J Wear 2008 12 020

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Wear 267 (2009) 350–355

Contents lists available at ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Effect of friction stir processing on the tribological performance of


high carbon steel
S.H. Aldajah a,∗ , O.O. Ajayi b , G.R. Fenske b , S. David c
a
United Arab Emirates University, PO Box 17555, Al-Ain, United Arab Emirates
b
Argonne National Laboratories, Argonne, IL 60439, USA
c
Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA

a r t i c l e i n f o a b s t r a c t

Article history: Friction stir processing (FSP) was applied to 1080 carbon steel as a means to enhance the near-surface
Received 23 August 2008 material properties. The process transformed the original pearlite microstructure to martensite, resulting
Received in revised form 9 December 2008 in significant increase in surface hardness. This surface hardening produced a significant benefit for fric-
Accepted 10 December 2008
tion and wear behavior of the steel as measured by unidirectional sliding ball-on-flat testing. Under dry
sliding, FSP reduced friction coefficient by approximately 25% and wear rate by an order of magnitude.
Keywords:
Under oil lubrication, FSP had only a marginal effect on friction, but it reduced wear rates by a factor of
FSP
4. The improvement in tribological performance of 1080 steel by FSP technique is attributed to reduced
Tribology
Friction
plasticity of the near-surface material during sliding contact.
Wear © 2009 Elsevier B.V. All rights reserved.

1. Introduction [4]. This action results in significant microstructural changes in the


processed zone as a result of severe plastic deformation, mechanical
Tribological performance of machine components is often mixing, and thermal exposure of the material. The characteristics
enhanced by various kinds of surface modification techniques. In of FSP have led to several applications for microstructural modi-
steel (which is perhaps the most common tribological material) fication in metallic materials, including enhanced superplasticity
for instance, surface modification can take the form of mechan- [1,2,5–8], surface composites [9], homogenization of nanophase
ical properties enhancement (mainly surface hardening) through aluminum alloys, metal matrix composites [10,11], and microstruc-
surface chemical changes such as carburing, nitriding, etc. In some tural refinement of cast aluminum alloys [12–14]. The near surface
cases, surface hardening can be achieved without a chemical mod- properties enhancement by FSP is expected to have effects on the
ification, but only by thermal treatment, such as heat hardening; tribological properties of the processed surface.
or by mechanical treatment, such as shot peening. Surface harden- Friction stir processing can be applied as a single-pass for pro-
ing is also achieved by the deposition of a distinctive surface layer cessing a small area. For large engineering components in which
(coating) of harder material such TiN, CrN, DLC, etc. Coatings are the contact areas are relatively large, single pass FSP may not be
usually deposited by a variety of processes such as chemical vapor adequate. In such cases, multi-pass FSP with a certain level of over-
deposition (CVD), physical vapor deposition (PVD), thermal spray, lap between the successive passes may be required. For both single
etc. and multi-pass processes aimed at tribological applications, it is
Friction stir processing (FSP) is a new solid-state process- important to assess the microstructural evolution and its influence
ing technique that can be used for surface hardening through on the tribological properties. This paper investigates the impact
microstructural modification [1–3]. This process is an off shoot of of single-pass FSP and multi-pass FSP applied to high carbon steel
the friction stir welding (FSW) process used for solid state joining (AISI 1080) on its friction and wear performance.
of two separate pieces of metallic materials. Basically, FSP involves
the use of a non-consumable rotating tool with a pin and shoul-
der inserted into a single piece of material and traversed along the 2. Friction stir processing
desired path to cover the region of interest as illustrated in Fig. 1
Fig. 2 shows a picture of the friction stir processing of 1080 car-
bon steel plate in progress. Refractory tungsten alloy tool was used
for processing at a rotating speed of 1000 rpm, normal force of 5 kN
∗ Corresponding author. and a translation speed of 15 mm/s. The Ar gas shield is to prevent
E-mail address: s.aldajah@uaeu.ac.ae (S.H. Aldajah). the oxidation of the processed surface layer. Fig. 3 shows the macro-

0043-1648/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2008.12.020
S.H. Aldajah et al. / Wear 267 (2009) 350–355 351

Fig. 3. Friction stir processed 1080 carbon steel sample.


Fig. 1. Schematic illustration of friction stir processing.

Fig. 4. Original pearlite microstructure of 1080 steel.

Fig. 2. Friction stir processing process with Ar gas shield.


