You are on page 1of 10

Wear 252 (2002) 832–841

A comparison of the reciprocating sliding wear behaviour of steel based


metal matrix composites processed from self-propagating
high-temperature synthesised Fe–TiC and Fe–TiB2 masteralloys
C.C. Degnan, P.H. Shipway∗
Advanced Materials Group, School of Mechanical, Materials, Manufacturing Engineering and Management,
University of Nottingham, University Park, Nottingham NG7 2RD, UK
Received 2 November 2001; received in revised form 25 February 2002; accepted 12 March 2002

Abstract
Steel matrix particulate composites were processed by direct addition of various powders to molten medium carbon steel. Fe–TiC and
Fe–TiB2 powders were produced using a self-propagating high-temperature synthesis (SHS) reaction and consisted of a dispersion of fine
TiC (5–10 ␮m) and TiB2 particles (2–5 ␮m), respectively in an iron binder.
Addition of the Fe–TiC powder to the steel resulted in the formation of a metal matrix composite containing a homogeneous dispersion
of TiC particles. However, addition of the Fe–TiB2 powder resulted in the formation of a parasitic Fe2 B phase and TiC within the steel
microstructure. In response to this an SHS masteralloy composed of Fe–(50% TiB2 + 50%Ti) was manufactured which, when added to
steel, prevented the formation of Fe2 B and resulted in a composite containing a mixture of TiB2 and TiC particles.
Dry reciprocating sliding wear behaviour of the three composite materials and their unreinforced counterpart was investigated at room
temperature against a white cast iron counterface. Relative wear behaviour of the materials varied as a function of load. In all cases, the
composite manufactured by addition of Fe–TiB2 (yielding Fe2 B and TiC phases in the steel) exhibited wear rates greater than three times
that of the unreinforced alloy. However, improvements in wear resistance over the base steel of up to two and a half times were observed
with the other composites where the desired TiC and/or TiB2 phases were retained in the steel. Scanning electron microscopy has been used
to interpret wear behaviour in relation to both the as-cast microstructures of the composites and the wear scar microstructures observed.
© 2002 Elsevier Science B.V. All rights reserved.
Keywords: Reciprocating sliding; SHS; TiB2 ; TiC; Composites

1. Introduction but more recently studies using copper [7,8] and ferrous
based matrices have been performed [9,10].
It is well established that the incorporation of hard, Conventional production routes used to fabricate these
second-phase particles deliberately added to ferrous matri- composite materials, such as powder metallurgy, are usu-
ces can significantly improve certain material properties. In ally energy and capital intensive and prove difficult to scale
this class of engineering materials, iron-based composites up for large components. Self-propagating high-temperature
containing TiC have received particular attention [1–3]. synthesis (SHS) [11] provides an economical and energy
They exhibit the toughness and machinability associated efficient process route for the preparation of various hard
with conventional alloy steels combined with significant ceramic particles that can be subsequently incorporated in
improvements in hardness and wear resistance. a metallic matrix. Numerous investigators have produced
The addition of titanium diboride (TiB2 ) to metal matrices iron–titanium carbide (Fe–TiC) by the thermal explosion
has also been observed to greatly increase stiffness, hardness mode of synthesis [12,13]; this results in a product consist-
and wear resistance. Historically, much of the work in this ing of TiC particles embedded in an iron matrix. Addition
field has been directed at aluminium based materials [4–6] of this masteralloy material to molten iron or steel results in
the iron binder melting which enables the particles of TiC
to disperse throughout the host metal.
∗ Corresponding author. Tel.: +44-115-951-3760; Analogous SHS routes have been successfully employed
fax: +44-115-951-3764. in the production of Fe–TiB2 masteralloys [14]. However,
E-mail address: philip.shipway@nottingham.ac.uk (P.H. Shipway). previous studies [15,16] have shown that addition of TiB2

