You are on page 1of 11

Tribology Letters 8 (2000) 237–247 237

Frictional characteristics of Al–SiC composite brake discs


K. Laden, J.D. Guérin, M. Watremez and J.P. Bricout ∗
Laboratoire d’Automatique et de Mécanique Industrielles et Humaines, Groupe de Recherche en Génie Mécanique, CNRS UMR 8530,
Université de Valenciennes et du Hainaut Cambrésis, BP 311, F-59304 Valenciennes Cedex, France
E-mail: Jean-Paul.BRICOUT@univ-valenciennes.fr

Received 10 November 1999; accepted 31 August 2000

The use of light materials such as aluminium matrix composites reinforced with silicon carbide (SiC) in railway braking devices is
considered. Four quarter-scale discs were produced by the vortex method with two distinct matrices and a SiC reinforcement with two
different shapes and rates. Continuous braking tests (120 s) were run with organic pads in a dry environment. The 390.0 matrix discs
exhibited a higher wear resistance than one produced from a 514.0 matrix. The use of a spherical SiC, instead of an angular one, very
markedly improved the wear resistance of the antagonist materials. During the braking tests, the wear fragments become oxidized and
their presence in tribocontact increases friction and pad wear but decreases the disc wear.
Keywords: railway braking, aluminium alloy, globular SiC, delamination, wear debris

1. Introduction high wear and a marked tendency to seizure. The recourse


to aluminium matrix composites (AMC) and SiC reinforce-
Weight reduction is one of the main objectives of railway ment could reduce these disadvantages. Al–SiC composites
rolling-stock designers. This includes the braking system are able to offer a good mass saving (about 45%) in spite
within the non-suspended masses. Regarding the DUPLEX of the greater disc size made unavoidable because of their
TGV (the new French high-speed train), the brake discs mechanical characteristics and thermal capacities, which are
of the carrying bogie (four per axle) made of forged steel lower than those of steel discs.
(28CDV5 0.8) weigh about 88 kg each and disperse about The present study concerns the production of AMC discs
69% of the total train energy [1,2]. Thus, each disc can be and the analysis of their frictional behaviour in continuous
required to dissipate 23 MJ in the worst situation, corre- braking tests. The tests were carried out with an organic pad
sponding to an emergency stop from 340 km/h on a 1.6% and discs differing in matrix composition and reinforcement
downhill incline. morphology. The braking performances were evaluated in
In order to reduce the mass of the system, a first ap- terms of stability of friction coefficient, surface temperature
proach consists in reducing the number of discs per axle by and wear resistance of antagonist materials.
increasing their thermal dispersion capacity. These “high-
energy discs”, should be able to endure extreme thermome-
chanical stresses. Promising experiments were conducted 2. Description of the materials
in our laboratory on steel discs covered with a NiCr–Cr3 C2
cermet layer coupled with aluminium titanate pads [3]. This 2.1. Production of the AMC discs
system offered remarkable performances in wear resistance
and friction coefficient stability even in wet conditions. The Al–SiC composite discs were produced by the
A second approach consisted in using light materials
“vortex” method. This technique was chosen because of
with high thermal conductivity, such as aluminium alloys.
its moderate cost and versatility. Indeed, thanks to this
This choice allowed significant mass reduction and lower
method, numerous parameters can be controlled, such as
temperature but required recourse to ventilated discs. The
concentration, size, shape and nature of the reinforcement
use of a ventilated device seems unavoidable in the case
for example. The principle consists in mixing the alu-
of low fusion temperature materials (507 ◦ C for Al 390.0)
minium alloy under a neutral gas to a pasty state at 5 ◦ C
and according to Missori and Sili [4], the decrease in sur-
below liquidus, then gradually introducing the reinforce-
face temperature is noticeable, especially when the braking
ment inside the vortex created by a propeller running at
has stopped. The interest in a high thermomechanical con-
about 250 rpm. After mixing, the temperature is increased
ductivity also lies in the reduction of hot spots and tem-
to 750 ◦C and the fluid mixture cast by pouring it into a
perature gradients, and thus in the reduction of buckling
sand mould (figure 1). The optimum parameters in the
and cracking risks [5]. Unfortunately, the frictional behav-
composite production, such as propeller rotation speed and
iour of aluminium alloy discs is not satisfactory because of
temperature during reinforcement incorporation, have to be
∗ To whom correspondence should be addressed. scrupulously observed. Pores are inevitably created by the

 J.C. Baltzer AG, Science Publishers


238 K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs

Table 1
Description and composition of the composites discs.

Disc Brinell λ Matrix Reinforcement


no. hardnessa (W m−1 K−1 ) material type Material type Content Average size Morphology
(HB) (wt%) (µm)

1 50 122.6 514.0 SiC 4 12 Angular


2 64 94.2 390.0 SiC 4 12 Angular
3 62 93.4 390.0 SiC 8 12 Angular
4 67 93.4 390.0 SiC 8 12 Spherical
a Load 31.75 kg.