2–5 range at a strain rate in excess of 10 S−1 . Such severe and rapid
deformation results in significant heating of the processed layer. In
picture of the processes plate with track 1 being the single pass and addition, significant frictional heating occurs at the rubbing inter-
tracks 3–8 are the multi-pass treatment with about 20% overlap face of the tool shoulder and work piece. These two heat generation
between passes. The FSP produces rapid, severe plastic deforma- processes produce a large increase in the temperature of the pro-
tion in the processed surface layer with plastic shear strain in the cessed region which has been shown to be as high as a 1000 ◦ C

Fig. 5. (a) Cross-sectional macrograph of the FSP processed 1080 steel. (b) Micrograph of processed layer after Nital etching showing martensite.

Fig. 6. A schematic of the four microstructural regions of a FSP sample: (a) unaffected material; (b) heat affected zone (HAZ); (c) upper stir zone; (d) lower stir zone.
352 S.H. Aldajah et al. / Wear 267 (2009) 350–355

during the FSP of high carbon steel [15]. Such a temperature is in


excess of the austentizing temperature for steel.
Cross-sectional metallographic samples were prepared from the
FSP plate to assess the impact of this process on the 1080 steel
microstructure. Fig. 4 shows the microstructure of the original
material which consists primarily of a pearlite phase. Fig. 5a shows
the macrograph of a polished and Nital etched cross-section of sin-
gle pass FSP. Fig. 5b clearly shows that the stir region consists of
martensite as revealed by Nital etching. Indeed, four distinctive
microstructural regions; namely unaffected material, heat affected
zone (HAZ), upper stir zone and lower stir zone, can be identified
as illustrated schematically in Fig. 6. The unaffected material is the
region remote from the FSP area and has neither been deformed nor
thermally modified in terms of microstructure or mechanical prop-
erties. The material in heat affected zone region, which lies closer
to the FSP centre, has experienced a thermal cycle which has mod-
ified the mechanical properties. In the upper stir zone region, the
material has been plastically deformed by the friction stir process- Fig. 7. Lower stir zone of the FSP processed 1080 steel.
ing tool, and the heat from the process exceeded the austenitizing
temperature of the 1080 steel; resulting in the formation of marten- Microhardness measurements were performed for the cross-
sitic phase from the rapid cooling after the FSP tool passes. The sectional samples from both single pass and multi-pass FSP (Fig. 8).
lower stir zone is the region in which the material also experienced In the single pass FSP, the surface hardness, to a depth of about 5 mm
severe plastic deformation and heating with the concomitant phase was about 750 HK (7.25 GPa) which is substantially harder than the
transformation from pearlite to martensite (Fig. 7). initial value of about 330 HK (3.2 GPa). This is expected because of

Fig. 8. Microhardness profiles of single and multi-pass FSP processed 1080 steel.
S.H. Aldajah et al. / Wear 267 (2009) 350–355 353

the formation of a harder martensite in the FSP region compared


to the initial pearlite phase. Because of the very high temperature
achieved by the processed material, it is safe to assume that none of
the hardness increase is due to work hardening from plastic strain.
Indeed, FSP has been shown to be accompanied by dynamic recrys-
tallization in some non-ferrous alloys [7–9]. In the multiple pass
treatment, a slight decrease was observed in the surface layer hard-
ness, 660 HK (6.4 GPa), as a result of tempering that occurred in the
surface martensitic phase during subsequent passes

3. Friction and wear evaluation

Friction and wear performance of FSP 1080 steel surface


were evaluated with a commercial pin-on-disc test machine
Fig. 9. Friction vs. time under dry conditions.
using a ball-on-flat contact configuration. Flat specimens of
25 mm × 25 mm × 10 mm dimensions were prepared from single
pass and multi-passes FSP as well as unprocessed 1080 steel mate-
rial as the baseline. Because of significant surface roughening by the
FSP, the flat specimens for both the baseline and processed material
were ground to the same surface finish with SiC abrasive paper to
a surface finish of about 0.3 ␮m Ra . The ball specimen is made of
12.7 mm diameter alumina ball. Because of the considerable higher
hardness of the alumina ball (20 GPa) compared to the 1080 steel
disc (3–7 GPa), most of the wear during testing is expected to occur
in the steel specimen. This will facilitate the assessment of the
impact of FSP on wear of 1080 steel.
During friction and wear testing, the ball is held stationary and
loaded against a rotating flat. The friction force (from which fric-
tion coefficient is calculated) was continuously monitored during
the entire duration of the test. All the tests were conducted at a
normal load of 10 N (which imposes a maximum Hertzian contact Fig. 10. Friction vs. time under lubricated conditions (10w30 Mobil).
pressure of 1.01 GPa, [16]), sliding speed of 0.05 m/s for a duration
of 1 h. Tests were conducted in ambient room air, lubricated with
a full formulated synthetic engine oil (Mobil 1 10W30) and lubri-
cated with a top-of-the-rail (TOR) lubricant. A summary of the test
conditions and parameters is shown in Table 1.
Fig. 9 shows the variation of friction coefficient with time during
friction and wear tests in the various environments. In dry slid-
ing contact, the steady state friction coefficient after wear-in for
both single pass and multi-pass FSP surfaces (0.68) is less than
the untreated baseline surface (0.85) as shown in Fig. 9. Although,
the run-in friction trend for the three surfaces during the first few
minutes of tests are similar to one another, the magnitude of the
run-in-friction coefficient for multi-pass FSP is the lowest (0.55)
while that of untreated surface is highest at about 0.7. The differ-
ence in the run-in friction may be due to differences in extent of
localized plastic deformation at real contact areas at the early stage
of test. Since the FSP surfaces are harder, less plastic deformation Fig. 11. Friction vs. time under lubricated conditions (TOR).