0043-1648/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 4 3 - 1 6 4 8 ( 0 2 ) 0 0 0 5 1 - 0
C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841 833

particles to steel results not in the direct dispersion of the


borides throughout the host matrix, but rather in the forma-
tion of iron boride (Fe2 B) and titanium carbide. Controversy
exists regarding the mechanism by which these phases orig-
inate. Sigl and Schwetz [17] and Jungling et al. [18] propose
that C reacts with TiB2 and Fe, which results in the forma-
tion of TiC and Fe2 B. Tanaka and Saito [15] are in agreement
with this but go further, proposing that oxygen (present as
an impurity) can consume Ti and lead to a deviation in TiB2
stoichiometry. This subsequently leads to excess B, which
reacts with Fe to form Fe2 B. On the contrary, other studies
[19] claim that a significant proportion of Ti dissolves in Fe
and is in equilibrium with TiB2 . They conclude that Fe2 B is
essential to retain the stiochiometric composition of TiB2 .
In the present study, Fe–TiC and Fe–TiB2 master alloys
were introduced into a medium carbon, low alloy steel matrix
using a liquid route. Experiments in which excess Ti was
incorporated into Fe–TiB2 master alloy in order to avoid the
formation of Fe2 B were also performed. This paper reports
on the comparative dry reciprocating sliding wear behaviour
of the resulting composites.

2. Experimental details

2.1. Materials

Stoichiometric iron–titanium carbide (Fe–TiC) and


Fe–TiB2 masteralloy powders were manufactured using
combustion mode SHS and contained 70 wt.% titanium
carbide and titanium diboride, respectively. The Fe–TiC
additive powder (Fig. 1a) consists of globular, non-faceted
TiC particles (5–10 ␮m in diameter) surrounded by a ma-
trix of iron. Fig. 1b shows the typical microstructure of
the Fe–70 wt.% TiB2 master alloy powder additive and
consists of a dispersion of 2–5 ␮m faceted TiB2 particles
(darker phase) in a matrix of iron. EDX chemical analysis
of the ferrous binder phase revealed it to contain around
3.5 wt.% Ti. The hexagonal based morphology exhibited by
the borides is consistent with the crystal structure of TiB2 ,
i.e. hexagonal. Fig. 1. SHS derived powders composed of (a) Fe–70 wt.% TiC, (b)
Fe–70 wt.% TiB2 and (c) Fe–70 wt.% (50 wt.% TiB2 + 50 wt.% Ti).
The masteralloy powder containing excess titanium was
composed of Fe–70 wt.% (50 wt.% TiB2 + 50 wt.% Ti) and
is shown in Fig. 1c. Its microstructure is analogous, in many
respects, to that of its stoichiometric counterpart. The mor- liquid medium carbon steel (BS970:080M15) with chemical
phology of the titanium borides, for instance, is comparable, composition, Fe–0.2 wt.% C, 0.25 wt.% Si, 0.7 wt.% Mn,
although their size is more varied, from <0.5 to 10 ␮m. The 0.06 wt.% S, 0.06 wt.% P. In order to compare like with
matrix surrounding the TiB2 dispersoids, however, is very like, masteralloy additions to the steel were made on a mole
inhomogeneous. Areas of almost pure titanium metal exist percent basis. Addition levels of 3.5 mol% TiB2 and TiC
alongside iron titanium intermetallics and mixtures which were used, which equate to approximately 4.3 and 3.7 wt.%,
encompass a broad spectrum of stoichiometry. This can be respectively. The same mol% TiB2 level was maintained for
mainly attributed to the titanium diluting the pre-mix, result- the addition of the masteralloy containing 50 wt.% excess Ti.
ing in lower combustion temperatures, incomplete melting Composite processing was performed in a solid state
and unstable combustion characteristics. induction furnace (Inductalec, Sheffield, UK, 300 kHz) in
The composite materials used for wear testing were recrystalised alumina crucibles. Masteralloys were incorpo-
produced by adding the SHS masteralloy powders to a rated into the steel by directly pouring the powders onto the
834 C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841

