Friction Stir Processing Route For Metallic Matrix Composite Production

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

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

net/publication/339536587

Friction stir processing route for metallic matrix composite production

Chapter · February 2020


DOI: 10.1016/B978-0-12-819724-0.11910-X

CITATIONS READS
0 340

3 authors:

Isaac Dinaharan N. Murugan


Tsinghua University PSG College of Technology
111 PUBLICATIONS   2,185 CITATIONS    198 PUBLICATIONS   5,039 CITATIONS   

SEE PROFILE SEE PROFILE

Esther Akinlabi
University of Johannesburg
563 PUBLICATIONS   2,081 CITATIONS   

SEE PROFILE

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

Aluminium MMC View project

Experimental study on characteristics of laser metal deposited Al-si-sn-cu/Ti-6Al-4v composite coatings View project

All content following this page was uploaded by Isaac Dinaharan on 27 February 2020.

The user has requested enhancement of the downloaded file.


11910 Friction Stir Processing Route for Metallic Matrix Composite
Production
Isaac Dinaharan, Tsinghua University, Beijing, China
N Murugan, PSG College of Technology, Coimbatore, India
ET Akinlabi, University of Johannesburg, Johannesburg, South Africa

F
© 2021.

OO
Introduction

Metal matrix composites (MMCs) have become one of the important class of engineering materials of the modern era. MMCs have drawn
a huge research focus in recent years. MMCs were initially conceived during the 1960s but the advancement of production technologies
made it possible to manufacture various combination of matrix and reinforcements. Particles of various materials are currently used as
reinforcement for MMCs. The size of particles varies from micron to nano levels. Ceramic and carbonaceous particles remain a popular

PR
choice. MMCs are characterized by many desirable properties and features such as low weight, high specific strength, superior wear
resistance, limited thermal expansion, high temperature stability, improved resistance to creep, etc. Therefore, MMCs are most sought after
in several industries to substitute components produced using conventional monolithic alloys. MMCs help to improve the performance of
industrial components without invoking additional weight into the system (Miracle, 2005; Sidhu et al., 2016; Lloyd, 1994; Kaczmar et al.,
2000).
MMCs are produced by combining a matrix material and a reinforcement. This presents lot of challenges to successfully manufacture
MMCs. The final MMCs product should have a homogenous distribution of the reinforcement phase to enhance the properties and

D
component service life. Liquid metallurgy routes are commonly used to make MMCs because of simplicity, applicability to mass
production, and economic reasons. The chosen matrix material is melted in the beginning of the process and the selected reinforcement
particles are fed to the molten metal. Although the description of the process looks simple, it involves several hurdles. Most of the
TE
ceramic particles do not wet the molten metal causing rejection from the melt. The density gradient causes either upward or downward
movement which results in settling or floating. Further, the higher temperature initiates interfacial reaction between the molten metal
and the reinforcement particle producing undesirable brittle intermetallic particles. The solidification pattern influences the distribution of
particles adversely. Several methods are used to improve the issues with wettability and other setbacks (Taha, 2001; Hashim et al., 1999,
2002; Yigezu et al., 2013; Ravi et al., 2007; Hashim et al., 2001). A solid-state method is highly preferred to overcome those limitations.
EC

Friction Stir Processing

Friction stir processing (FSP) has evolved as a potential solid-state method to produce surface and bulk MMCs effectively. FSP was
conceived from the latest solid-state welding technology namely friction stir welding (FSW) which was invented at The Welding Institute
RR

in 1991. Mishra et al. (2003) first published an article to prepare MMCs using FSP. The method is relatively simple. The working principle
can be visualized in Fig. 1 (Lee et al., 2006). The matrix material is usually taken in the form of plate which may be in extruded or rolled
form. A groove is machined on the top of the plate in order to accommodate the reinforcement particles. The size of the groove is chosen
based on the volume fraction and the depth of processing required. A milling machine or wire electric discharge machine can be used to
cut the groove. The particles are then stuffed into the groove manually or pre pressed powders can be packed into the groove. The opening
of the groove at the top leads to escape of particles during processing. This can be avoided by processing using a pinless tool initially. The
CO

pinless tool plasticizes a thin layer of material and moves from one side to another and closes the groove opening. Subsequent processing
is carried out using a conventional FSW tool. The tool is not consumed and has two parts namely shoulder and pin. The rubbing of tool
shoulder and shearing action of the pin generated frictional heat which causes plasticization of the material. The geometry and the material
of the tool determines the quantity of frictional heat generated and subsequently the degree of plasticization. The rotary and transverse
movement of the tool causes material flow from advancing side to retreating side. The stirring action of the tool mixes the packed particles
into the plasticized material and creates the composite which is consolidated at the back of the tool due to the application of axial force
UN

on the tool. The temperature raise is above the recrystallization temperature of the substrate material and insufficient to melt the matrix
material. The whole processing of making the composite is accomplished in solid state (Ma, 2008; Arora et al., 2012; Sharma et al., 2015;
Bajakke et al., 2019; Ratheea et al., 2018).
There are various strategies adopted to fabricate the composite as depicted in Fig. 2. Groove strategy is most common. The cross
section of the groove may be square, rectangle or V shaped (Liu et al., 2013; Avettand-Fènoël et al., 2014). The volume fraction is
maximum in square groove compared to V groove. If the tool does not provide sufficient vertical flow of material, a V groove may lead
to functionally graded distribution along the depth. A cover sheet is sometimes used to avoid scattering of particles in lieu of pinless tool.
The thickness of the cover plate is crucial to successfully plasticize and mix with the particles in the groove. A thicker cover plate may
lead to debonding at the edges of the groove and cause hook defect. Some investigators used an array of drilled holes instead of grooves
to pack reinforcement particles. There is a gap between holes in the array. Multiple columns were also used in few investigations. This
arrangement may avoid the need for a cover sheet or a pinless tool and reduce the processing time. However, the distribution will not

Encyclopedia of Materials: Composites, Volume ■ https://doi.org/10.1016/B978-0-12-819724-0.11910-X 1


2 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
Fig. 1 The FSP procedure: (a) cutting groove(s) and inserting particles; (b) using a flat tool to undertake the surface repair; (c) applying a tool
EC

with a fixed pin to undertake the FSP; and (d) conducting multiple FSP passes. Reproduced from Lee, C.J., Huang, J.C., Hsieh, P.J., 2006. Mg
based nano-composites fabricated by friction stir processing. Scripta Materialia 54, 1415–1420.

be constant along the FSP track. The volume fraction is high at the location of holes and low between the hole location. In all these methods,
RR

the depth of the groove or the hole determines the depth of the particle rich processing zone. It is possible to increase the depth of processing
by increasing the length of tool pin considerably to groove depth at the cost of deprived distribution. Another strategy is to precoat the plate
with the reinforcement particle and do the processing. The precoating can be done by mixing reinforcement with a solvent and apply using
a brush. This method produces surface composites to a lower thickness of less than 1 mm (Ratna Sunil, 2016).
FSP process is attractive and presents many benefits. It does not consume huge amount of energy compared to liquid metallurgy
methods. The physical and chemical properties of the reinforcement particles have little or no influence on the nature of the process.
CO

The density gradient or the wettability do not play a role in determining the final dispersion of the particles in the composite. Therefore,
the options of reinforcement using FSP is unlimited. Practically, any potential particle can be reinforcement with ease. The processing
temperature does not promote undesirable interfacial reactions between the matrix and the reinforcement. Absence of solidification avoids
unnecessary pushing of particles during consolidation (Zohoor et al., 2012; Kurta et al., 2011; Yuvaraj et al., 2017; Shyam Kumar et al.,
2015; Akramifard et al., 2014).
UN

Microstructural Features

The common microstructural features of MMCs produced using FSP are presented in this section. FSP was successfully applied to produce
composites based on aluminum, magnesium, copper, steel, and titanium as matrix materials. FSP produces a stir zone similar to FSW. The
stir zone houses the composite. Typical appearance of stir zone of various composites is presented in Fig. 3. It is essential to obtain a stir
zone without any macroscopic defects such as pin hole, worm hole, tunnels, and irregular voids. Optimization of process parameters is
necessary to obtain a sound stir zone. Nevertheless, a defect free stir zone does not guarantee proper dispersion of reinforcement particle.
Grains adjacent to stir zone may undergo coarsening due to frictional heat and constitute heat affected zone (HAZ). However, the size of
HAZ is less to that of conventional FSW because the reinforcements absorb heat and act as heat sink.
Friction Stir Processing Route for Metallic Matrix Composite Production 3

F
OO
PR
D
TE
Fig. 2 Strategies to reinforce the particles using FSP: (a) array of holes, (b) V groove and cover plate, and (c) precoated particles. Reproduced
from Liu, Q., Ke, L., Liu, F., Huang, C., Xing, L., 2013. Microstructure and mechanical property of multi-walled carbon nanotubes reinforced
aluminum matrix composites fabricated by friction stir processing Materials and Design 45, 343–348. Avettand-Fènoël, M.N., Simar, A., Shabadi,
R., Taillard, R., De Meester, B., 2014. Characterization of oxide dispersion strengthened copper based materials developed by friction stir
EC

processing. Materials and Design 60, 343–357. Ratna Sunil, B., 2016. Different strategies of secondary phase incorporation into metallic sheets by
friction stir processing in developing surface composites. International Journal of Mechanical and Materials Engineering 11, 12.

Aluminum Matrix Composites


RR

FSP was initially applied to aluminum alloys to make composites before extending to other nonferrous and ferrous materials. Aluminum
and its alloys can be plasticized with ease compared to other materials. An overview of various aluminum composites by FSP technique is
presented in Table 1. Several kind of reinforcement particles including SiC (Ghanbari et al., 2017; Rathee et al., 2017; Wang et al., 2009),
Al2O3 (Mazaheri et al., 2011; Zarghani et al., 2009), TiC (Thangarasu et al., 2015), B4C (Rana and Badheka, 2018; Zhao et al., 2015),
WC (Huang et al., 2018), TiB2 (Palanivel et al., 2016), TiO2 (Khodabakhshi et al., 2014; Joyson Abraham et al., 2019), ZrO2 (Mirjavadi
CO

et al., 2017), SiO2 (Joyson Abraham et al., 2016), fly ash (Dinaharan et al., 2016b), rice husk ash (Dinaharan et al., 2017b), CNT
(Lim et al., 2009; Hosseini et al., 2015), graphene (Maurya et al., 2016; Sharma et al., 2019), graphite (Mostafapour Asl and Khandani,
2013), NiTi (Dixit et al., 2007; Ni et al., 2014), W (Bauri et al., 2015), Ti (Huang and Shen, 2017), Ni (Yadav and Bauri, 2011), Mo
(Selvakumar et al., 2017a), SS (Selvakumar et al., 2017b), Cu (Yadav and Bauri, 2015), and fibers (Arab et al., 2015) were successfully
used to produce the composites. Such a variety of reinforcements cannot be tried using conventional stir casting methods. FSP is flexible
enough to reinforce any kind of particle irrespective of its physical and chemical properties. A reasonable distribution and an improvement
UN

of properties were reported in most of the literature.