Table 2
Mechanical and thermal characteristics of the materials.
Material characteristic Pad Steel disc Composites discs
organic 1038 Matrix Reinforcement
514.0 390.0 SiC

Melting temperature (◦ C) – 1530 590 507 2650–2950


r, density (kg m−3 ) 2500 7810 2670 2700 3150
HV (10 g) (kgf mm−2 ) 22–710 200a 74 75 2500
c, specific heat (J kg−1 K−1 ) 670 460 840 860 670
λ, thermal conductivity (W m−1 K−1 ) 4 50 145 110 90
k, thermal diffusivity (m2 s−1 ) 2.39 × 10−6 1.39 × 10−5 6.46 × 10−5 4.74 × 10−5 4.26 × 10−5
a Load 50 g.

ter the particles and to improve wetting, which helps in


reducing liquid–vapour tension. The aluminium–silicon al-
loy (390.0) was chosen for several reasons. Firstly be-
cause, for a content higher than 8–10%, the silicon helps
to limit chemical reactions between Al and SiC and thus
prevents the appearance of fragile Al4 C3 intermetallic com-
pounds, which may alter mechanical characteristics and re-
sistance to corrosion [7–10]. Moreover, this hypereutectic
matrix presents a pasty state necessary for the vortex pro-
duction process and gives a low coefficient of expansion,
Figure 1. Typical production diagram of the Al–SiC composite brake disc. good mechanical characteristics at high temperatures and a
high wear resistance.
vortex during the mixing and casting of the composite, the The choice of the nature, rate and morphology of the
resulting porosity being about 5% by volume. reinforcement is fundamental. SiC particles were chosen
because of their great hardness and relatively low price.
2.2. Disc and pad descriptions Generally, a particular reinforcement improves wear resis-
tance and protects the matrix from the hard “sticky” strain.
Tests using four AMC discs differing in type of ma- The use of SiC short fibres was not envisaged for this appli-
trix, and rate and morphology of the reinforcement, were cation because the high surface temperatures and the severe
carried out. The discs are classified in table 1 and the plastic strains of the disc subsurface would probably induce
main mechanical and thermal characteristics of the materi- breaking of the fibres. Indeed, MODI [11,12] has found that
als used are shown in table 2. For comparison the results short fibres reinforced composites present a lower wear re-
from a 1038 steel disc have also been reproduced. Equiva- sistance than particle-reinforced ones.
lent thermal conductivity values (and analytical method) of The lowest reinforcement rate (4 wt%) is still sufficient
composite materials developed by Laurent [6] are given in to induce a significant gain of wear resistance as noted
table 1. by several authors [13,14]. The differences in morphology
The composition of the matrix and the morphology of between angular and spherical SiC are clearly seen in the
the SiC reinforcement are determining factors in the per- micrographs shown in figure 2.
formances of AMC discs. Two aluminium alloys (514.0 An organic pad designed by the FLERTEX Company
and 390.0) were tested. The choice of alloy is fundamen- was used for the braking tests. It was the only one that
tal in ensuring stress transfer and support of the reinforce- presented acceptable levels of friction coefficients under
ment. The chemical composition of these alloys is shown water spraying (0.25) when the first AMC discs were stud-
in table 3. Magnesium is known for its capacity to scat- ied [15,16].
K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs 239

Table 3
Chemical compositions of the aluminium alloys (wt%).

ASTM AFNOR ISO Si Fe Cu Mn Mg Cr Ni Zn Ca P Pb Sn Ti

514.0 AG3T Al–Mg3 0.5 0.5 0.1 0.5 3 – 0.05 0.2 – – 0.05 0.05 0.15
390.0 AS18UNG Al–Si18 17.6 0.5 1.25 0.13 1.07 0.02 0.92 0.16 0.0003 0.0149 0.01 <0.01 0.02
a (Cu, Ni, Mg)
a Supplied data.

