thermal behavior of the molten pool, microstructural evolution, … · 2020. 8. 18. · particle...

14
Contents lists available at ScienceDirect Journal of Materials Processing Tech. journal homepage: www.elsevier.com/locate/jmatprotec Thermal behavior of the molten pool, microstructural evolution, and tribological performance during selective laser melting of TiC/316L stainless steel nanocomposites: Experimental and simulation methods Bandar AlMangour a, , Dariusz Grzesiak b , Jinquan Cheng c , Yavuz Ertas d a School of Engineering and Applied Sciences, Harvard University, Cambridge, MA, 02138, USA b Department of Mechanical Engineering and Mechatronics, West Pomeranian University of Technology, Szczecin, Aleja Piastów 17, Poland c Composite Solutions and Digital Manufacturing LLC, Chandler, AZ, 85248, USA d Department of Bioengineering, University of California, Los Angeles, CA, 90095, USA ARTICLE INFO Keywords: Selective laser melting Dispersion Particle size Simulation Marangoni convection Wear ABSTRACT Bulk-form nanocomposites of TiC-reinforced 316L stainless steel matrix were fabricated by selective laser melting (SLM), an emerging powder-bed additive manufacturing technique which allows the direct fabrication of usable end-products, using various volumetric laser energy densities (η). The microstructural features of the distribution and sizes of TiC nanoparticles, as well as the grain sizes and tribological performances of the SLM- processed nanocomposite parts, were sensitive to the applied η. Increasing η enhanced the dispersion state of nanoscale TiC owing to intensied Marangoni ow and the corresponding capillary force, which prevented TiC aggregation and promoted a uniform dispersion of reinforcements in the solidied matrix. However, with in- creasing η, the TiC particle size also increased, and some nanoparticles lost their initial nanostructure because of signicant thermal accumulation within the molten pool. Increasing η also caused increases in the grain sizes of the fabricated nanocomposite because of the decreasing cooling rate. A simulation model was developed to enhance understanding of the manufacturability of these new materials, as well as to predict the temperature evolution and thermal behaviors of the molten pool under various η. The simulation modeled the eects of various η values on the temperature distribution evolutions and the corresponding eects of Marangoni con- vection during the SLM process. The temperature distribution was signicantly inuenced by the applied η; the maximum temperature gradient within the molten pool was increased signicantly with increased η. The si- mulation results validated the experimental results and the underlying physical mechanism of the molten pool. The microhardness of the SLM nanocomposite decreased sharply with increased grain size due to the lower cooling rate, but increased with further increases in η because of the enhanced densication degree. Nanocomposites processed under the optimum condition of η = 200 J/mm 3 showed the lowest wear rates ac- companied by the formation of adherent and strain-hardened tribolayers on the worn surfaces of the nano- composites, suggesting improved tribological performance. 1. Introduction Many material-dependent sectors, including aircraft, spacecraft, biomedical devices, and electronics, require materials with extra- ordinary combinations of features not found in traditional pure cera- mics or metal alloys. Therefore, various composite materials have re- cently been developed with improved material properties. Metal alloys can be reinforced by homogenous dispersions of second-stage hard and ne ceramic particulates; when nanoscale reinforcement particles are present, the resulting materials are called nanocomposites. The strengthening mechanism involves both interparticle interactions and dislocations within the matrix. Pagounis and Lindroos (1998) demon- strated that stainless steel-matrix composites (SMCs) strengthened with stier and harder ceramic particles exhibited promising features, in- cluding high specic strength, specic stiness, and wear resistance. Numerous manufacturing methods have been reported for produ- cing nanoparticle-reinforced metal matrix composites (MMCs). Sheikhzadeh and Sanjabi (2012) successfully synthesized MMCs by mechanical alloying, while Saheb et al. (2012) and Chen et al. (2013) eectively used powder metallurgy and stir-casting, respectively. However, Viswanathan et al. (2006) and Moya et al. (2007) indicated that obtaining a homogenous distribution of nanoscale reinforcing https://doi.org/10.1016/j.jmatprotec.2018.01.028 Received 23 June 2017; Received in revised form 20 November 2017; Accepted 21 January 2018 Corresponding author. E-mail addresses: [email protected], [email protected] (B. AlMangour). Journal of Materials Processing Tech. 257 (2018) 288–301 Available online 02 February 2018 0924-0136/ © 2018 Elsevier B.V. All rights reserved. T

Upload: others

Post on 03-Feb-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Contents lists available at ScienceDirect

    Journal of Materials Processing Tech.

    journal homepage: www.elsevier.com/locate/jmatprotec

    Thermal behavior of the molten pool, microstructural evolution, andtribological performance during selective laser melting of TiC/316L stainlesssteel nanocomposites: Experimental and simulation methods

    Bandar AlMangoura,⁎, Dariusz Grzesiakb, Jinquan Chengc, Yavuz Ertasd

    a School of Engineering and Applied Sciences, Harvard University, Cambridge, MA, 02138, USAbDepartment of Mechanical Engineering and Mechatronics, West Pomeranian University of Technology, Szczecin, Aleja Piastów 17, Polandc Composite Solutions and Digital Manufacturing LLC, Chandler, AZ, 85248, USAd Department of Bioengineering, University of California, Los Angeles, CA, 90095, USA

    A R T I C L E I N F O

    Keywords:Selective laser meltingDispersionParticle sizeSimulationMarangoni convectionWear

    A B S T R A C T

    Bulk-form nanocomposites of TiC-reinforced 316L stainless steel matrix were fabricated by selective lasermelting (SLM), an emerging powder-bed additive manufacturing technique which allows the direct fabricationof usable end-products, using various volumetric laser energy densities (η). The microstructural features of thedistribution and sizes of TiC nanoparticles, as well as the grain sizes and tribological performances of the SLM-processed nanocomposite parts, were sensitive to the applied η. Increasing η enhanced the dispersion state ofnanoscale TiC owing to intensified Marangoni flow and the corresponding capillary force, which prevented TiCaggregation and promoted a uniform dispersion of reinforcements in the solidified matrix. However, with in-creasing η, the TiC particle size also increased, and some nanoparticles lost their initial nanostructure because ofsignificant thermal accumulation within the molten pool. Increasing η also caused increases in the grain sizes ofthe fabricated nanocomposite because of the decreasing cooling rate. A simulation model was developed toenhance understanding of the manufacturability of these new materials, as well as to predict the temperatureevolution and thermal behaviors of the molten pool under various η. The simulation modeled the effects ofvarious η values on the temperature distribution evolutions and the corresponding effects of Marangoni con-vection during the SLM process. The temperature distribution was significantly influenced by the applied η; themaximum temperature gradient within the molten pool was increased significantly with increased η. The si-mulation results validated the experimental results and the underlying physical mechanism of the molten pool.The microhardness of the SLM nanocomposite decreased sharply with increased grain size due to the lowercooling rate, but increased with further increases in η because of the enhanced densification degree.Nanocomposites processed under the optimum condition of η=200 J/mm3 showed the lowest wear rates ac-companied by the formation of adherent and strain-hardened tribolayers on the worn surfaces of the nano-composites, suggesting improved tribological performance.