Dinaharan (2016) fabricated AA6082 based AMCs by reinforcing several ceramic particles such as SiC, Al2O3, TiC, B4C, and WC
and investigated the microstructural features and the response of FSP process to fabricate variety of composites. Fig. 4 shows SEM
micrographs of AA6082 AMCs which clearly reveal the distribution of ceramic particles in the aluminum matrix. The distribution of
ceramic particles is observed to be fairly homogeneous. There are no clusters or agglomeration of particles seen. Moreover, there is
no segregation of particles along the grain boundaries. Some particles might be located on the grain boundaries due to smaller grain
size. But entrapment of particles within grain boundaries is absent. Hence, the distribution is considered to be roughly intragranular.
The mechanical and tribological properties of AMCs are influenced by the nature of distribution. A homogeneous and intragranular
distribution is essential to attain higher properties. The FSP process has resulted in the desirable dispersion. Stir casting
4 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
Fig. 3 Optical macrostructure of stir zone containing: (a) AA5083/WC; (b) AZ91/Al2O3; (c) Cu/TiB2. Reproduced from Huang, G., Hou, W.,
Shen, Y., 2018. Evaluation of the microstructure and mechanical properties of WC particle reinforced aluminum matrix composites fabricated by
friction stir processing. Materials Characterization 138, 26–37. Faraji, G., Asadi, P., 2011. Characterization of AZ91/alumina nanocomposite
produced by FSP. Materials Science and Engineering A 528, 2431–2440. Dinaharan, I., Saravanakumar, S., Kalaiselvan, K., Gopalakrishnan, S.,
EC

2017c. Microstructure and sliding wear characterization of Cu/TiB2 copper matrix composites fabricated via friction stir processing. Journal of
Asian Ceramic Societies 5, 295–303.

technique frequently produce inhomogeneous and intergranular dispersion owing to solidification related phenomena. The density gradient
RR

results in improper dispersion (Hashim et al., 2002). Since the aluminum matrix does not melt during FSP, density gradient does not cause
free movement of ceramic particles. This leads to proper dispersion. The dispersion of ceramic particles is a function of process parameters
such as tool rotational speed and traverse speed (Sharma et al., 2015). A fine and homogeneous distribution in the SEM micrographs
confirms that the chosen set of process parameters was sufficient to produce the desirable distribution. The compacted ceramic particles are
distributed throughout the stir zone.
FSP induced a change in the size and morphology of ceramic particles. The severe plastic deformation together with the rotating action
CO

of the tool is able to smash the ceramic particles. The strong stirring action of the tool knocks off the sharp corners of the ceramic particle.
Large size variation of ceramic particles (SiC, TiC, B4C, and WC) in Fig. 4 indicates the fragmentation. Fragmentation was observed by
many researchers (Sharma et al., 2015; Ratheea et al., 2018). The rate of fragmentation depends upon the initial size and shape of the
particles. Large size particles and irregular or polygonal shape particles have the tendency to break off during FSP. The retention of shape
and size of smaller Al2O3 particles after FSP which did not undergo much fragmentation confirms this statement. It is observed that the
large ceramic particles are not surrounded by debris generated due to fragmentation. There is no clustering of small debris either. This
UN

suggests that the debris also mixed well with the plasticized aluminum and dispersed homogeneously in the aluminum matrix. The size of
debris is remarkably low in the order of nanometer compared to the size of initially packed ceramic particles. The size variation leads to
functionally graded local areas within the AMC.
Thangarasu et al. (2015) presents some details on the interface in AA6082/TiC composites. The details of the interface between the
TiC particles and the aluminum matrix can be observed at higher magnification provided in SEM micrographs in Fig. 5. It is observed
from the figure that the interface is very clear without the presence of any reaction products or micro pores. A clean interface increases
the load bearing capacity of the MMC. The plasticized aluminum matrix might have wetted or spread over the entire surface of the TiC
particles during mixing, which may avoid the formation of micro pores. The temperature of the process plays a key role to initiate any kind
of reaction between the TiC particle and the matrix. The local temperature developed during FSP is very low compared to liquid metallurgy
routes to initiate interfacial reaction.
Friction Stir Processing Route for Metallic Matrix Composite Production 5

Table 1 Aluminum matrix composites fabricated by friction stir processing

Matrix Reinforcement Size Reference

AA2024-T351 SiC 50 nm (Ghanbari et al., 2017)


AA6063-T6 SiC 10 µm (Rathee et al., 2017)
A356 Al2O3 50–100 µm, 20–40 nm (Mazaheri et al., 2011)
AA6082 TiC 2 µm (Thangarasu et al., 2015)

F
AA7075 B4C 8–12 µm, 15–18 µm (Rana and Badheka, 2018)
AA5083 WC 27 µm (Huang et al., 2018)
AA6082 TiB2 20 µm (Palanivel et al., 2016)

OO
Al–Mg TiO2 30 nm (Khodabakhshi et al., 2014)
AA5083 ZrO2 110 nm (Mirjavadi et al., 2017)
AA6063-T6 SiO2 32 µm (Joyson Abraham et al., 2016)
AA6061 Fly ash 2 µm (Dinaharan et al., 2016b)
AA6061 Rice husk ash 8 µm (Dinaharan et al., 2017b)
Al 6111–T4/Al 7075–T6 CNT 30–50 nm (Lim et al., 2009)
AA6061 Graphene 15 nm (Maurya et al., 2016)

PR
AA5083 Graphite 10–50 mm (Mostafapour Asl and Khandani, 2013)
AA1100 NiTi 2–193 µm (Dixit et al., 2007)
AA5083 W 10 µm (Bauri et al., 2015)
AA5083 Ti 23 µm (Huang and Shen, 2017)
AA1050 Ni 70 µm (Yadav and Bauri, 2011)
AA6082 Mo 25 µm (Selvakumar et al., 2017a)
AA6082 SS 40 µm (Selvakumar et al., 2017b)

D
AA1100 Glass fiber 300 µm (Arab et al., 2015)
TE
Palanivel et al. (2016) recorded optical photo micrographs of AA6082/TiB2 aluminum composites at different regions within the stir
zone as presented in Fig. 6. TiB2 particles are dispersed effectively to all regions of the stir zone. There is no region without the dispersion
of particles. The distribution is nearly constant across the stir zone. The variation in the dispersion of TiB2 particles from the advancing side
to the retreating side or from the top to the bottom side is small. This led to a conclusion that the distribution is independent of the region
EC

in the stir zone. Certain researchers have, however, noticed substantial variation in the dispersion of ceramic particles within the stir zone
(Sharma et al., 2015). These distribution variations are the results of insufficient plasticization of the aluminum matrix and non-optimum
tool rotational speeds which does not facilitate and even dispersion of the TiB2 particles to all regions of the stir zone. It is difficult to obtain
constant dispersion of ceramic particles across the whole composite synthesized using casting techniques. The velocity of the solidification
front will be varying across the mold which induces substantial variation of reinforcement particles across the composite castings.
Selvakumar et al. (2017a) elaborated the grain size evolution in AA6082/Mo composites. EBSD images of the AA6082/Mo MMCs
RR

and the effect of Mo particles on the grain size are presented in Fig. 7. The grain structure present in the composites is clearly revealed
in the EBSD images. The grain structure of the aluminum matrix exhibits a combination of coarse elongated grains and fine grains. The
evolution of elongated grain structure is due to rolling process of the as-received aluminum plates. The average grain size was 31.66 µm.
The grains in the composite show (Fig. 7(b–d)) very fine and equiaxed structure. The grains are refined extensively in the composite.
It is well documented in the literature that dynamic recrystallization during FSP results in the formation of fine-grained structure (Ma,
2008). Metals which undergo hot working processes encounter dynamic recrystallization. Recrystallization takes place due to severe plastic
CO

deformation and frictional heating. Dynamic recrystallization is accomplished in the following ways: (a) continuous extended recovery
processes through extinction and rearrangement of dislocations and (b) discontinuous formation of relatively dislocation free grains through
recrystallization (Emami and Saeid, 2015). The mechanism of operation is depended upon the strain rate history in addition to stacking
fault energy of the material. Since aluminum possesses high stacking fault energy, it undergoes continuous dynamic recrystallization. It is
observed in Fig. 7(b–d) that the grain size decreased with an increase in Mo particle volume fraction. This observation suggests that Mo
particles might acted as grain refiners. Mo particles pinned the movement of grain boundaries which reduced the rate of grain growth. This
UN

pinning effect leads to refinement of grains. It is interesting to note that the average particle size of Mo particles is several times higher
than the average grain size in the composite. Although porous Mo particles broke during FSP, several large size particles were retained in
the composite. A large Mo particle can neither be retained within the grains nor between adjacent grain boundaries. Hence, a large particle
may be covering several grains. The grain size reduction is drastic from 0 vol% to 6 vol% and gradual from 6 vol% to 18 vol%. This result
is contrary to the trend observed in MMCs using liquid metallurgy techniques and can be related to the nature of FSP. A reinforcement
particle acts as a grain nucleating site within the solidifying composite which restricts the grain growth of aluminum. Hence, a uniform
reduction in grain size with increased content of homogeneously distributed reinforcement particles can be expected. Conversely, FSP
technique employs an additional mechanism due to intense plastic deformation which negates the pinning effect.
6 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
EC
RR
CO

Fig. 4 FESEM micrograph AA6082 MMCs reinforced with (a) SiC, (b) Al2O3, (c) TiC, (d) B4C, and (e) WC. Reproduced from Dinaharan, I.,
2016. Influence of ceramic particulate type on microstructure and tensile strength of aluminum matrix composites produced using friction stir
processing. Journal of Asian Ceramic Societies 4, 209–218.
UN

Bauri et al. (2015) recorded microstructural features of AA5083/W MMC using TEM as depicted in Fig. 8. Fine equiaxed grains
are visible in Fig. 8(a). The particle–matrix interface is presented in Fig. 8(b). The interface is sharp and free from intermetallics or any
other reaction products. Fine grains are formed by the dynamic recrystallization as discussed earlier. The high stacking fault energy of Al
leads to rearrangement of dislocations into sub-grain boundaries by dynamic recovery (DRV). TEM observations show that dislocations
are in the process of rearrangement into sub-grain boundaries (Fig. 8(c)). A magnified image of the sub-grain boundary in Fig. 8(d)
shows that the sub-grain boundary is indeed composed of a well-arranged array of dislocations. Incorporation of dislocations, generated
during the deformation process, into the sub-grain boundaries gradually increases the misorientation and they progressively turn into
low-angle grain boundaries. The diffraction contrast observed between many of the sub-grain boundaries (Fig. 8(e)) in TEM also confirms
that misorientation across these boundaries is close to high angle. Dislocation glide assisted lattice rotation may turn these boundaries
into high-angle grain boundaries producing a recrystallized fine-grained structure. It is known that continuous dynamic recrystallization
(CDRX) takes place by gradual transformation of low-angle sub-grain boundaries, formed by dislocation rearrangement, to high-
Friction Stir Processing Route for Metallic Matrix Composite Production 7

F
OO
Fig. 5 SEM micrograph of AA6082/TiC MMCs at higher magnification containing: (a) 18 vol%, and (b) 24 vol%. Reproduced from

PR
Thangarasu, A., Murugan, N., Dinaharan, I., Vijay, S.J., 2015. Synthesis and characterization of titanium carbide particulate reinforced AA6082
aluminum alloy composites via friction stir processing. Archives of Civil and Mechanical Engineering 15, 324–334.

angle boundaries. Therefore, it appears that a continuous type dynamic recrystallization process driven by dynamic recovery (DRV) has
resulted in the fine-grained structure. The basic mechanism of grain structure evolution is not changed due to the nature of reinforcement

D
particle.