The scale factor S is given by

λd λp
Ad1 √ + Ap1 √
P1 ad ap
S= = , (2)
P2 λd λp
Ad2 √ + Ap2 √
ad ap

which is satisfied when


Ad1 Ap1
S= = . (3)
Ad2 Ap2

The standard TGV-A pad is composed of nine sintered


iron–copper cylindrical blocks of 40 mm diameter fitted
to a steel sheet; two of these pads are applied per TGV-A
disc face. The 1/4-scale disc braking stand requires only
one block, which makes current material characterization
easier and leads to a scale factor of 18. Moreover, the
reduced-scale disc radius, fixed at 80 mm, allows the ratio
of disc-exchange surface to disc-friction track to be equal
to 1.6; this is similar to that of the actual TGV brake sys-
tem.
In order to respect the 1/4-scale reduction, the steel disc
thickness was fixed at 11.25 mm while the thickness of
the AMC discs was calculated so as to obtain the same
thermal capacity as the steel disc (MCp = 796 J K−1 ); it
Figure 2. Optical micrographs of bulk composite brake disc: Al–Si rein-
was fixed at 15.75 mm. The thickness is important, since
forced with 8 wt% angular (a) and spherical (b) SiC. for relatively short braking tests, which was the case in this
study, surface temperature decreases almost proportionately
with the thermal capacity of the disc.
2.3. The braking tests
Simulations of 120 s continuous braking tests were car-
ried out on a specially designed brake stand (figure 3). The
In order to study frictional materials from tests using
160 mm reduced disc was driven in rotation at constant
1/4-scale discs, it is essential to reproduce a temperature at
speed, the sliding velocities being variable up to 17 m s−1
the interface identical to the one found in the actual TGV
and the 40 mm diameter cylindrical pad was applied to
braking system. This implies the identity
the disc with four constant pressures (0.15, 0.3, 0.45 and
P1 P2 0.6 MPa). The test process included a running-in of the
= , (1) pad against the disc in order to accommodate the surfaces
λd λp λd λp
Ad1 √ + Ap1 √ Ad2 √ + Ap2 √ and thus obtain a maximal contact area between pad and
ad ap ad ap
disc [17].
in which a and λ denote, respectively, thermal diffusivity The measurements taken were: temperature inside the
and thermal conductivity of the materials, A is the contact pad (Tg5), temperature at the reverse side of the disc (Td),
area (Ad = 0.01319 m2 and Ap = 0.00126 m2 ), subscripts torque and wear of both pad and disc.
1 and 2 denote the actual and reduced braking systems, The temperature measurement inside the pad was ob-
respectively, and subscripts p and d denote pad and disc, tained with a K-thermocouple inserted 5 mm from the fric-
respectively, and where P is the total heat rate generated tion surface. The temperature of the reverse side of the disc
in the brake system (or braking power), was also measured with a bichromatic radiation pyrometer.
The frictional force was obtained from a torque tranducer
P = Pd + Pp . fitted into the rotating line. From the mechanical analysis
240 K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs

Figure 3. Representation of the braking stand for reduced scale discs: 1 – flexible coupling, 2 – torque transducer, 3 – transmitting shaft, 4 – radiation
pyrometer, 5 – 1/4 scale disc, 6 – pad, 7 – thermocouple and 8 – spring support device.

Figure 4. Wear of pads and discs (1–4) after the 120 s braking tests versus dissipated power. Wear of (—) composite and (- - -) reference steel discs
(µm), wear of (—) pad/composite and (- - -) pad/reference steel discs (µm ×100).

of the disc–pad system, the average friction coefficient can designed airtight plexiglass box surrounding the braking
be evaluated as being device.
Cm
µm = , (4)
pAp Rm 3. Results
where Cm is the average torque, Ap the pad area, p the con- 3.1. Disc and pad wear
tact pressure and Rm the average radius, i.e., the distance
between the disc axle and the application point correspond- The wear curves after 120 s continued braking tests in
ing to the average frictional torque (calculated at a value of a dry atmosphere are shown in figure 4. The curves per-
53.5 mm). taining to the steel disc are plotted as a dotted line for
Pad and disc wear were characterized by the loss of comparison only.
thickness H deduced from weighing measurements before Significant wear rates are observed for disc 1 and its
and after each test. This method allows a comparison of pad from the lowest dissipated powers (figure 4(a)). The
the wear of materials whatever their densities. recourse to the harder and more resistant 390.0 matrix very
The whole dissipated power P (W) was calculated as markedly improved the wear resistance of disc 2 and pad
being for the same reinforcement content, i.e., 4 wt% SiC (fig-
ure 4(b)). Nevertheless, because of its low reinforcement,
P = Cm ω (5)
disc 2 is still subject to the severe wear mechanisms which
with ω, the angular velocity (rad s−1 ). resulted in the formation of deep grooves. The average
Some specific tests were conducted in a non-oxidising arithmetic roughness (Ra) measured on the friction track
atmosphere thanks to injection of argon into a specially of the discs after five braking tests of 120 s carried out at
K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs 241

Figure 5. Optical micrograph of a typical worn surface of disc 2 after


holding braking tests carried at 0.6 MPa.

