavior of nanocrystalline metals, Acta Mater. 55 (2007)
(2009) 5848–5861. 4041–4065.
[49] S. Shibata, M. Taya, T. Mori, T. Mura, Dislocation punching from spherical inclu- [58] C.C. Koch, Optimization of strength and ductility in nanocrystalline and ultrafine
sions in a metal matrix composite, Acta Metall. Mater. 40 (1992) 3141–3148. grained metals, Scr. Mater. 49 (2003) 657–662.
[50] I.S. Lee, C.J. Hsu, C.F. Chen, N.J. Ho, P.W. Kao, Particle-reinforced aluminum

19

View publication stats

You might also like