    1. Introduction

    Many material-dependent sectors, including aircraft, spacecraft,biomedical devices, and electronics, require materials with extra-ordinary combinations of features not found in traditional pure cera-mics or metal alloys. Therefore, various composite materials have re-cently been developed with improved material properties. Metal alloyscan be reinforced by homogenous dispersions of second-stage hard andfine ceramic particulates; when nanoscale reinforcement particles arepresent, the resulting materials are called nanocomposites. Thestrengthening mechanism involves both interparticle interactions and

    dislocations within the matrix. Pagounis and Lindroos (1998) demon-strated that stainless steel-matrix composites (SMCs) strengthened withstiffer and harder ceramic particles exhibited promising features, in-cluding high specific strength, specific stiffness, and wear resistance.

    Numerous manufacturing methods have been reported for produ-cing nanoparticle-reinforced metal matrix composites (MMCs).Sheikhzadeh and Sanjabi (2012) successfully synthesized MMCs bymechanical alloying, while Saheb et al. (2012) and Chen et al. (2013)effectively used powder metallurgy and stir-casting, respectively.However, Viswanathan et al. (2006) and Moya et al. (2007) indicatedthat obtaining a homogenous distribution of nanoscale reinforcing

    https://doi.org/10.1016/j.jmatprotec.2018.01.028Received 23 June 2017; Received in revised form 20 November 2017; Accepted 21 January 2018

    ⁎ Corresponding author.E-mail addresses: [email protected], [email protected] (B. AlMangour).

    Journal of Materials Processing Tech. 257 (2018) 288–301

    Available online 02 February 20180924-0136/ © 2018 Elsevier B.V. All rights reserved.

    T

    http://www.sciencedirect.com/science/journal/09240136https://www.elsevier.com/locate/jmatprotechttps://doi.org/10.1016/j.jmatprotec.2018.01.028https://doi.org/10.1016/j.jmatprotec.2018.01.028mailto:[email protected]:[email protected]://doi.org/10.1016/j.jmatprotec.2018.01.028http://crossmark.crossref.org/dialog/?doi=10.1016/j.jmatprotec.2018.01.028&domain=pdf

  • particles in the matrix can be very challenging because extensive vander Waals interactions, coupled with the large surface areas of nano-particle reinforcements, can cause considerable microstructural in-homogeneity. Moreover, Kim et al. (2004) and Lee and Kang (2006)reported difficulties in preserving the nanostructures of loose powdersduring the consolidation of composites because of extreme grain growthduring high-temperature processes such as casting, sintering, and hotisostatic pressing. In addition, fabricating customized functional com-plex parts for high-tech applications by these conventional processingmethods is quite challenging. Therefore, it is important to develop anovel, non-traditional nanocomposite fabrication method that canovercome these complex processing issues.

    Selective laser melting (SLM) is an emerging additive manu-facturing process that provides substantial flexibility in the productionof parts with complex shapes that are suitable for various end uses.Parts can be synthesized directly from loose powders that cannot beprocessed readily by conventional processing methods of casting andsintering, as discussed by Gu et al. (2012b). SLM allows the fabricationof high-performance MMC parts with fine grain sizes because of theextremely high solidification rates stimulated by non-equilibrium laserscanning, as reported by Das et al. (2010) and Pei et al. (2005). How-ever, the SLM process involves multiple modes of heat, mass, and mo-mentum transfer, and is thus, as noted by Kruth et al. (2007), a complexmetallurgical process with strong Marangoni effects induced by thermalcapillary forces and very high cooling rates. The arrangement of na-noparticles and dispersion state are considerably influenced by fluidflow. However, existing knowledge mostly relies on experimental re-sults, which remain lacking; moreover, little theoretical work has beenperformed. Obtaining the transient temperature variation and under-standing the mechanisms underlying nanoparticle movements are dif-ficult in experimental approaches. Therefore, numerical simulationmethods are necessary to analyze and understand the influence of thelaser energy density on the thermal behavior and temperature dis-tribution evolution, as well as the fluid flow driven by surface tensiongradients and Marangoni convection, in the molten pool during SLM.

    Few studies have reported the possibility of processing Fe-matrixand 316L stainless steel-matrix nanocomposites by SLM. Song et al.(2014a) studied the SLM behavior of SiC/Fe bulk nanocompositesformed directly from a mixed powder of micro-sized Fe and SiC, re-porting that the addition of SiC to the Fe matrix caused significant in-creases in the local melt instability and viscosity, as well as enhancedtensile properties. Hao et al. (2009) and Wei et al. (2015) processedhydroxyapatite/stainless steel composites by SLM for medical applica-tions and reported the microstructural characteristics, element dis-tribution, and mechanical properties of the composites. The authors ofthe current study, AlMangour et al. (2016a; and 2016b), previouslyreported progressive improvements in the mechanical and tribologicalperformance of 316L nanocomposites with increasing TiC and TiB2volume contents. More recent publications by the authors, AlMangouret al. (2017 and 2018), analyzed the densification behavior, micro-structural evolution, and compression properties of TiC/316L nano-composites processed using various SLM scanning methods and energydensities. Nanoscale TiC was selected as the reinforcement based on its

    excellent thermodynamic stability and wettability in molten steel, aswell as its high hardness and elastic modulus as reported by Abenojaret al. (2002). In this study, the thermal behaviors and physical me-chanisms of the molten pool are investigated, as are the evolutions ofthe TiC dispersion state and particle size and the TiC/316L nano-composite grain size during SLM at various applied energy densities. Asimulation model is constructed to predict the thermal behaviors of theSLM processing. The resulting variations in the microstructures andcorresponding tribological performances of the nanocomposites arethen analyzed. A process–microstructure–properties correlation wasclarified to facilitate the effective production of SMC components.

    2. Experimental procedures and simulation setup

    2.1. Powder preparation and SLM processing

    In this study, 316L spherical gas-atomized powder with an averageparticle diameter of 45 μm (Fig. 1a) and nearly spherical TiC nano-powders with an average particle size of 50 nm were used as startingmaterials (Fig. 1b). Mechanical mixing of the TiC and 316L powderscomprising 15 vol.% TiC was performed in a Fritsch Pulverisette 4Vario-Planetary mill, according to the milling conditions discussed inAlMangour et al. (2016a). The milled nanocomposite powders wereseverely deformed, with TiC nanoparticles embedded within the 316Lmatrix powders (Fig. 1c).