Magnesium Matrix Composites


TE
Magnesium is the lightest material which is in demand in automotive industries to improve fuel economy by reducing the weight of the
structures. Therefore, magnesium composites are wanted to replace aluminum alloys and its composites in some applications. FSP is a
suitable method to produce such composites because it is so difficult to fabricate magnesium composites by liquid metallurgy routes (Ratna
Sunil et al., 2016). An overview of various magnesium composites by FSP technique is presented in Table 2. Several kind of reinforcement
particles including SiC (Naser and Darras, 2017; Sun et al., 2012), Al2O3 (Faraji et al., 2011; Faraji and Asadi, 2011), TiC (Balakrishnan
EC

et al., 2015), SiO2 (Lee et al., 2006; Khayyamin et al., 2013), ZrO2 (Navazania and Dehghani, 2016), CNT (Lu et al., 2013), fly ash
(Dinaharan et al., 2019), C60 molecule (Morisada et al., 2006), hydroxyapatite (Ratna Sunil et al., 2014), and carbon fibers (Mertens
et al., 2015) were successfully used to produce magnesium composites. Magnesium has hexagonal closed packed structure and a brittle
material. It is difficult to plasticize unlike aluminum. Therefore, the processing window to obtain successful composite is narrow. Fig. 8
shows typical defects occurring on the crown during the fabrication of AZ31/SiC MMC reported by Asadi et al. (2010) The microstructural
RR

features are similar to the features observed in aluminum composites. Appropriate choice of process parameters would provide a uniform
distribution of reinforcement particles across the stir zone with strong interfacial bonding (Fig. 9).

Copper Matrix Composites


Pure copper is used in many industries because of its high thermal and electrical conductivity, plasticity, softness, and formability.
CO

However, poor wear resistance, low hardness, and strength limit its applications and service life in components subjects to tensile loads
and sliding wear. Pure copper connectors undergo premature wear in electrical breakers. Therefore, various particles are reinforced to
produce copper composite. FSP is a good choice to produce copper composites compared to other liquid processing methods (Ebrahimi
and Par, 2019). An overview of various copper composites by FSP technique is presented in Table 3. Several kind of reinforcement
particles including SiC (Akramifard et al., 2014), Al2O3 (Suvarna Raju and Kumar, 2014), TiO2 (Heidarpour et al., 2019), Y2O3
(Avettand-Fènoël et al., 2014), TiC (Sabbaghian et al., 2014), B4C (Sathiskumar et al., 2013), WC (Khosravi et al., 2014), TiB2
UN

(Dinaharan et al., 2017c), AlN (Saravanakumar et al., 2017), fly ash (Kumar et al., 2018a), rice husk ash (Dinaharan et al.,
2017a), zircon sand (Kumar et al., 2018b), CNT (Jafari et al., 2015), graphite (Sarmadi et al., 2013), BN (Thankachan et al., 2018),
Polymer derived SiCN (Kumar et al., 2015), ZrC (Priyadharshini et al., 2017), and BaTiO3 (Thapliyal and Dwivedi, 2018) were
successfully used to produce copper composites. Pure copper (Akramifard et al., 2014; Suvarna Raju and Kumar, 2014; Heidarpour et
al., 2019; Avettand-Fènoël et al., 2014; Sabbaghian et al., 2014; Sathiskumar et al., 2013; Khosravi et al., 2014; Dinaharan et al., 2017c;
Saravanakumar et al., 2017; Kumar et al., 2018a; Dinaharan et al., 2017a; Kumar et al., 2018b; Jafari et al., 2015; Sarmadi et al., 2013;
Thankachan et al., 2018; Kumar et al., 2015) and its alloys such as Cu-Ni (Priyadharshini et al., 2017) and Cu–Al–Ni–Fe (Thapliyal
and Dwivedi, 2018; Thapliyal and Dwivedi, 2016) were used as matrix material. Since copper has high melting point, the FSP tool should
possess high hot hardness for successful processing. Many microstructural features are similar to aluminum and magnesium composite.
Fig. 10 shows micrographs of Cu/TiB2 MMCs fabricated by Dinaharan et al. (2017c) The micrographs show the distribution of
TiB2 particles at various volume fraction. The nature of distribution and reinforcement are seen in those micrographs. The distribution
of particles covers the total area of the micrographs. Particles are separated by various distances known as interparticle distance.
8 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
EC
RR
CO
UN
Friction Stir Processing Route for Metallic Matrix Composite Production 9

Therefore, the distribution of TiB2 particles in the copper matrix can be considered as nearly homogenous. The mechanical action of the
tool is accountable for the distribution of particles into the plasticized copper. The predominant process parameters of FSP process are
tool rotational speed and traverse speed. Both the parameters exert an influence on the distribution of particles. The interparticle distance
increases with an increase in tool rotational speed and decrease with traverse speed. Barmouz et al. (2011) observed good dispersion
of reinforcement particles at lower traverse speed and aggregation of particles at higher traverse speed. So, a better distribution of TiB2
particles can be ascribed to optimum conditions prevailing under the enforced experimental parameters. There was adequate time for the
plasticized copper layers to go around the tool several times to be reinforced with TiB2 particles before consolidation. There is also no chain

F
of TiB2 particles in the micrographs. Such an arrangement commonly forms in casting methods along the grain boundaries which is known
as segregation (Ebrahimi and Par, 2019). The velocity of the solidification front and the solidification pattern often result in segregation.
There is no solidification in FSP which does not result in particle movement once the plasticized composite is consolidated at the back of

OO
the tool. The particles within the forged composite will not move freely during the process of cooling to the ambient temperature. Absence
of segregation indicates that the distribution of TiB2 particles is not intergranular. Nevertheless, many TiB2 particles may be located on the
grain boundaries. Although, intragranular distribution, i.e., particles enclosed with in a grain is preferred to obtain higher properties, the
difference in the size of TiB2 particles and grain size decides the nature of distribution.
Fig. 11 represents a montage of the TEM micrographs of Cu/18 vol% rice husk ash MMC prepared by Dinaharan et al. (2017a) The
micrographs (Fig. 11(a) and (b)) are covered with ultrafine grains as well as high dislocation density. Inner regions of some grains are plain

PR
which indicates that the dislocation density is lower. The reason can be related to discontinuous dynamic recrystallization. The following
factors contribute to the generation of dislocation density in the composite. Primarily, copper is subjected to severe plastic deformation
throughout the FSP process. A deformed material automatically gives birth to dislocations. The grains in the stir zone experience a cycle
of recrystallization and deformation before forging at the back of the tool. Further, the coefficient of thermal expansion of copper and rice
husk ash particles are different. This variation creates an extra amount of dislocations. Entangled dislocations and pinned grain boundaries
are shown in Fig. 11(c). There are numerous rice husk ash particles in the composite which are below micro level in size. Those particles
improve the Zener-pinning effect which results in the formation of ultra-fine grains. Annealing twins are noticed in the composite (Fig.

D
11(d)). Conventionally etched and optical micrographs do not present a clear view of twins in the stir zone. Dynamically recrystallized
grains of copper go through annealing process during FSP (Xue et al., 2010). A growth fault may be initiated at the interaction points of
several grain boundaries. This growth fault turns to be a twin. The fine grains in the composite confirm that the annealing effect is not
TE
well pronounced to coarsen the grains. The interface between a rice husk ash particle and copper matrix is seen in Fig. 11(e). There is no
reaction layer around the particle.

Steel Matrix Composites


EC

Steel is used widely as a structural material in industry and construction. However, the wear resistance of steel is considered to be poor in
certain applications. Dispersion of hard ceramic particles in the steel matrix can improve strength and wear resistance compared to those
of the monolithic counterparts (Ram Prabhu et al., 2014; Mahathanabodee et al., 2013). Conventionally, power metallurgy route is used
to steel composites which is not economical. FSP is an attractive method to produce steel-based composites. An overview of various steel
composites by FSP technique is presented in Table 4. Several kind of reinforcement particles including Al2O3 (Kahrizsangi et al., 2015),
TiC (Kahrizsangi and Bozorg, 2012), B4C (Joshi et al., 2017), and TiB2 (Newishy et al., 2013) were successfully used to produce steel
RR

composites. Mild steel is the only material used for making the composites. The high melting point and high strength of steel imposes severe
challenge on the tool. The tool undergoes enormous wear within few meters of processing. WC (Kahrizsangi et al., 2015; Kahrizsangi
and Bozorg, 2012; Joshi et al., 2017) and WC–C (Newishy et al., 2013) based tools were used for processing. Therefore, limited works
were carried out on steel composites by FSP. Some of the microstructural features are presented in Figs. 12 and 13. Kahrizsangi et al.
(2015) observed that the preplaced Al2O3 nanoparticles were found to be non-uniformly dispersed in the stir zone (Fig. 12(a)) after the first
FSP pass. A subsequent FSP pass (second pass) resulted in improved dispersion of the clustered Al2O3 particles which is attributed to the
CO

stirring action of FSP. A limited number of regions associated with the clustered Al2O3 particles was discerned (Fig. 2(b)) after the fourth
FSP pass. Higher magnification SEM examination demonstrated the dispersion and cluster size of the Al2O3 particles to be a function of
the number of FSP passes. Multi FSP passes produced uniform dispersion of the nano-sized Al2O3 particles and consequently fabricated
an appropriate composite with respect to the microstructure. The grain structure evolution in the stir zone of TiC reinforced composites
observed Kahrizsangi and Bozorg (2012) by is shown in Fig. 13. The stir zone showed a fine grain microstructure (Fig. 13(a)). Matrix
mean grain size was measured to be ~5 µm after four numbers of passes. Fine grains of the fabricated surface layer in comparison to those
UN

of the unaffected mild steel substrate, reflect the occurrence of a dynamic recrystallization process. A steel/TiC nanocomposite after four
numbers of FSP passes exhibited a mean grain size of ~600 nm (Fig. 13(b)). Nano-sized distribution of TiC reinforcements had an extra
contribution in decreasing the matrix grain size by pinning effect which retarded the grain growth process.