0.6 MPa and 6.7 m s−1 was about 4.5 µm for disc 1, 2.5 µm
for disc 2 and only 1 µm for the steel disc.
By increasing the reinforcement rate up to 8 wt% SiC
(disc 3), disc wear was noticeably reduced at high power
without an increase in pad wear (figure 4(c)).
The recourse to a spherical SiC is very interesting since
disc and pad wear rates are lower than those obtained with
the steel disc.
After several braking cycles, flattening of the asperi-
ties of the AMC discs track were noticed as well as the
progressive formation of a blackish transfer layer. Never-
theless, the phenomenon was more marked with discs rein-
forced with 8 wt% SiC. Microhardness tests (0.1 N) carried
out on this transfer layer gave hardness levels higher than
that of the matrix in the 145–245 HV range and some val-
ues reached 700 HV. The energy dispersive spectrometry
(EDS) analysis of the disc 2 track (figure 6) showed mate-
rial transfer from the pad, in particular Fe, Zn, Cu and C.
The presence of oxygen resulting from the oxidation of
wear debris was also observed, which is in agreement with Figure 6. Energy dispersive X-ray spectrum of disc 2 before (a) and after
other studies [14,18,19]. Disc 2 was cut through a cross sliding (b).
section perpendicular to the sliding direction with a low
cutting speed and oil lubrication. It was then polished and
the micrograph obtained is shown in figure 7. A transfer
layer of about 15–20 µm thickness, in which were embed-
ded and scattered fine wear debris and silicon particles, was
observed.
For the high reinforced discs (discs 3 and 4), the black
transfer layer covered the entire friction track. On the
other hand, for the low reinforcement level (discs 1 and 2),
zones uncovered by the transfer layer were visible and
consequently contact with the pad occurred not only on
the layer but also on the composite friction surface. This
explains why wear resistance is rather poor (figure 4 (a)
and (b)).
The scanning electron micrograph of the cross section of
disc 2 after five consecutive braking cycles shows the for-
Figure 7. Optical micrograph of the cross section perpendicular to the
mation of sub-surface cracks around the angular SiC parti- worn surface of disc 2 after five consecutive 120 s holding braking tests
cles (figure 8). Moreover, broken angular particules appear carried at 6.74 m s−1 and 0.3 MPa; formation of a transfer layer composed
in the transfer layer, which reveals the poor toughness of of fine wear debris.
242 K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs

angular SiC reinforcement (figure 9(a)). In contrast, no bro- 3.2. Friction coefficient
ken spherical SiC particles have been observed with disc 4
(figure 9(b)). The variations of the friction coefficient versus sliding
velocity are given in figure 10 for the four discs studied.
The grey area represents the variation of the friction co-
efficient for a conventional steel disc tested under similar
conditions.
Discs 1 and 2 exhibited an average friction coefficient
similar to that of the steel disc (µ ≈ 0.3) and were not
very sensitive to velocity and contact pressure effects (fig-
ure 10 (a), (b) and (d)).
Disc 3 showed a particular behaviour distinguished by a
friction coefficient that was higher at low contact pressures
such as 0.15 MPa (figure 10(c)).
The micrographs of the cross section perpendicular to the
sliding direction (figure 11(a)) obtained after five consecu-
tive 120 s braking tests confirm the presence of a transfer
Figure 8. Scanning electron micrograph of the cross section perpendicular
layer of about 10–20 µm thickness at low contact pressure
to the worn surface of disc 2 after five consecutive 120 s holding braking (0.15 MPa). In contrast, at high contact pressure (0.6 MPa),
tests carried at 6.74 m s−1 and 0.6 MPa; formation of sub-surface cracks the transfer layer is not maintained. This layer is scattered
around the angular SiC particles. and uncovered areas are numerous (figure 11(b)).

Figure 9. Optical micrographs of the third body wear on disc 3 (a) and disc 4 (b); the disc 3 shows broken angular SiC particles; the disc 4 shows
intact globular SiC particles but broken Si particles.

Figure 10. Average friction coefficient during the 120 s braking tests for several discs 1–4 versus sliding velocity (grey shapes represent friction
coefficient obtained with the reference steel disc).
K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs 243

Figure 13. Mean friction coefficient µ obtained with disc 4 at 5.16 m s−1
sliding velocity in air and in argon atmospheres versus typical contact
pressures.

Figure 14. Wear of pad and disc 4 at 5.16 m s−1 sliding velocity in air and
in argon atmospheres versus typical contact pressures (wear of pad must
be changed by a multiplier factor of 100). Disc wear in (◦) argon and
(•) air atmospheres (µm); pad wear in () argon and () air atmospheres
(µm ×100).
Figure 11. Scanning electron micrographs of the cross section perpendic-
ular to the worn surface of disc 3 after five consecutive 120 s holding
braking tests carried out at 6.74 m s−1 and at 0.15 (a) and 0.6 MPa (b).
Transfer layer on sliding track.

Figure 15. Variation of the instantaneous friction coefficient µ due to the


application of sandpaper during the last 60 s of a braking test carried at
5.1 m s−1 and 0.6 MPa with disc 4 and organic pad.

effects on the frictional behaviour. Four braking tests were


run for each contact pressure and for a sliding velocity of
Figure 12. Variation of Vickers microhardness (10 g) of disc 3 matrix with 5.16 m s−1 . As shown in figures 13 and 14, the friction
the distance to the friction surface after five consecutive 120 s holding coefficient and pad wear increase with contact pressure in
braking tests carried at 6.74 m s−1 . an air environment. However, disc wear appears lower in
an air environment (figure 14).
The microhardness measurements performed on disc 3 In order to know the effect of the transfer layer on fric-
are plotted against depth in figure 12. A considerable hard- tional behaviour, sandpaper was applied for 60 s on the
ening of the surface (245 HV) is clearly displayed, partic- disc 4 track during a 6 min braking test carried out at
ularly in the case of low contact pressure (0.15 MPa). 0.6 MPa (figure 15). The removal of the transfer layer
Disc 4 exibited lower friction than disc 3 because of the resulted in a decrease and an instability of the friction co-
SiC spherical morphology. efficient.
Some specific tests were conducted with disc 4 in a non- It has been noted that the transfer layer forms during the
oxidising atmosphere in order to determine the oxidation running-in period; thus, five consecutive braking tests of
244 K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs

the stainless-steel (AISI 52100) pin was observed. On the


other hand, composites with a high weight per cent of non-
metals showed an abrasive wear mechanism on both disc
and pin. These wear mechanisms are not quite compara-
ble to those observed in our study because the phenolic
matrix composite pad is much more sensitive to abrasive
mechanisms than stainless steel.
The most extensive wear rates were obtained with disc 1
(figure 4(a)). It is also surprising that the wear rates are
higher than those obtained in the same conditions with a
non-reinforced disc [24]. So, these results show that the
Figure 16. Variation of the instantaneous friction coefficient µ, Vickers
mechanical resistance of the 514.0 matrix is not sufficient
microhardness (10 g) and mean average roughness Ra measured on disc 2 to retain the SiC particles. The free SiC particles form an
sliding track during the running in period (microhardness and roughness abrasive third body, which results in a severe abrasive wear
must be, respectively, changed by a multiplier factor of 300 and 3) (—) µ, of pad and disc. Moreover, according to some researchers,
(◦) HV (×300) and () Ra (µm ×3). the formation of fragile Al4 C3 composites resulting from
the reaction between non-oxidized SiC and aluminium dur-
120 s were performed with disc 2 at 0.6 MPa. The evolu-
ing the production process would be inevitable beyond a
tions of the friction coefficient, the surface microhardness
certain temperature and time and would lead to the deteri-
and the roughness versus the sliding distance are illustrated
oration of the reinforcement-matrix cohesion [7–10,25].
in figure 16. The friction coefficient presented a high value
Thanks to the higher mechanical resistance of the 390.0
(0.9) during the first 20 s of test (135 m) and then a high
matrix, the wear resistance of disc and pad is significantly
decrease to level off at about 0.3. Concerning the surface
improved. Moreover, the silicon present in the 390.0 matrix
roughness, its value increased during the first 1000 m of
helps to limit chemical reactions between Al and SiC and
sliding up to 3 µm then decreased progressively during the
thus prevents the appearance of fragile Al4 C3 intermetallic
next tests down to about 2.5 µm.
compounds.
The increase of the reinforcement rate up to 8 wt% SiC
4. Discussion (disc 3) results in an increase of the wear resistance of
the disc beyond 2300 W (figure 4(c)). This behaviour is
4.1. Disc and pad wear logically explained by the increase of mechanical properties
and by the fact that the more numerous SiC particles are
Except for disc 1, the variation of pad wear as a func- less loaded.
tion of braking power is close to an exponential relationship The cracks observed on the cross section of disc 2 after
because of the thermal degradation of the pad (and more five consecutive braking cycles (figure 8) are close and par-
particularly the phenolic resin) at high temperatures [20]; allel to the worn surface and can result in a delamination
here, it reached about 300 ◦ C for a dissipated power of wear mechanism [26].
1000 W. Therefore, the increase of applied loads and tem- The performance of disc 4 was very interesting. Because
peratures induce severe plastic strains at the disc surface of the spherical morphology of SiC particles, the friction
which probably accentuate disc wear (figure 5). Accord- mechanism may be regarded as sliding rather than plough-
ing to Martinez [21,22], who has also observed some high ing, as observed with angular SiC particles. In the latter
plastic strains with several composite materials and ma- case, the sharp corners of the SiC result in an abrasive fric-
trices (Al–Si matrices and SiC reinforcement) during slid- tion mechanism and the risk of reinforcement pulling out
ing against a stainless-steel antagonist, it results in the ap- is strongly increased [27].
pearence of delamination subsurface cracks, leading to an Moreover, broken angular particules which appear in the
increase in the wear rate. transfer layer reveal the poor toughness of angular SiC re-
Even if it is difficult to determine precisely the type inforcement (figure 9(a)). The severity of microploughing
of wear mechanism, the deep grooves observed on the disc produced by the free hard broken SiC particles is therefore
track area show that abrasion is dominant. This is explained strongly amplified. The higher ductility [28] and fracture
by the abrasive behaviour of the phenolic matrix composite toughness [29] of the AMC with spherical SiC results in
pad, which contains some hard particles (710 HV max). less damage when high plastic deformations occur.
Moreover, when damage occurs to the matrix, these pad
particles are not maintained any more and form an abra- 4.2. Friction coefficient
sive third body. By using the “vortex” production process,
Hosking [23] showed that AMC materials present different Discs 1 and 2 exhibited a friction coefficient not very
frictional behaviours depending on the reinforcement rate; sensitive to velocity and contact pressure effects (fig-
thus the wear mechanism of the low reinforced composites ure 10 (a), (b) and (d)). As noted before, the SiC particles
was consistently adhesive in nature; aluminium transfer to in discs 1 and 2 were not numerous enough to support the
K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs 245