and hence lower friction is expected. After run-in the friction coef-
ficient is then influenced by the wear process and possibly by the For engine oil lubricated tests, the friction coefficient is nearly
third bodies consisting of wear debris and/or oxides present at the constant for all the three surfaces as shown in Fig. 10 with the fric-
contact interface. tion in multi-pass FSP slightly lower than the other two. Similarly
the magnitude of about 0.11 is typical for boundary lubrication

Table 1
Ball-on-flat friction and wear test parameters.

Variable Value

Flat material 1080 steel (baseline; single pass FSP; multi-pass FSP)
Ball material 0.5 in. alumina ball
Load 10 N
Duration 1h
Sliding speed 0.05 m/s
Sliding distance 200 m
Environment Dry (open air) with relative humidity = 29–37%
Lubricated with Mobil 1 10W30 oil
Lubricated with top of rail (TOR) lubricant
Fig. 12. 3D profile of the wear track.
354 S.H. Aldajah et al. / Wear 267 (2009) 350–355

Fig. 13. Optical micrograph of wear track on the1080 steel disc after wear testing: (a) untreated disc; (b) multi-pass FSP disc.

Table 2 than just a phase transformation or formation of martensite on the


Wear rat of flat samples lubricated with TOR.
surface layer. It is known that the lower wear observed with FSP
Surface treatment Wear rate of flat (mm3 /N m) compared to laser glazed 1080 steel could be the combined effects
Dry Engine oil lubricant TOR lubricant of finer grain size and the formation of martensite in FSP as opposed
to only martensite formation in laser glazing.
Reference 1080 steel 2.3 × 10−4 1.31 × 10−6 2.04 × 10−6
FSP single pass 3.0 × 10−5 NMW 4.57 × 10−7
Although preliminarily, the results of the present study identi-
FSP multi-pass 3.1 × 10−5 NMW 4.82 × 10−7 fies a new avenue to significantly improve the friction and wear
behavior of high carbon steel and perhaps all steel in general. There
may be opportunities to further optimize the FSP for friction and
regime [17]. In the test with top-of-the rail (TOR) lubricant, the wear control.
friction behavior for the three disc surfaces is also similar to one It shall be noted also that the formation of martensite surface
another (Fig. 11). In all cases the test started with a friction coeffi- layer, by whatever means, is going to be accompanied by residual
cient of about 0.11 (boundary regime friction), but decreased in an stresses. This could have significant impact on some tribological
exponential manner for the duration of the test with a final value failures especially contact fatigue and other processes that may
of about 0.06. This frictional behavior is indicative of continuous involve fracture.
changes in the surface topography and/or surface chemistry as the
test progresses. 4. Conclusions
The summary of the wear rates on the flat specimens at the
conclusion of each test is shown in Table 2. Wear volume on the Friction stir processing when applied to 1080 carbon steel
disc was measured by optical profilometry. Fig. 12 shows a 3D seg- resulted in significant changes in the near surface microstructure
ment of the wear track on the disc. Volume of material removed by from pearlite to martensite up to a depth of 5 mm. This microstuc-
wear is calculated by the system software. Total wear of the entire tural change resulted in increased hardness of 1080 steel from 3 GPa
track is calculated from segment volume. Wear rate is determined to about 7 GPa as expected since martensite is much harder than
by normalizing wear volume with normal load and total sliding pearlite. The improvement in surface mechanical properties was
distance, i.e. wear rate = wear volume (mm3 )/(load (N) × distance observed to translate to improvement in friction and wear behavior
(m)). The wear rate under dry contacts is about two orders of mag- as measured by a unidirectional sliding ball-on-flat contact con-