Fig. 6 Optical photomicrograph of AA6082/TiB2 MMCs at various locations within the stir zone: (a) and (b) toward advancing side, (c) and (d)
toward retreading side, (e) middle portion, (f) interface between stir zone and aluminum matrix, (g) and (h) bottom portion. Reproduced from
Palanivel, R., Dinaharan, I., Laubscher, R.F., Paulo Davim, J., 2016. Influence of boron nitride nano particles on microstructure and sliding wear
behavior of AA6082/TiB2 hybrid aluminum matrix composites synthesized by friction stir processing. Materials and Design 106, 195–204.
10 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
EC
RR

Fig. 7 EBSD (IPF + grain boundary) maps of AA6082/Mo AMCs containing Mo particles: (a) 0 vol% (b) 6 vol%, (c) 12 vol%, and (d) 18
vol%. Reproduced from Selvakumar, S., Dinaharan, I., Palanivel, R., Ganesh Babu, B., 2017a. Characterization of molybdenum particles
reinforced Al6082 aluminum matrix composites with improved ductility produced using friction stir processing. Materials Characterization 125,
13–22.
CO

Titanium Matrix Composites


Titanium and its alloys are used in aerospace, petroleum, automobile, defense, and biomedical industries due to several covetable properties
including superior corrosion resistance. However, they suffer high friction coefficient, poor wear resistance loss in mechanical strength
at high temperatures. Reinforcing with various particles to produce titanium composites will improve the sliding wear and friction
UN

characteristics (Tjong and Mai, 2008; Morsi and Patel, 2007). Several methods such as powder metallurgy, spark plasma sintering, vacuum
arc remelting, reactive hot pressing, laser metal deposition, and induction skull melting are used for making titanium composites. Recently,
FSP has been applied to produce titanium composites. An overview of various titanium composites by FSP technique is presented in Table
5. Several kinds of reinforcement particles including SiC (Shamsipur et al., 2011), Al2O3 (Zarghani et al., 2015), TiO2 (Zhang et al.,
2017), TiC (Li et al., 2013), TiB2 (Wang et al., 2019a), TiN (Shamsipur et al., 2013), Ag (Wang et al., 2019b), and hydroxyapatite
(Rahmati and Khodabakhshi, 2018) were successfully used to produce titanium composites. The challenges are similar to steel composite
due to the high melting point of titanium and its alloys. WC tool was predominantly used for processing.
Some of the microstructural features are shown in Figs. 14 and 15. Morsi and Patel (2007) observed that a gradual break-up of
SiC clusters by increasing the number of FSP passes as presented in Fig. 14 in SiC reinforced titanium composites. Material flow is
Friction Stir Processing Route for Metallic Matrix Composite Production 11

F
OO
PR
D
TE
EC
RR

Fig. 8 TEM micrographs of 5083 Al–W composite showing: (a) fine equiaxed grains; (b) particle–matrix interface; (c) dislocation arrangement
into sub-grain boundaries, (d) array of dislocations in the sub-grain boundary and (e) diffraction contrast across the sub-grain
boundaries. Reproduced from Bauri, R., Yadav, D., Shyam Kumar, C.N., Balaji, B., 2015. Tungsten particle reinforced Al5083 composite with
high strength and ductility. Materials Science and Engineering A 620, 67–75.
CO

complex in FSP. Three types of motion are active in the stir zone. These are circumventing motion of surface material around the tool
shoulder, torsional motion due to rotational motion of surface material within the interaction layer under the tool shoulder, and vortex
motion associated with the flow of thickness material due to the action of the tool pin. Such material motions are responsible for the
break-up of SiC clusters and their dispersion in CP-Ti matrix. A uniform dispersion of nano-sized SiC particles was achieved after the
UN

fourth FSP passes (Fig. 14(c)). Fig. 15 shows the microstructures observed by Wang et al. (2019b) in Ti–6Al–4V/Ag composites. Ag
occupies most regions of the stir zone as seen from the morphologies in Fig. 15(b) and (c). The distribution of Ag streamline strips
is attributed to the spinning of FSP tool. Ag agglomeration is elongated along the spin direction and is inevitably formed by the likely
reason of overloaded Ag nanoparticles, while the formation of well-distributed Ag particles is attributed to the inherent merits of dispersing
introduced particles by FSP. The microstructure in SZ is composed of small grains with an average grain size of 10 µm (Fig. 15(c)). Smaller
grains around 5 µm are also observed as streamline strips. The rotation speed plays an important role in the heat input which can accelerate
the recrystallization process. Many other factors, such as the peak temperature, strain rate, cooling rate, and tool design can also affect
the average grain size directly. Small grains of around 5 µm are attributed to the Ag agglomeration, which can accelerate recrystallization
there by preventing grain coarsening, as indicated in Fig. 15(c). The pile-up and tangle of dislocations are easy to form around the nano
particle size of Ag. The dislocation is deposited at the original grain boundary, which can increase the storage energy of deformation and
provide the shape energy for dynamic recrystallization. The introduction of Ag agglomeration can increase the dislocation density and
12 Friction Stir Processing Route for Metallic Matrix Composite Production

Table 2 Magnesium matrix composites fabricated by friction stir processing

Matrix Reinforcement Size Reference

AZ31B SiC 250 µm (Naser and Darras, 2017)


AZ63 SiC 40 nm (Sun et al., 2012)
AZ91 Al2O3 3 µm, 0.3 µm and 30 nm (Faraji et al., 2011)
AZ31 TiC 4 µm (Balakrishnan et al., 2015)

F
AZ91 SiO2 10 nm (Khayyamin et al., 2013)
AZ31 ZrO2 40 nm (Navazania and Dehghani, 2016)
AZ31 CNT 30 nm (Lu et al., 2013)

OO
AZ31 Fly ash 10 µm (Dinaharan et al., 2019)
AZ31 C60 100 nm (Morisada et al., 2006)
Mg Hydroxyapatite 20 nm (Ratna Sunil et al., 2014)
AZ31B Carbon fiber 15–30 µm (Mertens et al., 2015)

PR
D
TE
EC
RR
CO
UN

Fig. 9 Effect of penetration depth (PD) on surface quality and defects in AZ31/SiC MMCs. Reproduced from Asadi, P., Faraji, G., Besharati,
M.K., 2010. Producing of AZ91/SiC composite by friction stir processing (FSP). International Journal of Advanced Manufacturing Technology
51, 247–260.
Friction Stir Processing Route for Metallic Matrix Composite Production 13

Table 3 Copper matrix composites fabricated by friction stir processing

Matrix Reinforcement Size Reference

Cu SiC 25 µm (Akramifard et al., 2014)


Cu Al2O3 20 µm (Suvarna Raju and Kumar, 2014)
Cu TiO2 40 nm (Heidarpour et al., 2019)
Cu Y2O3 3.5 µm (Avettand-Fènoël et al., 2014)

F
Cu TiC 25 µm (Sabbaghian et al., 2014)
Cu B4C 4 µm (Sathiskumar et al., 2013)
Cu WC 5 µm (Khosravi et al., 2014)

OO
Cu TiB2 6 µm (Dinaharan et al., 2017c)
Cu AlN 1 µm (Saravanakumar et al., 2017)
Cu Fly ash 8.2 µm (Kumar et al., 2018a)
Cu Rice husk ash 5 µm (Dinaharan et al., 2017a)
Cu Zircon sand 96 µm (Kumar et al., 2018b)
Cu CNT 20–30 nm (Jafari et al., 2015)
Cu Graphite 5 µm (Sarmadi et al., 2013)

PR
Cu BN 1 µm (Thankachan et al., 2018)
Cu Polymer SiCN 10 µm (Kumar et al., 2015)
Cu–Ni ZrC 5 µm (Priyadharshini et al., 2017)
Cu–Al–Ni–Fe BaTiO3 – (Thapliyal and Dwivedi, 2018)
Cu–Al–Ni–Fe Graphite – (Thapliyal and Dwivedi, 2016)

D
the storage energy, which can accelerate the recrystallization. Shorter excursion time at stir zone effectively restrains grain growth thereby
refining recrystallized grains. As seen from Fig. 15(d), narrowband-like structure is formed in TZ and the mean grain sizes in both stir zone
TE
and transition zone are much smaller than that in base metal. Fig. 15(e) indicates the microstructure of base metal with the average grain
size of 40 µm.

Effect of Process Parameters


EC

Fig. 16 shows a list of factors which can be varied and affect the resultant microstructure and properties (Sathiskumar et al., 2014). It is
essential to understand the role of various process parameters on the distribution of reinforcement particles in the stir zone. It was found
from several literature that the machine factors such as tool rotational speed, traverse speed, and number of passes influence the distribution
significantly.
RR

Tool Rotational Speed


Tool rotational speed refers to the rotating speed of the tool. The rotatory motion rubs the shoulder with the substrate plate and generate
frictional heat. The rotatory motion imparts a momentum to material flow from advancing side to retreating side. The amount of frictional
heat generated and the rate of material movement across the stir zone dictate the distribution of particles and grain evolution (Devaraju et
al., 2013; Moghaddas and Bozorg, 2013).
CO

Fig. 17 reveals the SEM micrographs of AA6082 AMCs prepared by Dinaharan et al. (2016a) as a function of tool rotational speed.
The influence of tool rotational speed on the distribution of TiC particles is visible. The distribution at 800 rpm (Fig. 17(a)) is poor. Particles
are closely packed in many regions. The distribution is fairly homogenous as tool rotational speed is increased to 1200 rpm. The distribution
further improved at 1600 rpm. The increase in tool rotational speed increased the mean interparticle distance. Apart from frictional heat
generation, tool rotation stirs the plasticized material around the pin and results transportation of the plasticized material across the stir
zone. The material flow from advancing side to retreating side at 800 rpm is inadequate causing poor distribution. The tool rotational speed
UN

is not sufficient enough to disperse the packed particles into all regions within the plasticized aluminum. FSP process induces plastic strain
on the processed aluminum. This plastic strain increases as tool rotation speed is increased. The enhanced plastic strain aids to disperse the
particles further into the particle free regions. The agglomeration of particles fades away.
Fig. 18 shows the effects of the rotational speed on grain size and hardness of AZ91/SiC MMC with the additions of various SiC
particles (Asadi et al., 2011). It can be inferred from Fig. 17(a) that addition of SiC particles fades the effect of generated heat on the
grain growth as the rotational speed increases. Also, the rotational speed has no significant effect on the grain size of the specimen with
30 nm SiC particles. Grains break into smaller sizes during deformation, regardless of SiC particles. Therefore, a large number of high
angle grain boundaries are produced. However, in this case, addition of SiC particles leads to inhomogeneous local deformation that assists
the break-up of the grains. According to Fig. 18(a), it seems that this phenomenon is highly affected by the number and size of SiC
14 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
Fig. 10 FESEM micrographs of Cu/TiB2 MMCs containing TiB2: (a) 6 vol%, (b) 12 vol%, and (c) 18 vol%. Reproduced from Dinaharan, I.,
Saravanakumar, S., Kalaiselvan, K., Gopalakrishnan, S., 2017c. Microstructure and sliding wear characterization of Cu/TiB2 copper matrix
EC

composites fabricated via friction stir processing. Journal of Asian Ceramic Societies 5, 295–303.

particles. While the grain size of the specimens with and without 5 µm SiC particles is rather close, there is a huge difference in the cases
of specimens with nano particles. The number of 30 nm SiC particles is much more than 5 µm ones. Consequently, the effect of grain
RR

breaking is more intensive. As shown in Fig. 18(b), the higher the rotational speed, the higher the grain size, and the less the microhardness.
The Hall–Petch relation expresses that the hardness has an inverse correlation with the grain size. But there is an exception. The stir zone
hardness of sample without SiC particles at the rotational speed of 1400 rpm was less than that of the base metal (63 HV), despite the
decrease in the grain size (from ~150 to ~27 µm). It was the result of dissolution of strengthening β (Mg17Al12) phase during the FSP.

