Elsevier

Wear

Volumes 448–449, 15 May 2020, 203169
Wear

Microstructures and intrinsic lubricity of in situ Ti3SiC2–TiSi2–TiC MAX phase composite fabricated by reactive spark plasma sintering (SPS)

https://doi.org/10.1016/j.wear.2019.203169Get rights and content

Highlights

  • Spark plasma sintering (SPS) was used to synthesize fully dense MAX phase composite Ti3SiC2–TiSi2–TiC from elemental powders.

  • Dry-sliding friction and wear test was carried out using a pin-on-disc configuration and room temperature.

  • The MAX phases exhibited intrinsic self-lubricity owing to the evolution TiC, TiCXOY, and graphitic carbon at the sliding surface.

  • TiC acts as a load bearing element by decentralizing the sliding load via plastic deformation, as well as introducing pinning effect of the Ti3SiC2 grains to inhibit deformation and grain pull-outs.

Abstract

MAX phase composite Ti3SiC2–TiSi2–TiC based on the Tin+1SiCn system was synthesized by spark plasma sintering (SPS) under vacuum sintering conditions. The microstructural evolution upon synthesis and Vickers indentation contact damage were characterized using scanning electron microscopy (SEM) and optical microscopy (OM). Tribological behaviour of the SPSed MAX phase composite was investigated under dry sliding ambient conditions for evidence of intrinsic lubricity as well as to understand the influence of second phase TiC particles on the wear behaviour of this composite system. Further, the underlying wear mechanisms was elucidated via detailed analyses of the worn surfaces using Raman spectroscopy, SEM-EDS and transmission electron microscopy (TEM). Exhaustive analyses of the worn surface revealed evidence of solid lubrication. Transition in friction and wear is attributed to change in wear mechanism from tribo-oxidative to deformation-induced wear due to the disruption of the tribofilm architecture.

Introduction

Early transition-metal ternary metalloceramics composed of hexagonal nanolaminated layered structure with a chemistry M(n+1)AX(n) configuration have attracted a lot of attention [1,2]. Since first discovered in the late 1960s, MAX phases have been recently further explored to investigate their synthesis and structure-property relation owing to their unusual set of metal-like (machinability, stiffness, electrical and thermal conductivities) and ceramic-like (damage tolerance, thermal stability, oxidation resistance) properties [3,4]. The MAX phases are so-called because of their general formula where M are mainly group-4, group-5, and group-6 transition metals (mainly Ti, Zr, Hf, V, Nb, Ta, Cr, and Mo), while A is mainly an A-group element from groups 9 (Ir), 10 (Pd), 11 (Cu, Au), 12 (Cd, Zn), 13 (Al, Ga, In, Tl), 14 (Si, Ge, Sn, Pb), 15 (P, As, Sb, Bi), X is either C or N and n = 1–3 and possibly higher [1,5,6]. Fig. 1 shows some of the elements in the periodic table that forms the Mn+1AXn phases.

These ternary phases (80 +) crystallize in a hexagonal structure (P63/mmc symmetry) with two formula units per unit cell, where Mn+1Xn layers are interleaved with pure A-group atoms thus resulting in a characteristic (Mn+1Xn)A(Mn+1Xn)A(Mn+1Xn) crystal structure [7]. The nature of their characteristic layered structure composed of stacking of n “ceramic” interposed with a “metallic” layer [8], coupled with the mixed covalent-metallic nature of the M–X bonds which are exceptionally strong and the relatively weak M–A bonds, endows the MAX phases with their signature mechanical, chemical, and electrical properties [4,9,10]. They represent the only class of ceramic-like material that deforms plastically via the nucleation and slip of basal dislocations (BDs) [11,12], incorporating a series of energy absorbing micro-scale events such as buckling of individual grains, diffuse micro-cracking, delamination of individual grains, kink and shear band formation followed by eventual grain push-outs and pull-outs [4,10,[13], [14], [15], [16], [17], [18], [19], [20]].

Ti3SiC2, a 312 compound, is the most studied representative member of the MAX phase family. It possesses unique metalloceramic properties such as low hardness, low density, high modulus, excellent thermal and electrical conductivity, high fracture toughness, damage tolerance and easy machinability [4,[21], [22], [23]]. Its hexagonal layered crystal structure similar to graphite and MoS2 [24] suggests it might be an excellent solid lubricant material with a low friction and wear properties suitable in a range of high-temperature structural engineering applications [25,26]. This is supported by the fact that Ti3SiC2 felt lubricious during machining as reported by Barsoum et al. [21]. Numerous research have since been conducted on the tribological behaviour of MAX phases for evidence of lubricity [27]. However, some researchers have reported that albeit Ti3SiC2 possessing layered hexagonal crystal structure similar to graphite, it is not intrinsically self-lubricating [28]. This was attributed to a three-body abrasive wear that stems from the fracture and pull-out of the Ti3SiC2 grains — representing the dominant wear mechanism at room temperature [27,29,30]. The ease of grain fracture and pull-outs has been linked to the weak grain boundary force of the Ti3SiC2 grains [31]. In addition, some other authors have highlighted the low hardness and oxidation resistance as the main factors deteriorating the friction and wear properties of monolithic Ti3SiC2 [32]. These observations have led to new studies focussing essentially on the incorporation of a second phase hard material in the soft matrix of Ti3SiC2 as an effective way to mitigate these weaknesses [32,33]. Some possible reinforcing materials are TiC and TiB2 as they possess high hardness, excellent oxidation resistance and close coefficient of thermal expansion (CTE) with Ti3SiC2 [32,34].