nitude higher than the wear rates under both TOR and engine oil figuration. Under dry contact, FSP reduced friction coefficient by
lubrication. This is expected as lubricants usually provide wear pro- about 25% and wear rate by an order of magnitude. Under oil lubri-
tection especially when formulated with anti-wear additives. Also, cated conditions, FSP had only a marginal impact on friction since
since the friction coefficients in the lubricated contacts are much it depends largely on the lubricant. Significant reduction in wear
lower than the dry contact, the reduced shear stresses at the con- rate was observed in FSP surfaces even under oil lubrication.
tact interface are expected to reduce the extent of damage and
wear and perhaps change the wear mechanism. The wear rate in Acknowledgement
the FSP process is about an order of magnitude lower than the
untreated surface under both dry and lubricated conditions. This This work was supported by the Department of Energy, Office of
can be attributed to differences in the flat surface hardness and Vehicle Technologies, under contract no. DE-AC02-06CH11357.
perhaps change in wear mechanism. The differences in the extent
of plastic deformation and wear damage as a result of hardness dif- References
ference is illustrated in Fig. 13, which is an optical micrograph of
wear track of untreated surface and multi-pass FSP. There is evi- [1] R.S. Mishra, M.W. Mahoney, S.X. McFadden, N.A. Mara, A.K. Mukherjee, Scripta
dence of more severe plastic deformation and material removal in Mater. 42 (2000) 163.
[2] R.S. Mishra, M.W. Mahoney, Mater. Sci. Forum 357-3 (2001) 507.
the untreated surface (Fig. 13a) compared to FSP surface (Fig. 13b).
[3] W.M. Thomas, E.D. Nicholas, J.C. Needham, M.G. Murch, P. Templesmith, C.J.
More work is needed and planned to develop a better under- Dawes, GB Patent Application No. 9125978.8, December 1991.
standing of the role of FSP on tribological performance of 1080 steel. [4] S. Mironov, Y.S. Sato, H. Kokawa, Acta Mater. 56 (2008) 2602.
[5] Z.Y. Ma, R.S. Mishra, M.W. Mahoney, Acta Mater. 50 (2002) 4419.
In a previous study in which a martensite phase layer was formed
[6] Z.Y. Ma, R.S. Mishra, M.W. Mahoney, Scripta Mater. 50 (2004) 931.
on the surface by a laser glazing process, the frictional behavior [7] Z.Y. Ma, R.S. Mishra, Scripta Mater. 53 (2005) 75.
and magnitude are similar to what was obtained with FSP. How- [8] Z.Y. Ma, R.S. Mishra, M.W. Mahohey, R. Grimes, Metall. Mater. Trans. A36 (2005)
ever, the wear rate in the material processed by FSP is about an 1147.
[9] R.S. Mishra, Z.Y. Ma, I. Charit, Mater. Sci. Eng. A341 (2003) 307.
order of magnitude less than the one for laser glazing [18]. This [10] P.B. Berbon, W.H. Bingel, R.S. Mishra, C.C. Bampton, M.W. Mahoney, Scripta
will suggest that the impact of FSP, particularly on wear, is more Mater. 44 (2001) 61.
S.H. Aldajah et al. / Wear 267 (2009) 350–355 355

[11] J.E. Spowart, Z.Y. Ma, R.S. Mishra, in: K.V. Jata, M.W. Mahoney, R.S. Mishra, S.L. [14] M.L. Santella, T. Engstrom, D. Storjohann, T.Y. Pan, Scripta Mater. 53 (2005) 201.
Semiatin, T. Lienert (Eds.), Friction Stir Welding and Processing II, TMS, Warren- [15] Y.S. Sato, H. Yamanoi, H. Kokawa, T. Furuhara, Scripta Mater. 57 (2007) 557.
dale, PA, 2003, pp. 243–252. [16] K.L. Johnson, Contact Mechanics, Cambridge University Press, 1987.
[12] Z.Y. Ma, S.R. Sharma, R.S. Mishra, M.W. Mahoney, Mater. Sci. Forum 426–432 [17] K.C. Ludema, Friction, Wear, Lubrication A Textbook in Tribology, CRC Press,
(2003) 2891. 1996.
[13] S.R. Sharma, Z.Y. Ma, R.S. Mishra, Scripta Mater. 51 (2004) 237. [18] S. Aldajah, O.O. Ajayi, G. Fenske, Trans. ASME, J. Tribol. 127 (July (3)) (2005) 596.

You might also like