Traverse Speed
CO

Traverse speed determines the dwell period of the frictional heat and reduce the material movement. The effect of traverse speed on the
microstructure of Cu/B4C MMC is shown in Fig. 19. The optical micrograph of the composite fabricated at 20 mm/min shows (Fig. 19(a))
homogeneous distribution of B4C particles. The distribution is not uniform (Fig. 19(c)) at 60 mm/min due to the poor distribution of B4C
particles at several places. When traverse speed was increased, the poor distribution of B4C particles gradually increased. The average
spacing between B4C particles decreased when traverse speed was increased. The traverse speed governs the available stirring of rotating
UN

tool per unit length of FSP and affects the transportation of material from advancing side to retreating side. The available stirring is more
at a traverse speed of 20 mm/min which results in higher plastic strain. The uniform distribution of B4C particles as seen in the micrograph
(Fig. 19(a)) can be attributed to intense stirring and sufficient material flow which reduces poor distribution. When traverse speed increases,
the material flow between advancing side and retreading side becomes inadequate. Proper mixing of plasticized copper and B4C particles
does not take place which aids poor distribution.
Microhardness behavior of Cu/SiC MMCs with and without SiC particles is shown in Fig. 20 (Barmouz et al., 2011). It is seen
that microhardness values in the side regions of stir zone are reduced because of annealing-induced grain growth. On the one hand,
stirring action of the pin leads to a dynamic recrystallization in the stir zone which reduces the grain size and enhances the dislocations
which eventually improves the microhardness values. On the other hand, the annealing effect of heat input decreases the microhardness
values. In the specimens processed without SiC particles at traverse speeds of 40 and 80 mm/min, despite the reduction in grain size,
Friction Stir Processing Route for Metallic Matrix Composite Production 15

F
OO
PR
D
TE
EC
RR
CO
UN
16 Friction Stir Processing Route for Metallic Matrix Composite Production

the annealing effect is predominant and microhardness is diminished as compared to the microhardness of the pure copper which was
measured to be 70 HV. It is also to be noted that according to the results, the microhardness values in the stir zone of the aforementioned
specimens are independent of grain size. Thus, other factors such as dislocation density could control the microhardness behavior. However,
in the traverse speed of 200 mm/min, dynamic recrystallization dominates the annealing effect which could be due to lower heat input
and higher thermomechanical stress resulting in enhancement of the microhardness values. Specimens processed with SiC particles, i.e.,
composites show higher microhardness values in comparison with those without SiC particles. This is due to the pinning effect and presence
of hard SiC particles. Fig. 20(b) implies that higher traverse speed causes gathering of SiC particles and thus reduction of the pinning

F
effect of SiC particles which results in lower increment of the microhardness values. Whereas for the specimens prepared at lower traverse
speeds, SiC particles were separated well and consequently an intense pinning effect occurs in stir zone leading to a further enhancement
of microhardness values. It was also found that due to heterogeneous distribution of SiC particles in the specimen produced at a traverse

OO
speed of 200 mm/min, the microhardness values in stir zone are not in the same range. In other words, there is a large difference between
microhardness values in this zone.

Number of Passes
Passes refer to doing the same processing over the previously processed track again. It is discovered in literature that many times it is

PR
difficult to achieve proper distribution in first pass. The distribution and material flow improve with successive passes. Fig. 21 shows the
SEM micrographs of the AA1060/W composites produced by different passes (Huang et al., 2016). It is clearly seen from Fig. 21(a) that
there exist obvious W clusters and only a small amount of aluminum is immersed into W clusters in the 1-pass. This is mainly due to the
following two reasons: (a) during compaction of W particles into the groove, the pressure causes the W particles to mechanically interlock
and cold weld together; (b) relatively large residual strain and stress of as-received rolled aluminum plate from its manufacturing process
result in insufficient material softening that leads to poor plastic flow during 1-pass FSP. As the number of FSP passes increases, the size
of W clusters is significantly decreased. As shown in Fig. 21(b), only a few W clusters are observed after the 3-pass. A more uniform

D
W distribution with almost no W clusters is obtained after the 5-pass. This is attributed to more plastic deformation and thorough mixing
caused by the accumulated plastic strain and the repeated thermal exposure. Nevertheless, no significant change in the size and morphology
of W particles are observed as the number of FSP passes increases. This observation is at odds with the results reported by Prater (2011)
TE
who recorded that the vigorous stirring action of the tool and the intense plastic strain could break the ceramic particles and change their
size and morphology. This is likely to be related to the initial smaller size of W particles. In addition, all SEM micrographs of composites
show no occurrence of particle detachment during the mechanical polishing process. This may imply that an effective interface bonding is
obtained. Further, this suggests the effectiveness of the 5-pass FSP process for the production of composites with uniform microstructure.
Fig. 10 also shows the stress–strain curves for the base AZ91 magnesium alloy and the specimens with nano SiO2 powder as the
EC

reinforcement fabricated in one, two and three passes (Khayyamin et al., 2013). It is expected that the as cast AZ91 magnesium alloy has
the lowest UTS and elongation because of the existence of large grains (140 µm) and hard Mg17Al12 precipitates at the grain boundaries.
The as-cast AZ91 magnesium alloy has a completely brittle fracture with low yield strength and UTS. It is observed in Fig. 10 that the UTS
and elongation of the composite produced in one pass increased from 139 to 193 MPa and 6.74%–12.56%, respectively, in comparison
with the base metal. The main reasons of the improvement in UTS and elongation are as follows: (a) by increasing the FSP pass number,
the distribution of the particles becomes more uniform; (b) hard precipitates of Mg17Al12 in the grain boundaries which are the favorable
RR

sites for starting the cracks are dissolved during processing; (c) porosities and voids of the cast alloys (Fig. 22).

Effect of Tool Design

Tool design refers to number of aspects including shoulder diameter, shoulder profile, pin diameter, pin profile, and tool material
CO

(Sathiskumar et al., 2014; Parumandla and Adepu, 2020; Shojaeefard et al., 2016; Azizieh et al., 2011). All the aforementioned aspects
influence the generation of frictional heat and the subsequent material flow. The particle distribution is greatly affected by the resultant
material flow of the tool design. Sarmadi et al. (2013) used five tools with different pin profiles (straight cylindrical (SC), tapered
cylindrical (TC), threaded cylindrical (TH), square (SQ), and triangular (TR)) as schematically shown in Fig. 23 to fabricate graphite
reinforced copper composites. Fig. 24 presents optical micrographs of particles distribution using different tool pin profiles. These images
were recorded from center of the stir zone. As seen in the micrographs, particles are aggregated in the center of stir zone in case pin
UN

profiles SC, TC, and TH. Particles totally dispersed in the matrix and the distribution of particles is adequate using pin profile SQ and
TR. This is related to flow patterns of copper substrate against different tools during FSP. The particle volume fraction was computed
using and Image Analyzer software. It was found that the volume percent was maximum in composite produced by SC pin being 13.95%.
In the case of composite produced by TR pin, the volume percent was 10.89 vol% which is the minimum value among all of the

Fig. 11 TEM micrographs of Cu/18 vol% rice husk ash MMCs showing: (a) ultra-fine grains, (b) dislocation density, (c) pinning of dislocations,
(d) annealing twins, and (e) interface. Reproduced from Dinaharan, I., Kalaiselvan, K., Akinlabi, E.T., Paulo Davim, J., 2017a. Microstructure and
wear characterization of rice husk ash reinforced copper matrix composites. Journal of Alloys and Compounds 718, 150–160.
Friction Stir Processing Route for Metallic Matrix Composite Production 17

Table 4 Steel matrix composites fabricated by friction stir processing

Matrix Reinforcement Size Reference

Mild Steel Al2O3 70 nm (Kahrizsangi et al., 2015)


Mild Steel TiC 70 nm (Kahrizsangi and Bozorg, 2012)
Mild Steel B4C – (Joshi et al., 2017)
Mild Steel TiB2 2 µm (Newishy et al., 2013)

F
OO
PR
D
TE
Fig. 12 SEM micrographs of mild steel composite produced using (a) one and (b) four FSP passes showing distribution of Al2O3 particles (with
EC

bright contrast) within the stir zone. Reproduced from Kahrizsangi, A.G., Bozorg, S.F.K., Javadi, M.M., 2015. Effect of friction stir processing on
the tribological performance of steel Al2O3 nanocomposites. Surface and Coatings Technology 276, 507–515.
RR

composites. Composite produced by SQ pin had 11.90 vol% graphite particles which is lower than composites produced by SC, TC, and
TH pins. Lower particles volume percent of composites produced by SQ and TR pins showed that particles did not aggregate at the center
of these composites and distribution of particles of these composites is better than three other composites. The size of largest cluster in
composites using SC, TC, and TH profiles is more than 40 µm which proves that particles agglomerated and as can be seen from Fig.
24(a–c). The number of clusters were high in these composites. In the case of composites using SQ and TR profiles, the size of largest
cluster is about 27 and 22 µm, respectively, which are almost the same. Consequently, better distribution of particle was reached using the
CO

tool with triangular pin.


Bahrami et al. (2014) employed five FSW tools with different pin geometries, i.e., threaded tapered, triangular, square, four-flute
square, and four-flute cylindrical (denoted as TT, T, S, FFS, and FFC) to fabricated AA7075/SiC MMCs as presented in Fig. 25. All tools
were machined out of H13 and heat treated to have 58 HRC. SEM micrographs of particle distribution in the stir zone are shown in Fig.
25. It can be inferred from these figures, that SiC particles clustered locally. Although triangular specimen showed finest cluster size, the
most uniform cluster distribution was achieved in threaded tapered specimen. This can be ascribed to the improvement of material flow
UN

associated with downward motion of materials along the probe threads. On the other hand, largest clusters were observed in four-flute
cylindrical specimen. This further supported the inappropriate material flow in four-flute cylindrical specimen. There was a remarkable
difference between the cluster size of four-flute cylindrical specimen and that of the other specimens. However, the remaining cluster sizes
were almost identical in size. Presence of large clusters in four-flute cylindrical specimen is a good reason for lack of pulsating stirring
action. Indeed, pulsating stirring action intensifies the mixing of particles in the matrix. Moreover, downward motion of materials around
the pin is eliminated by absence of threads in FFC pin. Consequently, lack of flat sides or threads are the main reasons behind severe
agglomeration of the reinforcements in four-flute cylindrical specimen (Fig. 26).
18 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
Fig. 13 The grain structure of stir zone (a) without particles and (b) after introduction of TiC particles by four FSP passes. Reproduced from
Kahrizsangi, A.G., Bozorg, S.F.K., 2012. Microstructure and mechanical properties of steel/TiC nano-composite surface layer produced by friction
stir processing. Surface and Coatings Technology 209, 15–22.

Table 5 Steel matrix composites fabricated by friction stir processing

D
TE
Matrix Reinforcement Size Reference

CP Ti SiC 50 nm (Shamsipur et al., 2011)


CP Ti Al2O3 20 and 80 nm (Zarghani et al., 2015)
Ti–6Al–4V TiO2 200 nm (Zhang et al., 2017)
EC

Ti–6Al–4V TiC 5.5 µm (Li et al., 2013)


Ti–6Al–4V TiB2 + TiC 50–200 nm (Wang et al., 2019a)
CP Ti TiN 2 µm (Shamsipur et al., 2013)
Ti–6Al–4V Ag 50 nm (Wang et al., 2019b)
CP Ti Hydroxyapatite 6 µm (Rahmati and Khodabakhshi, 2018)
RR

Summary and Future Outlook

This article presented an overview of the FSP process to fabricate MMCs. The microstructural aspects of various metallic based MMCs
were explored. The role of process parameters and tool design were also presented. FSP is an economical and energy efficient process to
CO

produce MMCs. It is possible to obtain a homogeneous distribution of reinforcement particles in the stir zone by optimizing the process
parameters and tool design. FSP overcomes most of the common defects encountered in liquid metallurgy routes. A sound interface without
any kind of reaction is obtained in FSP. The load bearing capacity of composites by FSP is high. The fine grains generated by dynamic
recrystallization are beneficial to obtain higher properties. FSP can be used to make composites to a desirable depth either at surface level
or bulk thickness by changing the tool design. Multi track grooves can be used to fabricate the composite over the entire substrate material.
There is a demand for MMCs based on high melting point alloys such as titanium, steel, and inconel. The tool design presents
UN

challenges to fabricate such composites due to the higher temperature and forces during processing. The tool design especially tool material
combination should be developed for successful processing over longer processing distance. The ceramic reinforcements used for making
composites interacts with the tool and caused abrasion wear. The pin geometry and profile are lost over few hundred meters of processing.
There is no detailed investigation on tool wear which needs to be explored by researchers. An interest is gradually arising to use the FSP
technique to produce composites based on non-metals. More research work is required to extend the process to develop polymer based
composites.
Friction Stir Processing Route for Metallic Matrix Composite Production 19

F
OO
PR
D
TE
EC
RR
CO
UN
20 Friction Stir Processing Route for Metallic Matrix Composite Production

Fig. 14 SEM micrographs of SiC clusters and their dispersion in the titanium matrix as a function of FSP passes: (a) one, (b) two, and (c)
four. Reproduced from Shamsipur, A., Bozorg, S.F.K., Hanzaki, A.Z., 2011. The effects of friction-stir process parameters on the fabrication of
Ti/SiC nano-composite surface layer. Surface and Coatings Technology 206, 1372–1381.