top of the molten steel and stirring with a ceramic impeller. set-up generates a sinusoidal motion with an amplitude
This was undertaken within the confines of the furnace’s of 40 mm.
melting chamber under a protective argon atmosphere. The
melt was held at 1600 ◦ C for 20 min after which time the 2.3. Test procedure and characterisation
furnace was switched off and the composite allowed to
solidify in the crucible. A more detailed description of the The experiments were performed under dry sliding condi-
preparation of the additive powders and the method by tions at a room temperature of 22–24 ◦ C with a relative hu-
which they were incorporated into the steel matrix are given midity of about 50%. All wear experiments were carried out
elsewhere [20]. at a frequency of 2.5 Hz with a mean velocity of 100 mm s−1
Wear tests were performed on pins (15 mm long × 6 mm over a sliding distance of 1000 m. The normal load was var-
diameter) which were machined from the composite mate- ied from 150 to 600 N (corresponding to nominal pressures
rials in the as-cast condition. The rubbing surface of the pin of 5.3–21.2 MPa, respectively).
was ground and polished to a finish of between 0.25 and After the tests, the change in pin length and the corre-
0.3 ␮m Ra prior to wear testing. In order to compare the sponding weight loss were recorded. These data were sub-
wear resistance of the base steel with the composites, unre- sequently used in the calculation of wear rates presented in
inforced pins in the same condition were also manufactured. Section 3 of this paper. The worn surfaces of the pins and
The bulk hardness of the TiC, TiB2 and the Ti/TiB2 com- wear debris were examined using a Philips XL30 environ-
posite materials was measured to be 230 HV, 298 HV and mental scanning electron microscope equipped with field
245 HV, respectively. The counterface plates were machined gun emission (ESEM FEG) and an energy dispersive X-ray
from a high chromium, white cast iron (nominal composi- (EDX) analyser. Cross-sections taken through the wear sur-
tion: Fe–2.7 wt.% C, 0.5 wt.% Si, 0.7 wt.% Mn, 27.0 wt.% face of the pin, parallel to the direction of sliding were also
Cr, 1.0 wt.% Ni, 0.5 wt.% Mo, 0.1 wt.% Cu, 0.007 wt.% examined in order to aid the identification of the surface
Al, 0.01 wt.% S, 0.03 wt.% P) with a measured hardness degradation process and the dominant wear mechanisms.
of 810 HV. Again, the surface of the plate was ground Phases present in the composites and the nature of the wear
to a roughness of between 0.25 and 0.3 ␮m Ra prior to debris generated during testing were determined by X-ray
testing. diffraction using a Siemens D500 diffractometer with Cu
K␣ radiation.
2.2. Test equipment

Wear tests were performed on a BICERI pin on plate 3. Results


type sliding machine which was linked to a data logging
system. This apparatus, shown schematically in Fig. 2, 3.1. Composite microstructure
consists of a stationary pin sliding on a reciprocating plate
with its axis perpendicular to the direction of motion. The Fig. 3a shows the as-cast 3.5 mol% TiC steel matrix com-
plate is fixed by a holder, which is mounted on a steel posite which consists of a homogeneous dispersion of car-
heating block, which in turn is supported on a sliding bed. bides in a matrix consisting of ferrite and a small quantity
The bed, guided by two cylindrical bars, is driven by a of pearlite. A back scattered electron image of the as-cast
crank mechanism. Wear pins were fixed in a loading arm composite material formed when Fe–TiB2 (3.5 mol% TiB2 )
to which static weights can be added or removed. This master alloy powder was introduced into the steel is shown
in Fig. 3b. The microstructure consists of a large quantity of
interdendritic eutectic phase together with small quantities
of fine particulate (located within or close to the interface
between the eutectic and the steel dendrites). Larger partic-
ulates can also be within the dendrites of the steel matrix.
Quantitative EDX chemical analysis and X-ray diffraction
revealed the eutectic phase to be Fe2 B whilst the particles,
both large and small, were identified as TiC. Using both these
techniques, no evidence of TiB2 was detected in the compos-
ite structure at this level of master alloy addition. The matrix
material consisted predominantly of ferrite and bainite.
Fig. 3c shows the result of adding the master alloy con-
taining excess 50 wt.% excess Ti, i.e. Fe–70 wt.% (50 wt.%
TiB2 + 50 wt.%Ti) to the medium carbon steel at the
3.5 mol% TiB2 level. The microstructure consists of very
fine dispersion of particles located at interdendritic sites
Fig. 2. Schematic diagram of reciprocating wear machine. where, in samples processed using Fe–TiB2 , the Fe2 B
C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841 835