    A commercial SLM machine (MCP HEK REALIZER II) was used,consisting of a Yb-fiber laser, automatic powder layering system, inertAr gas protection system, and a computer system for process control.Before starting the SLM process, a steel substrate was placed on thebuilding platform and leveled; then the building chamber was sealedand filled with Ar gas to minimize oxidation (an O2 content of < 1%was used to process the feedstock powder). The main SLM processingparameters were kept constant with a laser power (P) of 100W and aspot size of 180 μm. An alternate-hatching scanning pattern (with 90°rotations between successive layers) was applied. The powder was ap-plied by spreading it over the building platform using a wiper. Thebuilding platform was lowered to a distance equal to the layer thick-ness, and the resulting space was filled with the powder. The SLMparameters of powder layer thickness (d) and scan line hatch spacing(h) were kept constant at 50 μm and 120 μm, respectively. To alter theprocessing conditions during processing, different scanning speeds (v)of 250, 133, 83, and 55.6 mm/s were set using the SLM control pro-gram. Thus, four varying volumetric laser energy densities of η =67,125, 200, and 300 J/mm3, defined by Gu et al. (2012a) as:

    = Pv hd

    η(1)

    were used to evaluate the laser energy input to the powder layer beingprocessed.

    2.2. Microstructural characterization

    Specimens for microstructural characterization were ground and

    Fig. 1. Microstructure of the starting materials: (a) 316L stainless steel; (b) TiC; and (c) ball-milled TiC/316L feedstock powder.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    289

  • polished in accordance with standard metallographic practices, andthen etched using Marble’s reagent for 10 s. The relative densities of theSLM-processed nanocomposites were estimated based on theArchimedes principle, and the average value of three readings was ta-bulated for each set of processing conditions. The microstructure of theSLM-processed nanocomposite samples was examined at their crosssection (i.e., parallel to the building direction) using a Nano 230scanning electron microscope (SEM). An FEI T12 transmission electronmicroscope (TEM) at 120 kV was used to examine the internal micro-structures of the fabricated nanocomposites. TEM samples were pre-pared via focused-ion beam (FIB) milling. The grain orientations of thespecimens were characterized using electron backscatter diffraction(EBSD) at 15 kV using an FEI Quanta 600F SEM with a step size of40 nm at low magnification to collect broader statistical data. Thespecimens were polished using a diamond suspension and colloidal si-lica (100-μm step sizes to 0.01 μm) for approximately 8 h prior to theEBSD measurements.

    2.3. Modeling approach and numerical simulation

    2.3.1. Physical model description and governing equationsA layerwise additive manufacturing model using the relevant

    Layerwise Additive Manufacturing Predictions and Simulations(LAMPS) software was used to analyze the full-scale temperature fieldresponse and quantitatively explain the experimental results. The full-scale layerwise AM model is based on the governing behavior of con-ductive heat transfer in a laminated structure and the track-by-trackand layerwise deposition of AM, as schematically shown in Fig. 2. Here,the heat transfer-governing equation for each layer in the laminatedplate can be described by

    ⎜ ⎟∂

    ∂⎛⎝

    ∂∂

    ⎞⎠

    + ∂∂

    ⎛⎝

    ∂∂

    ⎞⎠

    + ∂∂

    ⎛⎝

    ∂∂

    ⎞⎠

    + = ∂∂

    xK T x y z t

    x yK T x y z t

    y zK T x y z t

    z

    q ρC T x y z tt

    ( , , , ) ( , , , ) ( , , , )

    ( , , , )int (2a)

    or

    ∇∙ ∇ =K T f( ) on V (2b)

    with the relevant boundary conditions as follows:

    =T T on S11 (2c)

    ∙∇ =K T q on S2s (2d)

    − ∼∙∇ = − + −∞ ∞Kn T h T T r T T( ) ( ) on S34 4 (2e)

    where T(x,y,z,t) is the temperature of the plate. K, ρ, and C are thelocation-dependent thermal conductivity, density, and specific heatcapacity of the plate, respectively. qint is the internal heat source and qs

    is the applied surface heat flux. Further details on the fundamentals ofthe theoretical model are given in the LAMPS user’s manual (CSDM,2017). Since the LAMPS software is a full-scale transient simulationtool, it provides detailed information on the effects of the processingparameters on the temperature field from track-to-track and layer-to-layer.

    2.3.2. Physical propertiesThe geometries of the full substrates of 6061 Al alloy and 316L

    stainless steel are shown in Fig. 3 and were used to predict and analyzethe real deposition. The thermophysical properties of the substrate arelisted in Table 1. The effective thermophysical properties of the TiCnanoparticle-reinforced steel were obtained based on the sub-component volume fraction. The effective thermophysical properties ofthe powder bed were estimated in terms of the powder-bed densifica-tion and volume fractions of the TiC and steel particles.

    2.3.3. Simulation parametersBecause the numerical analysis is intended to clarify the effect of

    varying η, only the strip deposition with 8mm length and 0.00102mmwidth under the specified processing condition was modeled to reducethe computational work and save time. Considering that the powderparticles are surrounded by or embedded in TiC particles, the absorp-tivity of η was set as 0.82. The processing parameters are listed inTable 2 for LAMPS simulation (Yuan and Gu, 2015). Since the LAMPSsoftware considers the effect of the machine movement accelerationand deceleration, here both the acceleration and deceleration were setas 5m/s2.

    2.4. Hardness and wear testing

    A Leco LM800AT micro-hardness tester was used to measure theVickers hardness values at a load of 0.2 kg and indentation period of10 s. Each sample was subjected to 15 indentations, and the averageVickers hardness was considered. Dry rotary wear tests were performedaccording to the ASTM G99 standard using a T50 ball-on-disk trib-ometer in air at room temperature. The specimen surfaces were groundand polished before wear testing to ensure identical surface roughnessvalues. The counterface material was a 3-mm-diameter 52,100 chromesteel ball, applied with a test load of 3 N. The friction unit was rotatedat 840 rpm for 20min with a fixed rotation radius of 2mm. A ST 400light profilometer with a vertical resolution of 8 nm, accuracy of 80 nm,and lateral resolution of 2 μm was used to measure the volume of ma-terials lost (V). The wear rate (w) of the samples was then determinedby:

    = − −w VF L

    (mm N m )3 1 1 (3)

    Fig. 2. Schematic of the laminated layerwise additive manufacturing model.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    290

  • where F is the contact load and L is the total sliding distance.