F
OO
PR
D
TE
EC
RR
CO
UN
Friction Stir Processing Route for Metallic Matrix Composite Production 21

F
OO
PR
D
TE
EC
RR
CO

Fig. 15 SEM images of Ti–6Al–4V/Ag composites of the FSP treated specimen: (a) schematic illustration position where the SEM images were
extracted; (b) and (c) are extracted from stir zone; (d) is extracted from transition zone; (e) is extracted from base metal. Reproduced from Wang,
L., Xie, L., Shen, P., et al., 2019b. Surface microstructure and mechanical properties of Ti-6Al-4V/Ag nanocomposite prepared by FSP. Materials
Characterization 153, 175–183.
UN
22 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
Fig. 16 FSP parameters influencing the properties of the composite. Reproduced from Sathiskumar, R., Murugan, N., Dinaharan, I., Vijay, S.J.,
2014. Prediction of mechanical and wear properties of copper surface composites fabricated using friction stir processing. Materials and Design
55, 224–234.

D
TE
EC
RR
CO
UN
Friction Stir Processing Route for Metallic Matrix Composite Production 23

F
OO
PR
D
TE
Fig. 17 SEM micrographs of AA6082 AMCs at a tool rotational speed of (a) 800 rpm, (b) 1000 rpm, and (c) 1600 rpm. Reproduced from
Dinaharan, I., Murugan, N., Thangarasu, A., 2016a. Development of empirical relationships for prediction of mechanical and wear properties of
AA6082 aluminum matrix composites produced using friction stir processing. Engineering Science and Technology, an International Journal 19,
EC

1132–1144.
RR
CO
UN

Fig. 18 Effect of the rotational speed on: (a) grain size of SZ and (b) microhardness of SZ of Az91/SiC MMC. Reproduced from Asadi, P.,
Besharati Givi, M.K., Abrinia, K., Taherishargh, M., Salekrostam, R., 2011. Effects of SiC particle size and process parameters on the
microstructure and hardness of AZ91/SiC composite layer fabricated by FSP. Journal of Materials Engineering and Performance 20, 1554–1562.
24 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
Fig. 19 Optical photomicrograph of Cu/B4C MMC at traverse speed: (a) 20 mm/min, (b) 40 mm/min, and (c) 60 mm/min. Reproduced from
Sathiskumar, R., Murugan, N., Dinaharan, I., Vijay, S.J., 2013. Effect of traverse speed on microstructure and microhardness of Cu/B4C surface
composite produced by friction stir processing. Transactions of Indian Institute of Metals 66, 333–337.
EC
RR
CO
UN

Fig. 20 Microhardness values of friction stir processed copper: (a) without SiC particles and (b) with SiC particles at different traverse
speeds. Reproduced from Barmouz, M., Givi, M.K.B., Seyfi, J., 2011. On the role of processing parameters in producing Cu/SiC metal matrix
composites via friction stir processing: Investigating microstructure, microhardness, wear and tensile behavior. Materials Characterization 62,
108–117.
Friction Stir Processing Route for Metallic Matrix Composite Production 25

F
OO
PR
D
TE
Fig. 21 SEM micrographs of AA1060/W composites after: (a) 1-pass, (b) 3-passes, and (c) 5-passes. Reproduced from Huang, G., Shen, Y.,
EC

Guo, R., Guan, W., 2016. Fabrication of tungsten particles reinforced aluminum matrix composites using multi-pass friction stir processing:
Evaluation of microstructural, mechanical and electrical behavior. Materials Science and Engineering A 674, 504–513.
RR
CO
UN

Fig. 22 Stress strain curves for base metal and AZ91/SiO2 MMCs processed with traverse speed of 63 mm/min and rotational speed 1250 rpm in
1, 2 and 3 passes. Reproduced from Khayyamin, D., Mostafapour, A., Keshmiri, R., 2013. The effect of process parameters on microstructural
characteristics of AZ91/SiO2 composite fabricated by FSP. Materials Science and Engineering A 559, 217–221.
26 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
Fig. 23 FSP tool pin profile (a) Straight cylinder, (b) Threaded cylindrical, (c) Tapered Cylindrical, (d) Square, and (e) Triangle. Reproduced
from Sarmadi, H., Kokabi, A.H., Reihani, S.M.S., 2013. Friction and wear performance of copper–graphite surface composites fabricated by

PR
friction stir processing. Wear 304, 1–12.

D
TE
EC
RR
CO
UN
Friction Stir Processing Route for Metallic Matrix Composite Production 27

F
OO
PR
D
TE
EC
RR
CO

Fig. 24 Optical microscopic images from center of stir zone containing Cu/graphite composite using tool pin profile: (a) SC, (b) TH, (c) TC, (d)
SQ, and (e) TR. Reproduced from Sarmadi, H., Kokabi, A.H., Reihani, S.M.S., 2013. Friction and wear performance of copper–graphite surface
composites fabricated by friction stir processing. Wear 304, 1–12.
UN
28 Friction Stir Processing Route for Metallic Matrix Composite Production

F
OO
PR
D
TE
EC

Fig. 25 Designed pin geometries: (a) threaded tapered, (b) triangular, (c) square, (d) four-flute square, and (e) four-flute cylindrical. Reproduced
from Bahrami, M., Givi, M.K.B., Dehghani, K., Parvin, N., 2014. On the role of pin geometry in microstructure and mechanical properties of
AA7075/SiC nano-composite fabricated by friction stir welding technique. Materials and Design 53, 519–527.
RR
CO
UN
Friction Stir Processing Route for Metallic Matrix Composite Production 29

F
OO
PR
D
TE
EC
RR
CO

Fig. 26 Dispersion of SiC nano-particles in AA7075/SiC composites of specimen using: (a) TT, (b) T, (c) S, (d) FFS, and (e) FFC pin
tool. Reproduced from Bahrami, M., Givi, M.K.B., Dehghani, K., Parvin, N., 2014. On the role of pin geometry in microstructure and mechanical
properties of AA7075/SiC nano-composite fabricated by friction stir welding technique. Materials and Design 53, 519–527.
UN

References

Akramifard, H.R., Shamanian, M., Sabbaghian, M., Esmailzadeh, M., 2014. Microstructure and mechanical properties of Cu/SiC metal matrix composite
fabricated via friction stir processing. Materials and Design 54, 838–844.
Arab, S.M., Karimi, S., Jahromi, S.A.J., Javadpour, S., Zebarjad, S.M., 2015. Fabrication of novel fiber reinforced aluminum composites by friction stir
processing. Materials Science and Engineering A 632, 50–57.
Arora, H.S., Singh, H., Dhindaw, B.K., 2012. Composite fabrication using friction stir processing – A review. International Journal of Advanced
Manufacturing Technology 61, 1043–1055.
Asadi, P., Besharati Givi, M.K., Abrinia, K., Taherishargh, M., Salekrostam, R., 2011. Effects of SiC particle size and process parameters on the
microstructure and hardness of AZ91/SiC composite layer fabricated by FSP. Journal of Materials Engineering and Performance 20, 1554–1562.
30 Friction Stir Processing Route for Metallic Matrix Composite Production

Asadi, P., Faraji, G., Besharati, M.K., 2010. Producing of AZ91/SiC composite by friction stir processing (FSP). International Journal of Advanced
Manufacturing Technology 51, 247–260.
Avettand-Fènoël, M.N., Simar, A., Shabadi, R., Taillard, R., De Meester, B., 2014. Characterization of oxide dispersion strengthened copper based materials
developed by friction stir processing. Materials and Design 60, 343–357.
Azizieh, M., Kokabi, A.H., Abachi, P., 2011. Effect of rotational speed and probe profile on microstructure and hardness of AZ31/Al2O3 nanocomposites
fabricated by friction stir processing. Materials and Design 32, 2034–2041.
Bahrami, M., Givi, M.K.B., Dehghani, K., Parvin, N., 2014. On the role of pin geometry in microstructure and mechanical properties of AA7075/SiC
nano-composite fabricated by friction stir welding technique. Materials and Design 53, 519–527.
Bajakke, P.A., Malik, V.R., Deshpande, A.S., 2019. Particulate metal matrix composites and their fabrication via friction stir processing – A review. Materials

F
and Manufacturing Processes 34, 833–881.
Balakrishnan, M., Dinaharan, I., Palanivel, R., Sivaprakasam, R., 2015. Synthesize of AZ31/TiC magnesium matrix composites using friction stir processing.
Journal of Magnesium and Alloys 3, 76–78.

OO
Barmouz, M., Givi, M.K.B., Seyfi, J., 2011. On the role of processing parameters in producing Cu/SiC metal matrix composites via friction stir processing:
Investigating microstructure, microhardness, wear and tensile behavior. Materials Characterization 62, 108–117.
Bauri, R., Yadav, D., Shyam Kumar, C.N., Balaji, B., 2015. Tungsten particle reinforced Al5083 composite with high strength and ductility. Materials Science
and Engineering A 620, 67–75.
Devaraju, A., Kumar, A., Kotiveerachari, B., 2013. Influence of rotational speed and reinforcements on wear and mechanical properties of aluminum hybrid
composites via friction stir processing. Materials and Design 45, 576–585.
Dinaharan, I., Kalaiselvan, K., Akinlabi, E.T., Paulo Davim, J., 2017. Microstructure and wear characterization of rice husk ash reinforced copper matrix
composites. Journal of Alloys and Compounds 718, 150–160.

PR
Dinaharan, I., Kalaiselvan, K., Murugan, N., 2017. Influence of rice husk ash particles on microstructure and tensile behavior of AA6061 aluminum matrix
composites produced using friction stir processing. Composites Communications 3, 42–46.
Dinaharan, I., Murugan, N., Thangarasu, A., 2016. Development of empirical relationships for prediction of mechanical and wear properties of AA6082
aluminum matrix composites produced using friction stir processing. Engineering Science and Technology, an International Journal 19, 1132–1144.
Dinaharan, I., Nelson, R., Vijay, S.J., Akinlabi, E.T., 2016. Microstructure and wear characterization of aluminum matrix composites reinforced with
industrial waste fly ash particulates synthesized by friction stir processing. Materials Characterization 118, 149–158.
Dinaharan, I., 2016. Influence of ceramic particulate type on microstructure and tensile strength of aluminum matrix composites produced using friction stir
processing. Journal of Asian Ceramic Societies 4, 209–218.