3.2. Composite wear behaviour

Fig. 4 plots sliding wear rate versus load for the three
composite materials and the equivalent unreinforced steel.
The most striking feature of this graph is the extremely poor
wear resistance exhibited by the composite formed by the
addition of stoichiometric titanium diboride in the form of
Fe–TiB2 . At the lowest test load (300 N) the wear rate of the
TiC reinforced composite, for example, is a almost a fac-
tor of 10 times lower whilst at the highest load this differ-
ence has decreased to around a factor of 7 times. Over the
range 300–400 N, the TiC reinforced composite exhibits the
best wear resistance of all the materials tested. Under these
loading conditions, the difference in wear rates exhibited by
the TiC reinforced composite, the Ti/TiB2 composite and
the unreinforced base steel is relatively substantial. Fig. 5,
showing a plot of the wear rates of only these three materials
illustrates this point. At 300 N, the wear rate of the unrein-
forced steel is almost 1.7 times greater than that of the TiC
composite, whilst that of the Ti/TiB2 composite is more than
2.4 times greater. At 400 N, this ranking is reversed, with the
Ti/TiB2 composite demonstrating better wear resistance (a
factor of 1.5 times) than the unreinforced steel. At the higher
loads of 500 and 600 N, however, only very small differences
in wear rate were apparent between the TiC composite, the
Ti/TiB2 composite and the unreinforced base steel.
Fig. 6 shows cross-sections through unreinforced and
TiC reinforced composite specimens tested under a load of
300 N. The oxide layer present on the wear surface of the
composite material is much thinner and more compacted
than that observed on the unreinforced material (Fig. 6b)
and was found to contain significant quantities of chromium
as well as iron and oxygen (identified by EDX analysis).
These so called “glazes” consist of very fine, crystalline ox-
ide particles (typically 10–100 nm in size) which have been
compacted onto either the composite’s surface, pre-existing
surface oxide layers, or compressed coarse wear debris.
Stott and Wood [21] have described the development of
such oxide glazes on sliding metal surfaces in detail. The
thicker layer observed on the unreinforced steel (Fig. 6a)
contains no chromium and has an analysis consistent with
iron oxide. Fig. 7, showing a prominent TiC particle, in-
dicates that at a load of 300 N, a percentage of carbide is
retained at the wearing surface which does not undergo
Fig. 3. Composite microstructures formed by the addition of (a)
fracture and/or pull out. A limited number of regions of thin,
Fe–70 wt.% TiC, (b) Fe–70 wt.% TiB2 and (c) Fe–70 wt.% (50 wt.%
TiB2 + 50 wt.% Ti) masteralloy powders to 0.2 wt.% C steel. compacted oxides, similar to those observed on the TiC re-
inforced composite, were also evident on the surface of the
composite containing the mixture of TiB2 and TiC parti-
phase was located. These particles were identified, by EDX cles (i.e. the Ti/TiB2 masteralloy addition). These features,
analysis, to be a mixture of TiC and TiB2 . Larger particles shown in plan view in Fig. 8, were typically 100–200 ␮m
were also evident in moderate quantities throughout the in size and were sparsely distributed over the pin’s wearing
microstructure and found to be almost exclusively TiB2 . surface. EDX analysis again revealed them to contain traces
XRD of this composite material confirmed the presence of of Cr. In a number of areas where the base composite could
TiC and TiB2 and, as expected, did not reveal any Fe2 B to be discerned through the layer of oxide particles adhered
be present. The steel matrix material consisted entirely of to the surface, striations indicative of third body abrasion
ferrite. occurring were evident (see Fig. 9).
836 C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841

Fig. 4. Plot of wear rate against applied load for composite and unreinforced materials.

The surface of the composite pin processed using of fracture, was also evident in various locations on the
Fe–TiB2 , containing Fe2 B and TiC as its predominant sec- wearing surface of the pin (see Fig. 11).
ond phases and tested at 300 N, exhibited large areas of Examination of the TiC, Ti/TiB2 and unreinforced com-
exposed metal where material had been removed as thin posite pins, tested at the higher loads of 500 and 600 N, again
sheets in several successive layers (see Fig. 10). Crack- revealed the presence of compacted oxide glazes on their
ing, which had an appearance indicative of a brittle mode wearing surfaces. These islands of glaze were much larger

Fig. 5. Plot of wear rate against applied load for TiC and TiC/TiB2 composites and unreinforced steel.
C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841 837

Fig. 7. TiC particle retained on the worn surface of a TiC composite


tested under a load of 300 N.

The wear surface of the stoichiometric TiB2 composite


tested at 500 and 600 N was similar to that of the equivalent
specimen tested under a load of 300 N, i.e. cracking of the
surface and pits where metallic sheets had been removed.