    3. Results and discussion

    3.1. Effects of applied η on the distribution state of TiC

    During the solidification of the molten pool in SLM, the laser beammoves, and the thermal energy dissipates quickly to the substrate or thepreviously solidified layer. Niu and Chang (1999) and Gu and Shen(2009) both attributed this to the higher thermal conductivity of thesolid relative to that of the surrounding powder, which creates tem-perature gradients within the molten pool. Simchi et al. (2001)

    suggested that gradients of temperature or chemical composition in themolten pool, which are strongly influenced by the applied η, maygenerate a surface tension gradient and corresponding Marangoniconvection, thus inducing a non-steady-state solidification process.Fig. 4 shows SEM images of the microstructures of the SLM-processednanocomposite parts, demonstrating that the dispersions of TiC nano-particles are significantly influenced by the applied η. At low η, the TiCnanoparticles tend to aggregate into clusters, forming microscale ag-glomerates of several nanoparticles formed in the matrix (Fig. 4a ande). As the applied η is increased to 125 J/mm3, the distribution of TiCnanoparticles is improved in uniformity and the TiC particles bondtogether as continuous ring-like structures (Fig. 4b). The high-magni-fication image, however, shows the presence of some inhomogeneousTiC agglomerates in the matrix (Fig. 4f). When η is increased further to200 J/mm3, the TiC reinforcements are uniformly and homogenouslydistributed throughout the matrix (Fig. 4c) with no visible aggregationof TiC nanoparticles within the SLM-processed part, even when ob-served under high magnification (Fig. 4g); however, the grain size in-creases markedly. At the highest η of 300 J/mm3, the TiC reinforce-ments remain uniformly distributed throughout the matrix (Fig. 4d).The microstructural development of the nanoscale TiC during SLM isclearly sensitive to the applied η. Using η≥ 125 J/mm3, the dispersionstate of the TiC reinforcements was enhanced (i.e., uniform andhomogenous distribution in the matrix) but at the expense of the for-mation of comparatively coarse grains. By decreasing the laser scanspeed, and thus the cooling rate, SLM-processed parts with increasinglycoarse microstructures were obtained (e.g., Fig. 4g). This is furtherdiscussed in Section 3.3.

    Simchi and Asgharzadeh (2004) and Gu et al. (2007) reported thatMarangoni convection within the molten pool, generated by hightemperature and high surface tension, induced a liquid capillary forcethat caused rearrangement of TiC reinforcements. Yuan and Gu (2015)found that Marangoni convection played an important role in in-tensifying the convective heat transfer and changing the molten poolgeometry. According to Arafune and Hirata (1999), the intensity ofMarangoni flow can be evaluated using the dimensionless Marangoninumber (Ma):

    Fig. 3. Substrate geometry and materials in the different zones for LAMPS simulation: (a) top view, (b) front view, (c) left view.

    Table 1Material properties of the substrate.

    Thermophysical Properties 6061 Al alloy 316L stainless steel

    Elastic modulus (GPa) 69 210Poisson’s ratio 0.3 0.28Shear modulus (GPa) 26 79Mass density (kg/m3) 2700 7800Tensile strength (MPa) 124.084 399.826Yield strength (MPa) 55.1485 220.594Thermal expansion coefficient (K−1) 2.40× 10−5 1.30×10−5

    Thermal conductivity (Wm−1 K−1) 170 43Specific heat (J kg−1 K−1) 1300 440

    Table 2The as-used relevant SLM processing parameters for LAMPS simulation.

    Parameters Value

    Laser power (W) 100Laser beam spot diameter (μm) 180Hatching spacing (μm) 120TiC absorptivity (×100%) 0.82Powder layer thickness (μm) 50Powder densification (×100%) 0.65Scan speed (mm/s) 55.6, 83, 133, 250Surface convection (Wm−1 K−1) 10

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    291

  • = σ Lv

    ΔMaμ k (4)

    where Δσ is the surface tension difference of Marangoni flow, L is thelength of the free surface, μ is the dynamic viscosity, and vk is the ki-nematic viscosity. The high temperature in the SLM process induced byhigh η causes a sharp decrease in μ, which intensifies Marangoni flowand turbulence within the molten pool, as well as the magnitude of theresulting capillary force. This, in turn, increases the rate of nanoscaleTiC rearrangement in the melt, preventing TiC aggregation and pro-moting uniform dispersion in the solidified matrix (Fig. 4c and d).

    Fig. 5 depicts simple mechanisms for the varied distribution statesof the TiC nanoparticles under decreased laser scan speeds and otherconstant processing parameters. In the laser-induced molten poolcomprising TiC and 316L particles developed by SLM, a sharp tem-perature gradient forms between the edge and center of the moltenpool. This produces a surface tension gradient and subsequent Mar-angoni flow, which, as reported by Simchi (2006), induces a liquid flowfor capillary forces. Whenever capillary forces act on the TiC nano-particles, a torque is generated around the particles, which Simchi(2008) attributed to particle center misalignment. The torque causesthe rotation of the TiC nanoparticles within the pool, enabling re-arrangement of the TiC distribution. Ma et al. (2013) reported that thesize of the temperature gradient determined the strength of the capil-lary forces. Moreover, an inadequate laser energy input appears toweaken both the SLM temperature gradient and the associated capillaryforce intensity; the TiC nanoparticles within the molten pool cannot besufficiently rearranged at low η. In this condition, the solidification rateof the liquid front in the molten pool is increased; therefore, there is aconsiderably shorter time for thermal Marangoni flow to rearrange theTiC nanoparticles. As a result, the TiC nanoparticle distribution is sig-nificantly non-uniform because of the formation of clusters of reinfor-cing particles at a comparatively low η of 67 J/mm3 (Fig. 5a).

    When η is increased by decreasing the scanning speed, the rate ofrearrangement of TiC reinforcing nanoparticles inside the molten poolis increased because of the intensification of Marangoni flow (Eq. 4).The torque force acts constantly on the TiC nanoparticles, which con-gregate around the core of the Marangoni flow pattern and form a ring-like structure (Fig. 5b and c). Meanwhile, repulsive forces tend to actbetween TiC reinforcing particles when an adequate quantity of 316Lmelt develops within the molten pool, as reported by Kruth et al.(2010). Although the converging flow pushes the TiC nanoparticlestoward the center, the collective impact of repulsion forces and sig-nificantly intensified Marangoni flow results in a much more well-de-veloped ring-like structure in the solidified matrix when a sufficiently

    high η (≥200 J/mm3) is applied (Fig. 5c). Under this condition, goodwetting of TiC by the liquid 316L alloy enhances the TiC dispersionstate. Nonetheless, when η is increased to the highest value of 300 J/mm3, the TiC reinforcement ring structure shows obvious coarsening(see Fig. 4d) because of considerable thermal accumulation within themolten pool, as well as subsequent grain growth and the large increasein the TiC particle size (as characterized in Sections 3.2 and 3.3).