D
Dinaharan, I., Saravanakumar, S., Kalaiselvan, K., Gopalakrishnan, S., 2017. Microstructure and sliding wear characterization of Cu/TiB2 copper matrix
composites fabricated via friction stir processing. Journal of Asian Ceramic Societies 5, 295–303.
Dinaharan, I., Vettivel, S.C., Balakrishnan, M., Akinlabi, E.T., 2019. Influence of processing route on microstructure and wear resistance of fly ash reinforced
AZ31 magnesium matrix composites. Journal of Magnesium and Alloys 7, 155–165.
TE
Dixit, M., Newkirk, J.W., Mishra, R.S., 2007. Properties of friction stir-processed Al 1100–NiTi composite. Scripta Materialia 56, 541–544.
Ebrahimi, M., Par, M.A., 2019. Twenty-year uninterrupted endeavor of friction stir processing by focusing on copper and its alloys. Journal of Alloys and
Compounds 781, 1074–1090.
Emami, S., Saeid, T., 2015. Effects of welding and rotational speeds on the microstructure and hardness of friction stir welded single-phase brass. Acta
Metallurgica Sinica 28, 766–771.
Faraji, G., Asadi, P., 2011. Characterization of AZ91/alumina nanocomposite produced by FSP. Materials Science and Engineering A 528, 2431–2440.
EC

Faraji, G., Dastani, O., Mousavi, S.A.A.A., 2011. Effect of process parameters on microstructure and micro-hardness of AZ91/Al2O3 surface composite
produced by FSP. Journal of Materials Engineering and Performance 20, 1583–1590.
Ghanbari, D., Asgarani, M.K., Amini, K., Gharavi, F., 2017. Influence of heat treatment on mechanical properties and microstructure of the Al2024/SiC
composite produced by multi–pass friction stir processing. Measurement 104, 151–158.
Hashim, J., Looney, L., Hashmi, M.S.J., 1999. Metal matrix composites: Production by the stir casting method. Journal of Materials Processing Technology
92–93, 1–7.
Hashim, J., Looney, L., Hashmi, M.S.J., 2002. Particle distribution in cast metal matrix composites – Part I. Journal of Materials Processing Technology 123,
RR

251–257.
Hashim, J., Looney, L., Hashmi, M.S.J., 2001. The enhancement of wettability of SiC particles in cast aluminium matrix composites. Journal of Materials
Processing Technology 119, 329–335.
Heidarpour, A., Mazaheri, Y., Roknian, M., Ghasemi, S., 2019. Development of Cu-TiO2 surface nanocomposite by friction stir processing: Effect of pass
number on microstructure, mechanical properties, tribological and corrosion behavior. Journal of Alloys and Compounds 783, 886–897.
Hosseini, S.A., Ranjbar, K., Dehmolaei, R., Amirani, A.R., 2015. Fabrication of Al5083 surface composites reinforced by CNTs and cerium oxide nano
particles via friction stir processing. Journal of Alloys and Compounds 622, 725–733.
CO

Huang, G., Hou, W., Shen, Y., 2018. Evaluation of the microstructure and mechanical properties of WC particle reinforced aluminum matrix composites
fabricated by friction stir processing. Materials Characterization 138, 26–37.
Huang, G., Shen, Y., 2017. The effects of processing environments on the microstructure and mechanical properties of the Ti/5083Al composites produced
by friction stir processing. Journal of Manufacturing Processes 30, 361–373.
Huang, G., Shen, Y., Guo, R., Guan, W., 2016. Fabrication of tungsten particles reinforced aluminum matrix composites using multi-pass friction stir
processing: Evaluation of microstructural, mechanical and electrical behavior. Materials Science and Engineering A 674, 504–513.
Jafari, J., Givi, M.K.B., Barmouz, M., 2015. Mechanical and microstructural characterization of Cu/CNT nanocomposite layers fabricated via friction stir
processing. International Journal of Advanced Manufacturing Technology 78, 199–209.
UN

Joshi, J., Pandey, A.K., Pandey, S., Aravindan, S., 2017. Development of mild steel – B4C surface composite by friction stir processing. Indian Journal of
Science and Technology 10, 1–6.
Joyson Abraham, S., Chandra Rao Madane, S., Dinaharan, I., John, Baruch, L., 2016. Development of quartz particulate reinforced AA6063 aluminum matrix
composites via friction stir processing. Journal of Asian Ceramic Societies 4, 381–389.
Joyson Abraham, S., Dinaharan, I., David Raja Selvam, J., Akinlabi, E.T., 2019. Microstructural characterization and tensile behavior of rutile (TiO2)
reinforced AA6063 aluminum matrix composites prepared by friction stir processing. Acta Metallurgica Sinica 32, 52–62.
Kaczmar, J.W., Pietrzak, K., WøosinÂski, W., 2000. The production and application of metal matrix composite materials. Journal of Materials Processing
Technology 106, 58–67.
Kahrizsangi, A.G., Bozorg, S.F.K., Javadi, M.M., 2015. Effect of friction stir processing on the tribological performance of steel Al2O3 nanocomposites.
Surface and Coatings Technology 276, 507–515.
Kahrizsangi, A.G., Bozorg, S.F.K., 2012. Microstructure and mechanical properties of steel/TiC nano-composite surface layer produced by friction stir
processing. Surface and Coatings Technology 209, 15–22.
Friction Stir Processing Route for Metallic Matrix Composite Production 31

Khayyamin, D., Mostafapour, A., Keshmiri, R., 2013. The effect of process parameters on microstructural characteristics of AZ91/SiO2 composite fabricated
by FSP. Materials Science and Engineering A 559, 217–221.
Khodabakhshi, F., Simchi, A., Kokabi, A.H., et al., 2014. Microstructure and texture development during friction stir processing of Al–Mg alloy sheets with
TiO2 nanoparticles. Materials Science and Engineering A 605, 108–118.
Khosravi, J., Givi, M.K.B., Barmouz, M., Rahi, A., 2014. Microstructural, mechanical, and thermophysical characterization of Cu/WC composite layers
fabricated via friction stir processing. International Journal of Advanced Manufacturing Technology 74, 1087–1096.
Kumar, H., Prasad, R., Srivastava, A., Vashista, M., Khan, M.Z., 2018. Utilisation of industrial waste (Fly ash) in synthesis of copper based surface composite
through friction stir processing route for wear applications. Journal of Cleaner Production 196, 460–468.
Kumar, P.A., Raj, R., Kailas, S.V., 2015. A novel in-situ polymer derived nano ceramic MMC by friction stir processing. Materials and Design 85, 626–634.

F
Kumar, H., Vashista, M., Yusufzai, M.Z.K., 2018. Microstructure and wear Behavior of zircon reinforced copper based surface composite synthesized by
friction stir processing route. Transactions of the Indian Institute of Metals 71, 2025–2033.
Kurta, A., Uygur, I., Cetec, E., 2011. Surface modification of aluminium by friction stir processing. Journal of Materials Processing Technology 211, 313–317.

OO
Lee, C.J., Huang, J.C., Hsieh, P.J., 2006. Mg based nano-composites fabricated by friction stir processing. Scripta Materialia 54, 1415–1420.
Li, B., Shen, Y., Luo, L., Hu, W., 2013. Fabrication of TiCp/Ti–6Al–4V surface composite via friction stir processing (FSP): Process optimization, particle
dispersion-refinement behavior and hardening mechanism. Materials Science Engineering A 574, 75–85.
Lim, D.K., Shibayanagi, T., Gerlich, A.P., 2009. Synthesis of multi-walled CNT reinforced aluminium alloy composite via friction stir processing. Materials
Science and Engineering A 507, 194–199.
Liu, Q., Ke, L., Liu, F., Huang, C., Xing, L., 2013. Microstructure and mechanical property of multi-walled carbon nanotubes reinforced aluminum matrix
composites fabricated by friction stir processing. Materials and Design 45, 343–348.
Lloyd, D.J., 1994. Particle reinforced aluminium and magnesium matrix composites. International Materials Reviews 39, 1–23.

PR
Lu, D., Jiang, Y., Zhou, R., 2013. Wear performance of nano-Al2O3 particles and CNTs reinforced magnesium matrix composites by friction stir processing.
Wear 305, 286–290.
Ma, Z.Y., 2008. Friction Stir Processing Technology: A review. Metallurgical and Materials Transactions A 39, 642–658.
Mahathanabodee, S., Palathai, T., Raadnui, S., Tongsri, R., Sombatsompop, N., 2013. Effects of hexagonal boron nitride and sintering temperature on
mechanical and tribological properties of SS316L h-BN composites. Materials and Design 46, 588–597.
Maurya, R., Kumar, B., Ariharan, S., Ramkumar, J., Balani, K., 2016. Effect of carbonaceous reinforcements on the mechanical and tribological properties of
friction stir processed Al6061 alloy. Materials and Design 98, 155–166.
Mazaheri, Y., Karimzadeh, F., Enayati, M.H., 2011. A novel technique for development of A356/Al2O3 surface nanocomposite by friction stir processing.

D
Journal of Materials Processing Technology 211, 1614–1619.
Mertens, A., Simar, A., Adrien, J., et al., 2015. Influence of fibre distribution and grain size on the mechanical behaviour of friction stir processed Mg–C
composites. Materials Characterization 107, 125–133.
Miracle, D.B., 2005. Metal matrix composites – From science to technological significance. Composites Science and Technology 65, 2526–2540.
TE
Mirjavadi, S.S., Alipour, M., Hamouda, A.M.S., et al., 2017. Effect of multi-pass friction stir processing on the microstructure, mechanical and wear properties
of AA5083/ZrO2 nanocomposites. Journal of Alloys and Compounds 726, 1262–1273.
Mishra, R.S., Ma, Z.Y., Charit, I., 2003. Friction stir processing: A novel technique for fabrication of surface composite. Materials Science and Engineering
A 341, 307–310.
Moghaddas, M.A., Bozorg, S.F.K., 2013. Effects of thermal conditions on microstructure in nanocomposite of Al/Si3N4 produced by friction stir processing.
Materials Science Engineering A 559, 187–193.
EC

Morisada, Y., Fujii, H., Nagaoka, T., Fukusumi, M., 2006. Nanocrystallized magnesium alloy – uniform dispersion of C60 molecules. Scripta Materialia 55,
1067–1070.
Morsi, K., Patel, V.V., 2007. Processing and properties of titanium–titanium boride (TiBw) matrix composites – A review. Journal of Materials Science 42,
2037–2047.
Mostafapour Asl, A., Khandani, S.T., 2013. Role of hybrid ratio in microstructural, mechanical and sliding wear properties of the Al5083/Graphitep/Al2O3p a
surface hybrid nanocomposite fabricated via friction stir processing method. Materials Science and Engineering A 559, 549–557.
Naser, A.Z., Darras, B.M., 2017. Experimental investigation of Mg SiC composite fabrication via friction stir processing. International Journal of Advanced
RR

Manufacturing Technology 91, 781–790.