4. Discussion

The poor wear performance over the entire range of load-


Fig. 6. Cross-sections through (a) an unreinforced pin and (b) a TiC
ing employed in this study of the composite material formed
composite pin, both worn under a load of 300 N.
by the addition of Fe–TiB2 powder to 0.2 wt.% carbon steel
can be attributed directly to the formation of Fe2 B and TiC
than those observed on the specimens tested at 300 N. Glaze in the microstructure. The reaction to form TiC, which has
features measuring around 1 mm across were not uncommon been observed by other workers [17,18], consumes carbon
(see Fig. 12) and interestingly, in the case of the compos- from the steel and leads to a reduction in the hardness and
ite materials, retained some TiC particles in their structure strength of the matrix. The simultaneous formation of sig-
(see Fig. 13). It is believed that these particles were initially nificant quantities of brittle and comparatively soft Fe2 B at
removed from the bulk material along with the oxide de- interdendritic locations also seriously degrades the ability of
bris and trapped in the glaze material during its formation. the material to resist wear (hardness of Fe2 B was measured
XRD analysis of the debris collected from these high load at 820 kgf mm−2 compared to TiB2 which has a reported
wear tests revealed it to contain predominantly iron oxides hardness of around 3600 kgf mm−2 ). The cracks observed
(Fe2 O3 and Fe3 O4 ) along with, in the case of the compos- on the surface (Fig. 11), which lead to the formation and
ites, quantities of Cr2 O3 . Significantly, unoxidised metallic removal of large sheets of material during the wear process,
flakes were visually observed in the debris and these were are believed to initiate and propagate through this phase. The
confirmed by XRD to consist of steel. mechanism by which this composite wears does not change

Fig. 8. Island of glaze on the surface of a composite containing TiC and TiB2 , worn under a load of 300 N.
838 C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841

Fig. 9. Striations present on the surface of a composite containing TiC and TiB2 , worn under a load of 300 N.

Fig. 10. Surface of a composite containing Fe2 B and TiC, worn under a load of 300 N, showing material removal in the form of layers.

over the range of test conditions employed in this study. this field have observed similar improvements and proposed
It is only the rate at which these processes take place that various mechanisms to explain this behaviour. In the main,
increases with applied load. increased wear resistance of TiC composites is attributed to
At the lowest level of loading (300 N), the TiC reinforced load bearing by the hard second phase particles [22–25] and
steel demonstrated significant improvements in wear resis- the formation of protective glazes [21,26,27], formed by
tance over its unreinforced counterpart. Many workers in the compaction of oxidised material. In this study, at 300 N,

Fig. 11. Cracks detected on the surface of a composite containing Fe2 B and TiC, worn under a load of 300 N.
C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841 839

Fig. 12. Large islands of glaze present on the surface of a composite containing TiC and TiB2 , worn under a load of 600 N.

these glaze type features were observed on the surfaces of Any hard particles, removed under the action of sliding ei-
the TiC and TiC/TiB2 composites as well as on the unre- ther by plucking, cracking or oxide engulfment, would, if
inforced steel. However, these areas were small in size and trapped at the wear interface, lead to preferential third body
number, only partially formed and would contribute little to abrasion of the regions in the microstructure without hard
the overall wear resistance of the respective materials. It is particles present. This effect, illustrated in Fig. 9, may pro-
the load bearing capability of the TiC retained at the sliding vide some explanation for the poorer wear performance of
surface of the composite and its ability to reduce metal to this composite at 300 N compared to that of its TiC rein-
metal contact during sliding that is directly responsible for forced and base steel counterparts.
the improved wear performance of the TiC composite at this At a load of 400 N, increases in both pressure and quan-
low level of loading. tities of wear debris generated become more dominant in
Given that similar quantities of hard second phase parti- governing wear rates. The hard particles present in the
cles are present, it may be expected that the composite ma- composite materials gouge the counterface disc, providing
terial containing the mixture of TiC and TiB2 would display material for glaze formation, whilst the increased pressure
analogous improvements to that of its TiC reinforced equiv- more readily leads to the formation of tenacious glazes. At
alent. However, this was not observed to be the case and this load, it is predominantly these glazes which take the
its wear rate was observed to be worse than both the TiC bulk of the applied load and not the hard, second phase par-
composite and the unreinforced steel. Unlike the TiC rein- ticles. Consequently, particle removal, leading to third body
forced composite, the dispersion of the hard reinforcement abrasion in the case of the inhomogeneous TiC/TiB2 com-
particles present in this composite was inhomogeneous and posite, is greatly reduced. The unreinforced material, on the
tended to be concentrated around the steel’s interdendritic other hand, generates less debris (giving rise to less mate-
regions. Wear on particles at this load is relatively low, but rial being available for the formation of glazes) and has no
eventually they will be worn down, fractured or pulled out. particles present for load bearing capability. When the small