    3.2. Effects of applied η on TiC particle size

    In addition to affecting the TiC dispersion state within the matrix,the applied η also affects the evolution of nanoscale TiC particle size bychanging the undercooling degree and the resulting solidification rate.Fig. 6a and b show TEM micrographs of the nanocomposites fabricatedunder low and high η and the effect of applied η on TiC particle size. Athigh scan speeds, η is low, which increases the solidification rate of theliquid front within the molten pool, causing large temperature gradientswithin the pool and high degrees of undercooling in the melt. Typically,the time available before solidification is insufficient for TiC graingrowth, leading to the favorable formation of refined TiC nanos-tructures (Fig. 6a). By examining the TEM image in Fig. 6a, we canconfirm that TiC nanoparticles approximately maintain their initialrefined size of 60–100 nm (Fig. 6c).

    Conversely, at the lowest scan speeds in this study, η is high; hence,the temperature in the molten pool increases, resulting in significantthermal accumulation within the molten pool. The longer molten timeand lower cooling rate significantly enhance the activity of grainswithin TiC nanoparticles in the molten pool, allowing grain growth forsome TiC particles. The statistical results show that some TiC particlesin the fabricated nanocomposite with η of 300 J/mm3 become verycoarse and lose their initial standard nanostructure, reaching a value of400 nm (Fig. 6b). In addition, the matrix of the composite shows anincreased dislocation density (Fig. 6d). Orowan bypassing of hard fineparticles by dislocations, which contributes to composite strengthening,is clearly revealed by the TEM image (Fig. 6b and d). Grain boundariesand particle/matrix interfaces are major sources of dislocations becauseof the large differences in the thermal expansion coefficients betweenTiC and 316L. At higher η, the particles are more homogenously dis-tributed, creating a larger number of particle-matrix interfaces. Underthese conditions, all particle clusters are broken, further increasing thenumber of such interfaces. At higher energy densities, more grainboundaries are available, increasing the total number of dislocationsources.

    Fig. 4. SEM images showing the evolution of nanoscale TiC reinforcement dispersion states in SLM-processed TiC/316L nanocomposites at various applied η: (a,e) η=67 J/mm3; (b,f)η=125 J/mm3; (c,g) η=200 J/mm3; (d,h) η=300 J/mm3.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    292

  • 3.3. Effects of applied η on grain size

    EBSD orientation maps of the nanocomposites using various η va-lues, taken from the top views of the fabricated cylinders, are shown inFig. 7. Analysis of the cross-sectional surfaces (i.e., parallel to thebuilding direction) was undertaken in our previous study (AlMangouret al., 2018). η clearly influences the thermal history and liquid flow byaltering the solidification rate, which in turn determines the final so-lidified microstructure.

    During the solidification process, the local growth velocity of thesolidification front, vs, can be calculated directly by measuring theangle (θ) between vs and the laser scan speed v using the followingequation proposed by Liu et al. (2004):

    =v v θcos( )s (5)

    The thermal undercooling ΔTt is expressed as follows defined bySchwarz et al. (1997):

    =TΔHc

    F PΔ tf

    pl Iv e

    (6)

    where Pe= v RD2s

    Tis the thermal Péclet number, ΔHf is the heat of fusion

    (J mol−1), clp is the specific heat of the liquid (J mol−1 K−1), FIv is theIvantsov function, R is the radius of curvature of the crystal tip (m), andDT is the thermal diffusivity (m2 s−1).

    The kinetic undercooling ΔTk within the pool, defined by Schwarzet al. (1997), is expressed as:

    =T vλ

    Δ k s (7)

    where λ= ΔH vk T2

    f

    B L

    0 is the interfacial kinetic coefficient, v0 is the speed ofsound (m s−1), kB is the Boltzmann constant, and TL is the liquidustemperature (K). At low v, the input η during line-by-line scanning ishigh, causing an elevated thermalization of the energy and a higherSLM temperature. With increased v, the solidification rate of the liquidfront within the pool is increased (Eq. 5), thereby enhancing both ΔTt(Eq. 6) and ΔTk (Eq. 7). Significantly enhanced temperature gradientswithin the pool further refine the grain size (Fig. 7a).

    The TiC nanoparticles have smaller grain sizes and higher meltingpoints than those of the 316L stainless steel matrix. Smallman and Ngan(2013) reported that these remained partially melted as impurityphases, and acted as nucleation sites for the molten stainless steel, fa-cilitating heterogeneous nucleation and consequently causing furthergrain refinements. The main driving force in the molten pools of theSLM process is, as reported by Niu and Chang (1999), convectionformed by a combination of surface tension gradients, viscous shearstress, and buoyancy forces. Furthermore, during SLM, laser scanning isperformed line-by-line, followed by layer-by-layer. These sequentialpasses influence the grain formation in multiple directions, as observedin the EBSD images, because of their corresponding thermal flows. Asthe temperature gradient decreases, the grain size increases.

    Microstructural development in these alloys is dependent on thethermal history experienced by different regions of a component. In thesolidification process, G is the temperature gradient in the liquid phaseand R is the growth rate of the interface, or the solidification rate. These

    Fig. 5. Schematic of TiC nanoparticle movement in the melt pool under various scanning speeds: (a) high scan speed; (b) moderate scan speed; and (c) low scan speed.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    293

  • Fig. 6. TEM images showing the TiC particle size of SLM-processed TiC/316L nanocomposites obtained at (a,c) η=67 J/mm3; and (b,d) η =300 J/mm3. Note the size of the nearlyspherical TiC particles distributed in the 316L matrix, as well as the increased dislocation density of the 316L austenitic matrix.

    Fig. 7. EBSD orientation maps of SLM-processed TiC/316L nanocomposites obtained at (a) η=67 J/mm3, (b) η=125 J/mm3, (c) η=200 J/mm3, (d) η=300 J/mm3. Note the increasein grain size.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    294

  • two factors determine the grain morphology. Zhang et al. (2014) re-ported that a high G and low R caused columnar growth, while a low Gand high R caused equiaxed growth. Rapid solidification and hence,high R values, is inherent to the SLM process. The strategy used in thisstudy of alternating the hatched laser scans results in shorter scantracks, which, as noted by Vrancken et al. (2014), produce lowertemperature gradients (low G). Therefore, the low G/R ratio in thisstudy produces equiaxed grains in the SLM-processed nanocomposite(Fig. 7a).