friction stresses and the abrasive nature of the broken parti- hardness of the wear fragments increases greatly, and par-
cles added to the delamination process described above (fig- ticularly so for aluminium, where the increase is more than
ure 8). The continuous regeneration of the friction surfaces thirty times higher after oxidation [31], making them more
explains the quite stable friction coefficient. The results abrasive and increasing friction and pad wear. In contrast,
are in agreement with Straffelini’s work [30]; the lowest disc wear appears lower in an air environment because the
friction coefficients were found in the cases in which the oxidised wear fragments are embedded in the transfer layer,
AMC showed the highest wear rates. which protects the disc from wear (figure 14).
Some factors showed that the formation of a transfer
layer results in an increase of the friction coefficient. In- 4.3. Surface temperatures
deed, as regards disc 3, the formation of a transfer layer at
low contact pressure (0.15 MPa) has been observed and the Theoretical surface temperatures can be calculated from
friction tends to be high. Observations of the cross section the physical characteristics of the materials [3,33,34]. If
of this disc showed that wear fragments are embedded in the we consider disc and pad as two semi-infinite solids in fric-
transfer layer. Moreover, a clear hardening of the surface tional contact, division of heat between disc and pad may
is distinctly displayed (245 HV). The matrix strain hard- be deduced from the equation of Carslaw and Jaeger [35]
ening cannot explain this effect, otherwise the maximum and the temperature at the interface may then be expressed
hardening would be observed for higher contact pressures. as
It is likely that the large oxidised wear fragments are re- √
2 P t
sponsible for the surface hardening, thereby increasing the T (0, t) = √ , (6)
π λd λp
friction coefficient. At high contact pressure (0.6 MPa), the Ad √ + Ap p
kd kp
stresses are too high so that the transfer layer is not main-
tained. In that case, the layer is scattered and uncovered where k, λ are, respectively, thermal diffusivity and con-
areas are numerous and friction tends to be low. Moreover, ductivity and P the total absorbed power. The friction track
it is probable that the layer becomes viscous with the rise area of the disc (Ad ) and the pad (Ap ) are 13.19×10−3 and
in temperature, which reduces friction. 1.256 × 10−3 m2 , respectively. As the steel and AMC discs
Lastly, the removal of the transfer layer obtained thanks and the pad are 11.25, 18.75 and 20 mm thick, respectively,
to the application of sandpaper on the friction track resulted the semi-infinite solid assumption must be considered as
in a decrease and instability of the friction coefficient and, approximate and only valid at the first step of the braking
according to other researchers [31,32], the phenomenon ap- test. For continuous braking, the constant heat rate assump-
pears frequently with clean metal. Thus, the presence of tion seems adequate. It should be noted that this equation
this layer appears essential since it stabilizes the friction disregards thermal exchange between the materials and the
coefficient. It has been noted that the transfer layer forms exterior environment.
during the running-in period. The high friction coefficient The pad surface temperatures measured at 2500 W were
value (0.9) measured during the first 135 m of test is due much higher than those predicted from equation (4) (ta-
to the severe wear mechanisms. The building of the trans- ble 4). Two principal factors are responsible for this phe-
fer layer contributes to the increase of friction since the nomenon: true contact area being clearly lower than the
aluminium tends to cover the wear fragments embedded geometrical area and thermocouple location. Tirovic and
in the superficial layer. This process is favoured by the Day [36] showed that for a standard organic friction mater-
moderate hardness of the disc (∼85 HV), the wear debris ial, only 75% of the pad surface area is in contact, and for a
being easily embedded in the matrix. The flattening of the stiff friction material this was reduced to 60%. By consid-
asperities of the disc track is revealed by the decrease in ering that the true contact area is about 60% of the geomet-
the surface roughness, and the formation of the compact rical area, then the gap between the experimental and the
layer (∼155 HV) then induces a decrease in the friction theoretical results becomes insignificant. Lastly, the single
coefficient down to about 0.3. K-thermocouple gives a localized surface temperature and
Disc 4 exibited a lower friction than disc 3 because of not an average one. The thermocouple is inserted at the
the SiC spherical morphology resulting in a sliding contact. center of the pad (figure 17), which is a zone presenting a
Indeed, disc 3 showed the highest values of the friction higher surface temperature than average. Indeed, the fric-
coefficient (up to 0.5); the high angular SiC content (8 wt%)
Table 4
logically induces an increase in the abrasive behaviour.
Thermal results obtained at 2500 W dissipated power with
The oxidation of the wear fragments is a determining steel disc and 390.0 matrix disc.
process, as for the friction and wear mechanisms. Some
Steel 1038 AMC
specific tests showed that the friction coefficient and pad
wear increase with contact pressure in an air environment MCp (J K−1 ) 796 798
and were higher than in a non-oxidising atmosphere (ar- Experimental Ts (◦ C) 325 266
Calculated Ts (◦ C) 192 157
gon). The wear fragments are oxidised during high-pressure Real surface/apparent surface 0.59 0.59
braking because of induced high temperatures, the maximal in order to have Ts(exp) = Ts(cal)
values reaching about 350 ◦C under 0.6 MPa. Thus, the
246 K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs

casting would have limited this porosity considerably. The


squeeze casting technique, which is recognized for its ca-
pacity to limit porosity and to improve the wetting of the
reinforcement, would seem to be an interesting alternative.
A reduction in mass of the device (around 45%) could
be obtained in spite of an increase of the disc thickness
made necessary by the low mechanical properties and by
the lower thermal capacity compared to steel.

5. Conclusions
Figure 17. Representation of the pad and localization of the thermocouple. Four AMC discs differing in matrix composition (514.0
and 390.0), in reinforcement morphology (angular and
spherical) and in reinforcement rates (4 and 8 wt%) were
made by the “vortex” method and tested in continuous brak-
ing against an organic pad. The braking performances were
estimated in terms of stability of friction coefficient, surface
temperature and wear resistance of antagonist materials.
The tribological behaviour of AMC discs, compared
with the classical steel discs shows lower friction coeffi-
cients and higher wear rates. The 514.0 matrix disc induced
high abrasive wear because of poor mechanical resistance
not permitting efficient maintenance of the SiC particles.
For the same reinforcement rate, the use of the harder and
more resistant 390.0 matrix very markedly improved the
wear resistance of the antagonist materials.
Figure 18. Maximal temperatures of disc sliding track reached during By increasing the reinforcement rate up to 8 wt% an-
the 120 s holding braking tests in contact with the organic pad versus gular SiC, the disc exhibited a low wear rate, particularly
dissipated power P . at high braking power, and a relatively high friction coef-
ficient (µ ≈ 0.4) because of the high angular SiC content.
tion distance covered by a disc asperity is maximal around With the same SiC content, the spherical SiC reinforced
the center of the pad, and thus temperature is too. disc exhibited a lower friction coefficient (µ ≈ 0.3) due
The temperatures measured at the friction surfaces of to the spherical morphology resulting in a sliding contact
the pads are presented in figure 18. In agreement with instead of a ploughing one. The wear rates were close to
equation (2), the steel disc exibited a linear increase in those of the steel disc; in fact they were slightly lower. The
temperature with increasing braking power. In contrast, the risk of reinforcement pulling out and breaking was strongly
AMC discs exibited non-linear behaviour. Indeed, the rate decreased and the higher fracture toughness of the AMC
of increase in surface temperature tended to fall at higher resulted in lower damage when high plastic deformations
braking powers, with temperatures becoming much lower occurred.
than the equivalent ones for steel; the gap at 2500 W was Some specific tests conducted in a non-oxidising at-
about 60 ◦ C. This behaviour may be explained by the fact mosphere have shown that the presence of oxidised wear
that the thermal conductivity of aluminium increases with fragments in tribocontact increased friction and pad wear
temperature, unlike that of steel. Therefore, the drop in but decreased disc wear because they contributed to the
temperature is probably caused by an increase in the real formation of a blackish transfer layer on the disc friction
surface area of the aluminium alloy, the fusion temperature surface. Indeed, the transfer layer, consisting of oxidised
being relatively low (∼507 ◦C). wear debris coming from the pad and the disc and embed-
Even if the temperatures of the AMC discs were lower ded in the matrix, protected the disc and improved wear
than those of the steel discs, several factors limit the differ- resistance. Moreover, it was noticed that the presence of
ence in temperature such as the presence of third body wear this abrasive transfer layer slightly increased friction.
on its friction surface. Indeed, the low thermal conductivity Because of their relatively high thermal conductivity, the
of the third body wear acts as a heat barrier because of its temperatures of the AMC discs were lower than those of
being partially composed of oxides and elements derived the steel disc, the difference being about 60 ◦ C at 2500 W.
from the pad. For the steel disc, the formation of this layer Al–SiC composites are able to offer a good “mass
is very limited. Moreover, during the production of the economy-effectiveness” compromise in spite of the need
composite, the vortex method creates porosity in the disc for the discs to be thicker because of their mechanical prop-
and induces a heat barrier effect as well (table 1). In our erties and thermal capacities, which are lower than those of
case, the degassing of the molten aluminium bath before steel.
K. Laden et al. / Frictional characteristics of Al–SiC composite brake discs 247

Acknowledgement [15] H. Ruppert, Journées Européennes du Freinage (JEF), Pole Frein,