Fig. 13. TiC particles retained within glazes: composite containing TiC and TiB2 and worn under a load of 600 N.
840 C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841

amount of glaze material that does initially form is worn in wear resistance over both the unreinforced steel and
away or removed under the action of the higher stress acting the composite containing the TiC/TiB2 mixture. It is the
at this load, wear of the base material proceeds unchecked load bearing capability of the TiC and its ability to re-
until such times as enough wear debris is generated to re- duce metal to metal contact during sliding that is directly
form them. This debris can only originate from the pin ma- responsible for this improved wear performance. The
terial itself and leads to the comparatively high wear rates comparatively poor wear resistance of the TiC/TiB2 com-
observed. posite can be attributed to the inhomogeneous dispersion
At the higher loads of 500 and 600 N, all three materials of these particles within the microstructure.
undergo a period of severe wear at the start of the test where • At the higher levels of loading utilised (500 and 600 N)
metallic debris is produced and wear rates are high. The the presence of oxide glazes, formed from the comminu-
quantity of debris produced in this period will be similar for tion of metallic debris generated during severe wear at
all three materials since, in this mode of wear, hard particles the start of testing, dominate the wear behaviour of the
have little influence in preventing material removal. During TiC, TiC/TiB2 composites and the unreinforced steel.
the initial stages of testing much of this material is removed Under this regime, little difference in wear rate of these
from the wear interface under the action of sliding (evident three materials was evident.
from the metallic flakes observed in the wear debris after
completion of the test) but some will be retained between the
wearing surfaces. Under the influence of the high pressure Acknowledgements
at these loads, the trapped metallic debris will be quickly
comminuted to a size where it undergoes almost spontaneous The authors acknowledge funding for this research from
oxidation. Localised temperatures at the interface will be the EPSRC (Grant No. GR/M13954).
high and this will further increase the rate of oxidation.
At this level of loading the glazes form rapidly from the
References
oxide generated and remain at the wear interface for the
duration of the test. There is no metal to metal contact at this
[1] T.Z. Kattamis, T. Suganuma, Solidification processing and tribolo-
stage and the corresponding wear rates for all three materials gical behaviour of particulate TiC-ferrous matrix composites, Mater.
decreases. Sci. Eng. A 128 (1990) 241–252.
The very small differences in overall wear rate exhibited [2] V.K. Rai, R. Srivastava, S.K. Nath, S. Ray, Wear in cast titanium
by these materials at high loads may be due to TiC (or carbide reinforced ferrous composites under dry sliding, Wear 231
(1999) 265–271.
TiB2 ) particles observed in the glazes acting as third body
[3] R.K. Galgali, H.S. Ray, A.K. Chakrabarti, Preparation of TiC
abradants and removing glaze material or acting as points of reinforced steel composites and their characterisation, Mater. Sci.
weakness from which the glaze material can be preferentially Tech. 15 (1999) 437–441.
removed. [4] A.V. Smith, D.D.L. Chung, Titanium diboride particle-reinforced
aluminium with high wear resistance, J. Mater. Sci. 31 (1996) 5961–
5973.
[5] S.Q. Wu, H.G. Zhu, S.C. Tjong, Wear behaviour of in situ Al-based
5. Conclusions composites containing TiB2 , Al2 O3 and Al3 Ti particles, Metall.
Mater. Trans. A 30 (1999) 243–248.
• The addition of an SHS produced masteralloy, containing [6] S.C. Tjong, K.C. Lau, Properties and abrasive wear of TiB2 /Al
a dispersion of TiC in an iron matrix, to a medium carbon 4% Cu composites produced by hot isostatic pressing, Comp. Sci.
Technol. 59 (1999) 2005–2013.
steel (0.2 wt.% C) results in the formation of a composite [7] C. Biseli, D.G. Morris, N. Randall, Mechanical alloying of high-
containing a homogeneous dispersion TiC particles. strength copper alloy containing TiB2 and Al2 O3 dispersoid particles,
• The addition of an SHS produced masteralloy, containing Scripta Metall. Mater. 30 (1994) 1327–1332.
a dispersion of TiB2 in an iron matrix, to a the same steel [8] Z.Y. Ma, S.C. Tjong, High temperature creep behaviour of in situ
TiB2 particulate reinforced copper-based composites, Mater. Sci.
results in the formation of a composite containing TiC and
Eng. A 284 (2000) 70–76.
parasitic Fe2 B. In order to avoid this reaction occurring, [9] E. Pagounis, V.K. Lindroos, Processing of particulate reinforced steel
a masteralloy composed of Fe–(50%TiB2 + 50%Ti) was matrix composites, Mater. Sci. Eng. A 246 (1998) 221–234.
manufactured which, when added to steel, prevented the [10] S.C. Tjong, K.C. Lau, Abrasion resistance of stainless-steel
formation of Fe2 B and resulted in a composite containing composites reinforced with hard TiB2 particles, Comp. Sci. Technol.
60 (2000) 1141–1146.
a mixture of TiB2 and TiC particles. [11] A.G. Merzhanov, Self-Propagating High Temperature Synthesis
• The composite material manufactured using Fe–TiB2 ex- Combustion and Plasma Synthesis of High Temperature Materials,
hibited poor wear performance over the entire range of VCH Publications, New York, 1990, p. 1.
loading employed in this study. This behaviour can be di- [12] A. Saidi, A. Chrysanthou, J.V. Wood, J.L.F. Kellie, Characteristics of
rectly attributed to the presence of Fe2 B and TiC in the the combustion synthesis of TiC and Fe–TiC composites, J. Mater.
Sci. 29 (1994) 4493–4998.
microstructure. [13] C.C. Degnan, J.V. Wood, in: P.K. Rohhatgi (Ed.), Proceedings of
• At the lower levels of loading (300 and 400 N), the TiC TMS Fall Meeting on Processing and Applications of Cast Metal
reinforced steel demonstrated significant improvements Matrix Composites, Cincinnati, OH, 1996, p. 317.
C.C. Degnan, P.H. Shipway / Wear 252 (2002) 832–841 841