    In addition to columnar grains, fine equiaxed grains are also ob-served. The fine grains surround the unmelted TiC particles, growingparallel to the surface of the particles due to the high thermal con-ductivity of TiC relative to that of 316L. This generates a higher Gsurrounding the TiC particles, as heat dissipates faster from the liquidphase along the TiC particles. When the melt tracks overlap, G is furtherincreased by re-melting, resulting in the formation of bimodal grains inthe solidified nanocomposite. Zhu et al. (2014) attributed the formationof these directional microstructures to the Gaussian laser heat source,defined by a non-uniform power distribution and a fluctuating energyoutput. This created multiple temperature gradients in different direc-tions within the molten pool. Because of the coupling of multiple nu-cleation sites in the molten pool and the different thermal con-ductivities in the liquid, solid, and powder, sub-structural anisotropywas generated.

    3.4. Thermal behavior prediction

    Because LAMPS simulated the transient temperature field responsetrack-to-track and layer-by-layer for the whole structure system, onlysome time-step results were extracted for comparison and discussed.Figs. 8 and 9 show the transient temperature field changes of the firstseven tracks in the first layers deposited at η=300 J/mm3 and 67 J/mm3, respectively. A higher η value corresponds to a higher peaktemperature, as shown in Figs. 8 and 9.

    The peak temperature can surpass 3200 K at the boundary betweenthe deposited zone and loose powder zones; this is because of the largedifferences in the thermophysical properties of these zones, whichhinder the transfer of laser energy between zones. Based on Takamichiand Roderick’s results, the dynamic viscosity μ of the melt liquid (unit:Pa s) is defined by (Yuan, 2015):

    =μγ m

    k T1615 B (8)

    where γ is the surface tension (Nm−1), m is the atomic mass, kB re-presents the Boltzmann constant, and T is the liquid temperature. Thedynamic viscosity decreases with an increase in the temperature of theliquid. Thus, in the current study, a higher η induces higher tempera-tures in the molten pool. The higher temperatures allow decreases inthe liquid dynamic viscosity and increases in the fluid flow, which isgoverned by the following equations:

    − ∂∂

    =∂∂

    ∂∂

    μ uz

    γT

    Tx (9a)

    − ∂∂

    =∂∂

    ∂∂

    μ vz

    γT

    Ty (9b)

    where u and v are the fluid flow velocities (m s−1) along the x and ydirections.

    From Eq. (9a,b), the fluid flow velocities show strong dependencieson the dynamic viscosity μ, surface tension temperature coefficient ∂

    ∂γT,

    and the temperature gradients ∂∂

    Txand ∂

    ∂Ty. In general, the surface tension

    temperature coefficient is an inherent constant for a given material. Thetemperature gradient becomes the main factor dominating the moltenpool fluid-flow velocity. Here, Fig. 10 shows the detailed effects of η onthe temperature gradient changes during the deposition. Clearly, higherη creates higher temperature gradients. Fig. 11 clarifies this effect byillustrating the relationship between η and the maximum temperaturegradient in three directions. As indicated in Eq. (9a,b) for the fluid flowvelocity, the higher temperature gradient permits higher fluid flowvelocity under the given conditions.

    Since LAMPS simulates the whole deposition procedure, the moltenpool size changes from location to location because of the edge effectand structural geometry. However, the simulations confirm that themolten pool size for higher laser energy densities is larger than that forlower densities. Fig. 12 displays the change in molten pool size at onesimulation step for different laser energy densities.

    Furthermore, since the TiC melting point is much higher than that of316L steel at 3430 K, the TiC particles are not melted and are insteadsuspended in the melt pool. Thus, the TiC particles must be affected byboth the buoyancy effect and the Marangoni effect in the melt pool.Assuming a TiC particle with a radius r in the molten pool, the speed ofthe flotation vBuoy, or buoyancy-driven flow, is given by Stokes’ law as(Bird et al., 1960):

    Fig. 8. Transient temperature field change from track to track in the first layer for η=300 J/mm3.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    295

  • ⎜ ⎟= ⎛⎝

    − ⎞⎠

    vg ρ ρ

    μr

    29Buoy

    melt TiC

    melt (10)

    where ρmelt is the density of the 316L melt, ρTiC is the density of the TiCparticle with the positive direction being against gravity g, and μmelt isthe dynamic viscosity of the molten pool.

    On the other hand, the Marangoni-driven flow vMar of a TiC particleis presented as (Bergman et al., 1988):

    ⎜ ⎟= ⎛⎝ +

    ⎞⎠

    ∂∂

    ∂∂

    v rμ μ

    γT

    TX

    23 2 3Mar melt TiC (11)

    where μTiC is the viscosity of the TiC. For a solid TiC particle, it is equalto zero.

    From Eq. (10), the buoyancy effect always allows TiC to float to thesurface, because the density of TiC is much lower than that of themolten pool. From Eq. (11), the flow speed and direction depend on thetemperature coefficient of surface tension and the temperature gra-dient. For a given condition and subcomponent materials, Eq. (11)shows that the speed of the Marangoni-driven flow depends only on thetemperature gradient, which means that higher temperature gradientscause higher flow speeds. The temperature coefficient of surface tensiondepends on many factors. Because the effects of TiC nanoparticles onthe surface tension of molten pools are unclear, further study is ne-cessary to clarify the flow speed. Regardless, there are only two casesfor the temperature coefficient of surface tension: a negative-value caseand a positive-value case.

    For a negative temperature coefficient of surface tension ∂∂

    γT, the

    Marangoni-driven flow is always directed upward to the surface andthen outward to the molten pool edge. Because the buoyancy-drivenflow speed is always positive, when driving the TiC particle toward thesurface, both the buoyancy and Marangoni effects promote TiC particleflow in the same direction. However, for a positive temperature coef-ficient of surface tension ∂

    ∂γT, the Marangoni-driven flow is directed

    toward the molten pool center and downward to the bottom of the pool.In this case, the buoyancy effect partially or completely cancels theMarangoni effect. Thus, the speed ratio of the Marangoni- and buoy-ancy-driven flows can be introduced to analyze the TiC particle flowcharacteristics, as the ratio of Eq. (11) to Eq. (10):

    = =−

    ∂∂

    ∂∂

    V vv g ρ ρ

    γT

    TX

    32 ( )ratio

    Mar

    Buoy melt TiC (12)

    From Eq. (12), the speed ratio depends strongly on the temperaturegradient as well as other inherent factors. A particle can be suspendedin the molten pool under a suitable temperature gradient; in SLM, thetemperature gradient is controllable, depending on the processingparameters.