Nevertheless, little is known on the exact wear mechanism(s) as well as the intrinsic self-lubricating behaviour of monolithic Ti3SiC2 and Ti3SiC2-based material due to the lack of detailed investigation undertaken on the worn surface irrespective of the varying testing conditions reported to date. The scope of this work is to determine comprehensively the wear mechanism sequence of this solid and its associated composites during dry sliding friction at ambient conditions, in order to establish the existence of intrinsic self-lubricating behaviour as speculated [30,35] and to further elucidate the wear mechanism.

Section snippets

Powder preparation

Commercially available titanium powder (100 mesh, 99.7% purity, Aldrich), silicon powder (200 mesh, 99% purity, Acros organics) and graphite powder were used as starting elemental powders. 5.53 g of titanium powder and 0.98 g of graphite powder were dry-milled in a SPEX 8000 mill continuously for 2 h and subsequently mixed with 1.08 g of silicon powder according to a 3:1:2 stoichiometry.

Consolidation by spark plasma sintering (SPS)

The stoichiometric powder mixture was poured into an electrically and thermally conductive cylindrical

SPS sintering cycle

Fig. 3 shows the variation of temperature, applied force (pressure), punch displacement (piston movement) and sintering speed during the SPS cycle. As shown, the punch displacement has been divided into distinctive segments (I–IV) to represent the respective sintering stages and corresponding sintering events. In segment I, particle rearrangement initiated by the applied force resulted in positive punch displacement due to powder compression. A corresponding increase in sintering speed (~

Friction and wear

The evolution of friction coefficient as a function of time for the contact condition [0.5 N/50 rpm/60 min] and [0.5 N/100 rpm/30 min] are shown in Fig. 12. The noticeable features of these plots are the friction transition(s) and mild stick-slip phenomenon. The friction plots have been divided into three regimes (I, II, and III). In regime I, the friction was initially very low with no visible wear scar. This was then followed by a transition in friction to a high friction regime II, where the

Wear mechanisms

Following microstructural and chemical analyses before and after the wear tests, the wear mechanisms (Fig. 24) of this MAX phase composite system for the test conditions are designated oxidative–deformation–reoxidation as explained thus:

Conclusions

Dense polycrystalline MAX phase composite Ti3SiC2–TiSi2–TiC was successfully synthesized by spark plasma sintering via the elemental powder route. The following conclusions can be drawn upon exhaustive characterization of the deformation microstructure and tribological behaviour of this MAX phase composite system:

  • 1.

    Deformation microstructure revealed evidence of room temperature plasticity, toughening and anisotropy in mechanical response.

  • 2.

    Evidence of intrinsic solid lubrication was observed due

Declaration of competing interest

We have no conflict of interest to disclose.

Acknowledgement

We wish to acknowledge the Henry Royce Institute for Advanced Materials for equipment access at Royce@sheffield.

References (58)

  • S. Gupta et al.

    On the tribology of the MAX phases and their composites during dry sliding: a review

    Wear

    (2011)
  • W. Dang

    Influence of Cu on the mechanical and tribological properties of Ti3SiC2

    Ceram. Int.

    (2016)
  • T. El-Raghy et al.

    Effect of grain size on friction and wear behavior of Ti3SiC2

    Wear

    (2000)
  • J. Yang

    Friction and wear properties of in situ (TiB2+TiC)/Ti3SiC2 composites

    Wear

    (2011)
  • W. Dang

    The tribological properties of Ti3SiC2/Cu/Al/SiC composite at elevated temperatures

    Tribol. Int.

    (2016)
  • S. Konoplyuk

    Synthesis of Ti3SiC2/TiC composites from TiH2/SiC/TiC powders

    Mater. Lett.

    (2005)
  • S. Gupta

    Tribological behavior of select MAX phases against Al2O3 at elevated temperatures

    Wear

    (2008)
  • I. Kero et al.

    Phase reactions associated with the formation of Ti3SiC2 from TiC/Si powders

    Ceram. Int.

    (2011)
  • H. Hashimoto

    Synthesis of Ti3SiC2 from powder blend of Ti, Si and TiC

    J. Alloy. Comp.

    (2006)
  • W.K. Pang

    15 - phase and thermal stability in Ti3SiC2 and Ti3SiC2/TiC/TiSi2 systems

  • P.J. Counihan et al.

    Influence of dynamic densification on nanostructure formation in Ti5Si3 intermetallic alloy and its bulk properties

    Mater. Sci. Eng. A

    (1999)
  • J. Lis

    Ti3SiC-based materials prepared by HIP-SHS techniques

    Mater. Lett.

    (1995)
  • C. Magnus et al.

    Influence of sintering environment on the spark plasma sintering of Maxthal 312 (nominally-Ti3SiC2) and the role of powder particle size on densification

    J. Alloy. Comp.

    (2019)
  • C. Magnus et al.

    Spark plasma sintering (SPS) synthesis and tribological behaviour of MAX phase composite of the family Tin+1SiCn (n = 2)

    Wear

    (2019)
  • W.G. Sloof

    Repeated crack healing in MAX-phase ceramics revealed by 4D in situ synchrotron X-ray tomographic microscopy

    Sci. Rep.

    (2016)
  • A. Joulain et al.

    Revisiting the defect structure of MAX phases: the case of Ti4AlN3

    Philos. Mag.

    (2008)
  • M.W. Barsoum et al.

    Elastic and mechanical properties of the MAX phases

  • H. Fashandi

    Synthesis of Ti3AuC2, Ti3Au2C2 and Ti3IrC2 by noble metal substitution reaction in Ti3SiC2 for high-temperature-stable Ohmic contacts to SiC

    Nat. Mater.

    (2017)
  • P. Eklund et al.

    Layered ternary Mn+1AXnphases and their 2D derivative MXene: an overview from a thin-film perspective

    J. Phys. D Appl. Phys.

    (2017)
  • Cited by (26)

    View all citing articles on Scopus
    View full text