Navazania, M., Dehghani, K., 2016. Fabrication of Mg-ZrO2 surface layer composites by friction stir processing. Journal of Materials Processing Technology
229, 439–449.
Newishy, M., Elkousy, M., Rahem, N.A., Morsy, M., Khodir, S., 2013. Study on production of steel/TiB2 composite surface layer by friction stir processing.
Academic Journal of Manufacturing Engineering 11, 80–85.
Ni, D.R., Wang, J.J., Zhou, Z.N., Ma, Z.Y., 2014. Fabrication and mechanical properties of bulk NiTip Al composites prepared by friction stir processing.
Journal of Alloys and Compounds 586, 368–374.
CO

Palanivel, R., Dinaharan, I., Laubscher, R.F., Paulo Davim, J., 2016. Influence of boron nitride nano particles on microstructure and sliding wear behavior of
AA6082/TiB2 hybrid aluminum matrix composites synthesized by friction stir processing. Materials and Design 106, 195–204.
Parumandla, N., Adepu, K., 2020. Effect of tool shoulder geometry on fabrication of Al/Al2O3 surface nano composite by friction stir processing. Particulate
Science and Technology 38 (1), 121–130. https://doi.org/10.1080/02726351.2018.1490361.
Prater, T., 2011. Solid state joining of metal matrix composites: A survey of challenges and potential solutions. Materials and Manufacturing Processes 26,
636–648.
Priyadharshini, G.S., Subramanian, R., Murugan, N., Sathiskumar, R., 2017. Surface modification and characterization of zirconium carbide particulate
reinforced C70600 CuNi composite fabricated via friction stir processing. Journal of Mechanical Science and Technology 31, 3755–3760.
UN

Rahmati, R., Khodabakhshi, F., 2018. Microstructural evolution and mechanical properties of a friction-stir processed Ti-hydroxyapatite (HA) nanocomposite.
Journal of the Mechanical Behavior of Biomedical Materials 88, 127–139.
Ram Prabhu, T., Varma, V.K., Vedantam, S., 2014. Effect of reinforcement type, size, and volume fraction on the tribological behavior of Fe matrix
composites at high sliding speed conditions. Wear 309, 247–255.
Rana, H., Badheka, V., 2018. Influence of friction stir processing conditions on the manufacturing of Al-Mg-Zn-Cu alloy/boron carbide surface composite.
Journal of Materials Processing Technology 255, 795–807.
Ratheea, S., Maheshwaria, S., Siddiquee, A.N., Srivastava, M., 2018. A Review of recent progress in solid state fabrication of composites and functionally
graded systems via friction stir processing. Critical Reviews in Solid State and Materials Sciences 43, 334–366.
Rathee, S., Maheshwari, S., Siddiquee, A.N., Srivastava, M., 2017. Investigating effects of groove dimensions on microstructure and mechanical properties
of AA6063/SiC surface composites produced by friction stir processing. Transactions of the Indian Institute of Metals 70, 809–816.
Ratna Sunil, B., 2016. Different strategies of secondary phase incorporation into metallic sheets by friction stir processing in developing surface composites.
International Journal of Mechanical and Materials Engineering 11, 12.
32 Friction Stir Processing Route for Metallic Matrix Composite Production

Ratna Sunil, B., Pradeep Kumar Reddy, G., Patle, H., Dumpala, R., 2016. Magnesium based surface metal matrix composites by friction stir processing.
Journal of Magnesium and Alloys 4, 52–61.
Ratna Sunil, B., Sampath Kumar, T.S., Chakkingal, U., Nandakumar, V., Doble, M., 2014. Friction stir processing of magnesium–nanohydroxyapatite
composites with controlled in vitro degradation behavior. Materials Science and Engineering C 39, 315–324.
Ravi, K.R., Sreekumar, V.M., Pillai, R.M., et al., 2007. Optimization of mixing parameters through a water model for metal matrix composites synthesis.
Materials and Design 28, 871–881.
Sabbaghian, M., Shamanian, M., Akramifard, H.R., Esmailzadeh, M., 2014. Effect of friction stir processing on the microstructure and mechanical properties
of Cu–TiC composite. Ceramics International 40, 12969–12976.
Saravanakumar, S., Gopalakrishnan, S., Dinaharan, I., Kalaiselvan, K., 2017. Assessment of microstructure and wear behavior of aluminum nitrate reinforced

F
surface composite. Surface and Coatings Technology 322, 51–58.
Sarmadi, H., Kokabi, A.H., Reihani, S.M.S., 2013. Friction and wear performance of copper–graphite surface composites fabricated by friction stir processing.
Wear 304, 1–12.

OO
Sathiskumar, R., Murugan, N., Dinaharan, I., Vijay, S.J., 2013. Characterization of boron carbide particulate reinforced in situ copper surface composites
synthesized using friction stir processing. Materials Characterization 84, 16–27.
Sathiskumar, R., Murugan, N., Dinaharan, I., Vijay, S.J., 2014. Prediction of mechanical and wear properties of copper surface composites fabricated using
friction stir processing. Materials and Design 55, 224–234.
Selvakumar, S., Dinaharan, I., Palanivel, R., Ganesh Babu, B., 2017. Characterization of molybdenum particles reinforced Al6082 aluminum matrix
composites with improved ductility produced using friction stir processing. Materials Characterization 125, 13–22.
Selvakumar, S., Dinaharan, I., Palanivel, R., Ganesh Babu, B., 2017. Development of stainless steel particulate reinforced AA6082 aluminum matrix
composites with enhanced ductility using friction stir processing. Materials Science and Engineering A 685, 317–326.

PR
Shamsipur, A., Bozorg, S.F.K., Hanzaki, A.Z., 2011. The effects of friction-stir process parameters on the fabrication of Ti/SiC nano-composite surface layer.
Surface and Coatings Technology 206, 1372–1381.
Shamsipur, A., Bozorg, S.F.K., Hanzaki, A.Z., 2013. Production of in-situ hard Ti/TiN composite surface layers on CP-Ti using reactive friction stir
processing under nitrogen environment. Surface and Coatings Technology 218, 62–70.
Sharma, V., Prakash, U., Manoj Kumar, B.V., 2015. Surface composites by friction stir processing: A review. Journal of Materials Processing Technology
224, 117–134.
Sharma, A., Sharma, V.M., Sahoo, B., Pal, S.K., Paul, J., 2019. Effect of multiple micro channel reinforcement filling strategy on Al6061-graphene
nanocomposite fabricated through friction stir processing. Journal of Manufacturing Processes 37, 53–70.

D
Shojaeefard, M.H., Akbari, M., Khalkhali, A., Asadi, P., 2016. Effect of tool pin profile on distribution of reinforcement particles during friction stir
processing of B4C/aluminum composites. Proceedings of the Institution of Mechanical Engineers Part L: Journal of Materials: Design and Applications
232, 637–651.
Shyam Kumar, C.N., Yadav, D., Bauri, R., Janaki Ram, G.D., 2015. Effects of ball milling and particle size on microstructure and properties 5083Al–Ni
TE
composites fabricated by friction stir processing. Materials Science and Engineering A 645, 205–212.
Sidhu, S.S., Kumar, S., Batish, A., 2016. Metal matrix composites for thermal management: A review. Critical Reviews in Solid State and Materials Sciences
41, 132–157.
Sun, K., Shi, Q.Y., Sun, Y.J., Chen, G.Q., 2012. Microstructure and mechanical property of nano-SiCp reinforced high strength Mg bulk composites produced
by friction stir processing. Materials Science and Engineering A 547, 32–37.
Suvarna Raju, L., Kumar, A., 2014. Influence of Al2O3 particles on the microstructure and mechanical properties of copper surface composites fabricated by
EC

friction stir processing. Defence Technology 10, 375–383.


Taha, M.A., 2001. Practicalization of cast metal matrix composites (MMCCs). Materials and Design 22, 431–441.
Thangarasu, A., Murugan, N., Dinaharan, I., Vijay, S.J., 2015. Synthesis and characterization of titanium carbide particulate reinforced AA6082 aluminum
alloy composites via friction stir processing. Archives of Civil and Mechanical Engineering 15, 324–334.
Thankachan, T., Prakash, K.S., Kavimani, V., 2018. Investigations on the effect of friction stir processing on Cu BN surface composites. Materials and
Manufacturing Processes 33, 299–307.
Thapliyal, S., Dwivedi, D.K., 2016. Microstructure evolution and tribological behavior of the solid lubricant based surface composite of cast nickel aluminum
RR

bronze developed by friction stir processing. Journal of Materials Processing Technology 238, 30–38.
Thapliyal, S., Dwivedi, D.K., 2018. Barium titanate reinforced nickel aluminium bronze surface composite by friction stir processing. Materials Science and
Technology 34, 366–377.
Tjong, S.C., Mai, Y.W., 2008. Processing-structure-property aspects of particulate- and whisker-reinforced titanium matrix composites. Composites Science
and Technology 68, 583–601.
Wang, T., Gwalani, B., Shukla, S., Frank, M., Mishra, R.S., 2019. Development of in situ composites via reactive friction stir processing of Ti–B4C system.
Composites Part B 172, 54–60.
CO

Wang, W., Shi, Q., Liu, P., Li, H., Li, T., 2009. A novel way to produce bulk SiCp reinforced aluminum metal matrix composites by friction stir processing.
Journal of Materials Processing Technology 209, 2099–2103.
Wang, L., Xie, L., Shen, P., et al., 2019. Surface microstructure and mechanical properties of Ti-6Al-4V/Ag nanocomposite prepared by FSP. Materials
Characterization 153, 175–183.
Xue, P., Xie, G.M., Xiao, B.L., Ma, Z.Y., Geng, L., 2010. Effect of heat input conditions on microstructure and mechanical properties of friction-stir-welded
pure copper. Metallurgical and Materials Transactions A 41, 2010–2021.
Yadav, D., Bauri, R., 2011. Processing, microstructure and mechanical properties of nickel particles embedded aluminium matrix composite. Materials
Science and Engineering A 528, 1326–1333.
UN

Yadav, D., Bauri, R., 2015. Development of Cu particles and Cu core-shell particles reinforced Al composite. Materials Science and Technology 31, 494–500.
Yigezu, B.S., Jha, P.K., Mahapatra, M.M., 2013. The key attributes of synthesizing ceramic particulate reinforced Al-based matrix composites through stir
casting process: A review. Materials and Manufacturing Processes 28, 969–979.
Yuvaraj, N., Aravindan, S., Vipin, 2017. Comparison studies on mechanical and wear behavior of fabricated aluminum surface nano composites by fusion
and solid-state processing. Surface and Coatings Technology 309, 309–319.
Zarghani, A.S., Bozorg, S.F.K., Gerlich, A.P., 2015. Strengthening analyses and mechanical assessment of Ti/Al2O3 nano-composites produced by friction
stir processing. Materials Science and Engineering A 631, 75–85.
Zarghani, A.S., Kashani Bozorg, S.F., Hanzaki, A.Z., 2009. Microstructures and mechanical properties of Al/Al2O3 surface nano-composite layer produced
by friction stir processing. Materials Science and Engineering A 500, 84–91.
Zhang, C., Ding, Z., Xie, L., et al., 2017. Electrochemical and in vitro behavior of the nanosized composites of Ti-6Al-4V and TiO2 fabricated by friction stir
process. Applied Surface Science 423, 331–339.
Friction Stir Processing Route for Metallic Matrix Composite Production 33

Zhao, Y., Huang, X., Li, Q., Huang, J., Yan, K., 2015. Effect of friction stir processing with B4C particles on the microstructure and mechanical properties of
6061 aluminum alloy. International Journal of Advanced Manufacturing Technology 78, 1437–1443.
Zohoor, M., Besharati Givi, M.K., Salami, P., 2012. Effect of processing parameters on fabrication of Al–Mg/Cu composites via friction stir processing.
Materials and Design 39, 358–365.

F
OO
PR
D
TE
EC
RR
CO
UN

View publication stats

You might also like