    As mentioned above, the effects of TiC nanoparticles on the surfacetension of molten 316L are poorly characterized. However, the existingdata permits some basic analysis. The temperature coefficient of surfacetension for 316L is −4.0× 10−4 Nm−1 K−1 and the viscosity is6.7× 10−3 kgm−1 s−1 (Mills, 2002). The buoyancy-driven flow speedfor a 60-nm-diameter TiC particle is 0.0280m s−1; the respectivespeeds of the Marangoni-driven flow are shown in Table 3. From thisdata, the speed of particle movement is much lower than 0.05m s−1.However, higher laser energy densities (lower scan speeds) entail in-creased molten pool sizes, providing sufficient time for Marangoni-driven flow to finish the flow cycle at lower scan speeds before solidi-fication occurs.

    3.5. Microhardness and tribological performance

    Fig. 13 depicts the effect of the applied energy density on the mi-crohardness values measured on polished cross-sections of SLM-pro-cessed TiC/316L parts. Clearly, the incorporation of TiC particles intothe 316L matrix (average hardness of ∼215 HV0.2) increases thehardness values, irrespective of the applied η. Hardness is influenced byvariations in the cooling rates at various energy densities. The highesthardness values are obtained in samples produced at the highestscanning speed and lowest η (67 J/mm3). This is associated with a finermicrostructure, as observed in the EBSD images (Fig. 7a). Aboveη=67 J/mm3, a sharp reduction in hardness occurs because of theincrease in grain size (Fig. 7b).

    In general, SLM processing at higher η causes two opposing effectson the hardness values of the processed parts (Fig. 13). The reducedscanning speed improves layer melting, which improves binding be-tween the molten particles, creating higher densification and reducedporosity, and hence, better mechanical properties. The average relative

    Fig. 9. Transient temperature field change from track to track in the first layer for η=67 J/mm3.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    296

  • densities of the nanocomposites processed using various η values areshown in Table 4; a detailed discussion of their densification behaviorhas been reported in an earlier study (AlMangour et al., 2018). How-ever, the grain refinement caused by rapid solidification after lasermelting enhanced the mechanical properties according to the Hall–-Petch relationship. Song et al. (2014b) proposed that the high me-chanical strength of SLM-processed specimens was related to the finegrains and high dislocation densities resulting from high cooling rates,while Abe et al. (2003) suggested that high porosities caused low

    ductility in SLM samples. In this study, the lowest energy density re-sulted in higher hardness, despite its low average relative density(Table 4).

    At lower scanning speeds and higher η, the size of the molten poolincreased, and the thermal gradient decreased, causing lower thermaland residual stresses, which decreased the hardness of the sample. Thehigher temperature of the molten pool and the lower cooling rate alsoincrease the grain size due to the longer melt time, which reduceshardness. However, the homogenous distribution of TiC particles

    Fig. 10. Temperature gradient distributions at the center of the second layer deposited at different laser energy densities: (a) η=67 J/mm3, (b) η=125 J/mm3, (c) η=200 J/mm3, and(d) η=300 J/mm3.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    297

  • throughout the matrix enhances nanocomposite hardness.These findings demonstrate the competition between densification

    degree and grain growth in an SLM-processed part. Therefore, in somecases, the microhardness values of SLM nanocomposites increase withincreased η. The increased relative density obtained at higher η canincrease hardness (as was seen in samples processed with higher η va-lues of 200 and 300 J/mm3).

    Fig. 14 shows variations in the average wear rate values of the SLM-processed TiC/316L nanocomposite parts. Although the applied η cri-tically affects wear performance, the trend in wear rate versus η is notnecessarily inversely proportional to the hardness of the nanocompo-sites. The samples processed using the highest scanning speed andlowest η exhibit the highest wear rates of 3.106×10−4 mm3 N−1 m−1,

    which is attributed to the more limited densification level and poreformation relative to other samples. As observed in the SEM images, atthis low η, the TiC nanoparticles form clusters within the 316L matrix,leading to poor microstructural homogeneity (Fig. 4a). The pores andnanoparticle aggregates are both prone to cracking due to stress con-centration during sliding. This forms debris, which separate from thewear surface when the expanded cracks become connected, resulting insignificant increases in the wear rate.

    The characteristic morphologies of the worn surfaces of the testedsamples are shown in Fig. 15. At low η, the wear surface is disrupted,showing deep parallel grooves, large agglomerates, and some ultrafineparticles. The observed abrasive worn morphology in this condition,characterized by the presence of irregular fragments at the grooveedges with severe local deformation and plowing of the surface duringsliding, indicates a relatively high wear rate. By increasing η to 125 J/mm3, a significant amount of loose abrasive fragments remains on theworn surface, but many fewer large agglomerates are observed on theworn surface (Fig. 15b). The worn surface shows slightly shallowergrooves with localized deep portions. This is accompanied by a corre-sponding wear rate of 2.959× 10−4 mm3 N−1 m−1 (Fig. 14). Forη=200 J/mm3, the worn surface of the SLM-processed sample be-comes smooth, with reinforcing nanoparticles embedded in the ad-hesive tribolayer. An adherent and strain-hardened tribolayer com-pletely covers the worn surface of the sample (Fig. 15c). Thischaracteristic led to the lowest wear rate of2.328×10−4 mm3 N−1 m−1 (Fig. 14) obtained in this study.

    The TiC/316L nanocomposite part produced at the optimum η of200 J/mm3 demonstrates superior wear performance. It is reasonable tosuggest that, as the TiC reinforcing nanoparticles are homogeneouslydispersed within the coherent adhesive tribolayer in the SLM-processedpart, the abrasive wear mechanism during prolonged sliding changes totribolayer adhesion. This transition to tribolayer formation is expectedto reduce the wear rate in the sliding tests, as was suggested by Tjong(2007) to explain the uniform distribution of nanoscale TiC reinforce-ments with lower interfacial stresses in the matrix, and thus minimalnanocomposite breakage during sliding. Jain et al. (2010) reported thatimproved densification and increased hardness of a sample allows easyplastic smearing of the protective tribolayer on the worn surface.

    When η is increased further to 300 J/mm3, although some localizedtribolayer formation is observed on the worn surface, it shows signs ofsurface plowing. Parallel grooves as well as much debris were formedon the worn surface after sliding. The wear mechanism in this instanceis micro-plowing rather than adhesion. Some reinforcing particles arealso observed at groove edges on the worn surface. All of this explainsthe slight increase in the average wear rate of2.558×10−4 mm3 N−1 m−1 (Fig. 14) relative to that of the sampleprocessed at η=200 J/mm3. These characteristics of the worn mor-phology observed in this condition could be attributed to the coarsemicrostructure, as well as the resultant disappearance of nanos-tructured TiC reinforcement. TiC tends to split more easily under severesliding wear, which in turn increases the wear rate. Similar observa-tions were reported by Gu et al. (2014).