Lille, 1995, pp. 291–300.
The present research work has been supported by the [16] T. Tsujimura, S. Manabe, A. Watanabe and Y. Kobayashi, in: Proc.
1st Jap. Int. SAMPE Symp. (1989) pp. 1031–1036.
CNRS, the European Community, the Conseil Régional [17] J.D. Guérin, K. Laden, M. Watremez and J.P. Bricout, J. Int. Fran-
Nord Pas de Calais, the Ministère de l’Education nationale cophones de la Société Tribologique de France, Actes de la Société
and the Délégation Régionale à la Recherche et à la Tech- Tribologique de France 2 (1996).
nologie; the authors gratefully acknowledge the support of [18] G.J. Howell and A. Ball, Wear 181–183 (1995) 379.
these institutions. [19] M. James and Jr. Herring, paper 670146 presented at the SAE Auto-
motive Eng. Congr., Detroit, January 1967.
[20] S.K. Rhee and J.E. Byers, SAE Transactions (1972), paper 720930.
[21] M.A. Martinez, A. Martin and J. Llorca, Scripta Metall. Mater. 28
References (1993) 207.
[22] A. Martin, M.A. Martinez and J. Llorca, Wear 193 (1996) 169.
[1] J. Raison, Journées Européennes du Freinage (JEF), Pole Frein, Lille, [23] F.M. Hosking, F. Folgar Portillo, R. Wunderlin and R. Mehrabian,
December 1992. J. Mater. Sci. 17 (1982) 477.
[2] P. Cauwel, Journées Européennes du Freinage (JEF), Pole Frein, [24] K. Laden, J.D. Guérin, J.P. Bricout and J. Oudin, in: Int. Conf. on
Lille, December 1995, pp. 331–336. Advanced Materials EMRS’97, Strasbourg, 16–20 June 1997.
[3] M. Watremez, J.P. Bricout, B. Marguet and J. Oudin, J. Tribol. 118 [25] J.C. Viala, P. Fortier, C. Bernard and J. Bouix, AECM, Bordeaux
(1996) 457. (1985).
[4] S. Missori and A. Sili, Proc. Instn. Mech. Engrs. 202 (1988) 91. [26] A.T. Alpas and J.D. Embury, Metall. Mater. 24 (1990) 931.
[5] P. Dufrenoy, Thèse de Doctorat, Lille, 22 December 1995. [27] M. Bai and Q. Xue, Tribol. Int. 30 (1996) 261.
[6] M. Laurent, M. Lostie and D. Staicu, La Revue de Métallurgie [28] J. Llorca and C. Gonzalez, Key Eng. Mater. 127–131 (1997) 111.
SF2M-JA 97 (1997). [29] L. Lu, J.K.M. Kwok, M.O. Lai, Y.B. Liu and S.C. Lim, J. Mater.
[7] V. Laurent, D. Chatain, N. Eustathopoulos and X. Dumant, in: Cast Proc. Techn. 37 (1993) 453.
Reinforced Metal Composites, ed. A.K. Dhingra, Conf. Proc., ed. [30] G. Straffelini, F. Bonollo, A. Molinari and A. Tiziani, Wear 211
S.G. Fishman (ASM International, 1988) pp. 27–31. (1997) 192.
[8] M. Skibo, P.L. Morris and D.J. Lloyd, in: Cast Reinforced Metal [31] E. Rabinowicz, Friction and Wear of Materials, 2nd Ed. (Wiley–
Composites, ed. A.K. Dhingra, Conf. Proc., ed. S.G. Fishman (ASM Interscience Lavoisier, New York, 1995).
International, 1988) pp. 257–261. [32] Y.M. Chen, J.C. Pavy, B. Rigaut and J.P. Peyre, in: Matériaux et
[9] R. Mitra and Y.R. Mahajan, Bull. Mater. Sci. 18 (1995) 405. Mécanique, Cetim Informations, No. 139 (June 1994) p. 26.
[10] D.J. Lloyd, Composites Sci. Technol. 35 (1989) 159. [33] J.D. Guérin, J.P. Bricout, K. Laden and M. Watremez, Tribol. Lett.
[11] O.P. Modi, B.K. Prasad, A.H. Yegneswaran and M.L. Vaidya, Mater. 3 (1997) 257.
Sci. Eng. A 151 (1992) 235. [34] J. Martinet, Eléments de thermocinétique – Conduction de la chaleur,
[12] N.P. Suh, Wear 44 (1977) 1. Technique et Documentation (Lavoisier, Paris, 1989).
[13] A. Watanabe, Y. Sugai, T. Tsujimura and K. Takao, Int. Conf. Recent [35] H.S. Carslaw and C. Jaeger, Conduction of Heat in Solids, 2nd Ed.
Advances in Science and Engineering of Light Metals, Raselm’91, (Clarendon Press, Oxford, 1958).
Mitsubishi Aluminium Co., Sendai Jap. (1991) pp. 525–530. [36] M. Tirovic and A.J. Day, J. Automobile Eng., Proc. Instn. Engrs.
[14] J. Masounave and L. Scheed, in: Matériaux et Revêtements Com- 205 (1991) 137.
posites Pour Applications Tribologiques, Mécanique et Matériaux,
Finishing Publications of CETIM (1996) pp. 17–36.

You might also like