[14] J.B. Holt, D.D. Kingman, G.M. Bianchini, Kinetics of the combustion [21] F.H. Stott, G.C. Wood, The influence of oxides on the friction and
synthesis of TiB2 , Mater. Sci. Eng. 71 (1985) 321–327. wear of alloys, Tribol. Int. 11 (1978) 211–218.
[15] K. Tanaka, T. Saito, Phase equilibria in TiB2 -reinforced high modulus [22] J.V. Wood, K. Dinsdale, P. Davies, J.L.F. Kellie, Production and
steel, J. Phase Equilib. 20 (3) (1999) 207–214. properties of steel–TiC composites for wear applications, Mater. Sci.
[16] I. Smid, E. Kny, Evaluation of binder phases for hardmetal systems Technol. 11 (1995) 1315–1320.
based on TiB2 , Int. J. Refrac. Met. Hard Met. 7 (1988) 135– [23] O.N. Dogan, J.A. Hawk, Abrasion resistance of in situ Fe–TiC
138. composites, Scripta Metall. Mater. 33 (1995) 953 –958.
[17] L.S. Sigl, K.A. Schwetz, Pressureless sintering of TiB2 –Fe-materials, [24] M. Vardavoulias, The role of hard second phases in the mild
Powder Metall. Int. 23 (1991) 221–223. oxidational wear mechanism of high-speed steel-based materials,
[18] Th. Jungling, L.S. Sigl, R. Oberacker, F. Thummler, K.A. Schwetz, Wear 173 (1994) 105–114.
New hardmetals based on TiB2 , Int. J. Refrac. Met. Hard Met. 12 [25] C.C. Degnan, P.H. Shipway, Elevated temperature sliding wear
(1993-1994) 71–88. behaviour of TiC–reinforced steel matrix composites, Wear 251
[19] A.D. Panasyuk, A.P. Umansky, Phase evolution in the iron–TiB2 (2001) 1444–1451.
system, J. Less-Common Met. 117 (1986) 335–339. [26] J.L. Sullivan, S.S. Athwal, Mild wear of a low alloy steel at a
[20] C.C. Degnan, The processing and wear behaviour of a W(TiC) temperature up to 500 ◦ C, Tribol. Int. 16 (3) (1983) 123–131.
reinforced steel matrix composite, PhD Thesis, University of [27] J. Molgaard, A discussion of oxidation, oxide thickness and oxide
Nottingham, Nottingham, 1995. transfer in wear, Wear 40 (1976) 277–291.

You might also like