    Under the optimum η, the nanoscale TiC reinforcements are uni-formly distributed in the matrix (Fig. 4c and g). During sliding wear, thehomogeneously dispersed ultrafine TiC particles stick together, drivenby small reinforcement/matrix interfacial stresses. Friction on the sur-face causes plastic deformation, which produces a strain-hardened layerof refined TiC nanoparticles on the wear surface (Fig. 15c) and therebyimproves the wear properties. At higher η, although the densificationlevel is maintained (Table 4), TiC loses its nanostructure during SLM(Fig. 5). The coarsened TiC particles are thus prone to spalling andsplitting under sliding. The plowing action between the TiC particlesand 316L matrix occurs with a tangential force, producing the deepgrooves on the surface that suggest limited wear performance(Fig. 15d).

    The established relationships between SLM process parameters,

    Fig. 11. Effect of the laser energy density η on the maximum temperature gradient: (a) X,(b) Y, (c) Z.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    298

  • Fig. 12. Molten pool size at the 19th simulation step for different laser energy densities: (a) η=67 J/mm3, (b) η=125 J/mm3, (c) η=200 J/mm3, and (d) η=300 J/mm3.

    Table 3Maximum speed vs. laser energy density.

    Laser energy density (J/mm3) Maximum speed (m/s)

    x-direction y-direction

    67 −0.000635224 −0.001346866125 −0.000597612 −0.001362388200 −0.000770149 −0.001402985300 −0.000853731 −0.001444179

    Fig. 13. Effect of applied η during SLM processing of TiC/316L nanocomposite parts onmicrohardness.

    Table 4Average relative densities of the TiC/316L nanocomposite parts obtained using variousenergy density values during SLM.

    Energy density η (J/mm3) 67 125 200 300Relative density (%) 90.39 93.01 96.10 98.22

    Fig. 14. Effect of applied η during SLM processing of TiC/316L nanocomposite parts ontheir average wear rate.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    299

  • solidification microstructure, and sample hardness assist in under-standing the tribological performance of the TiC/316L nanocomposites.The SLM energy densities strongly influence the tribological perfor-mance of the TiC/316L nanocomposites. Laser processing parametersshould be carefully selected from an optimum processing window toensure refined microstructure with adequate densification in order toachieve the best tribological properties under harsh wear conditions.

    4. Conclusion

    (1) In this study, the distribution state and particle size of nanoscaleTiC reinforcements, as well as the grain size of the nanocompositesin a 316L matrix, can be tailored by regulating η used during SLM.Increasing η from 67 to 300 J/mm3 homogenized the dispersion ofnanoscale TiC reinforcing particles from agglomerates, whilecausing partial disappearance of the reinforcement nanostructureand significant increase in grain size.

    (2) Simulations of the dynamics of the molten pool showed that themaximum temperature gradient within the molten pool was in-creased significantly with increasing η.

    (3) Both microhardness and wear rates were influenced by the degreeof densification, grain coarsening due to lower cooling rates, andthe size and dispersion state of TiC particles in the matrix. Thehighest microhardness was obtained at η=67 J/mm3, which wasattributed to the finest grain structure. The nanocomposites pro-cessed under optimum processing conditions (η=200 J/mm3)showed the lowest wear rate of 2.328×10−4 mm3 N−1 m−1, and astrain-hardened adherent tribolayer formed on the worn surface.

    (4) The experimental study, accompanied by the numerical simulationof temperature evolutions in SLM processing, may clarify thethermal behaviors active in this method of fabrication. Thus, theprocessing parameters during SLM can be optimized to obtain adesirable microstructure and good tribological performance.

    Acknowledgments

    One of the authors, Bandar AlMangour, gratefully appreciates thefinancial support from the Saudi Basic Industries Corporation (SABIC).

    References

    Abe, F., Santos, E.C., Kitamura, Y., Osakada, K., Shiomi, M., 2003. Influence of formingconditions on the titanium model in rapid prototyping with the selective laser meltingprocess. Proc. Inst. Mech. Eng. C J. Mech. Eng. Sci. 217, 119–126.

    AlMangour, B., Grzesiak, D., Yang, J.-M., 2016a. Selective laser melting of TiC reinforced316L stainless steel matrix nanocomposites: Influence of starting TiC particle size andvolume content. Mater. Des. 104, 141–151.

    AlMangour, B., Grzesiak, D., Yang, J.-M., 2016b. Rapid fabrication of bulk-form TiB2/316L stainless steel nanocomposites with novel reinforcement architecture and im-proved performance by selective laser melting. J. Alloys Compd. 680, 480–493.

    AlMangour, B., Grzesiak, D., Yang, J.-M., 2017. Scanning strategies for texture and ani-sotropy tailoring during selective laser melting of TiC/316L stainless steel nano-composites. J. Alloys Compd. 728, 424–435.

    AlMangour, B., Grzesiak, D., Borkar, T., Yang, J.-M., 2018. Densification behavior, mi-crostructural evolution, and mechanical properties of TiC/316L stainless steel na-nocomposites fabricated by selective laser melting. Mater. Des. 138, 119–128.

    Abenojar, J., Velasco, F., Torralba, J.M., Bas, J.A., Calero, J.A., Marcè, R., 2002.Reinforcing 316L stainless steel with intermetallic and carbide particles. Mater. Sci.Eng. A 335, 1–5.

    Arafune, K., Hirata, A., 1999. Thermal and solutal Marangoni convection in In–Ga–Sbsystem. J. Cryst. Growth 197, 811–817.

    Bergman, A., Fredriksson, H., Shahani, H., 1988. The effect of gravity and temperaturegradients on precipitation in immiscible alloys. J. Mater. Sci. 23, 1573–1579.

    Bird, R.B., Stewart, W.E., Lightfoot, E.N., 1960. Transport Phenomena. Wiley, New York.Chen, L.-Y., Peng, J.-Y., Xu, J.-Q., Choi, H., Li, X.-C., 2013. Achieving uniform distribution

    and dispersion of a high percentage of nanoparticles in metal matrix nanocompositesby solidification processing. Scr. Mater. 69, 634–637.

    Composite Solutions and Digital Manufacturing, LLC, 2017. User’s Manual of LayerwiseAdditive Manufacturing Predictions and Simulations (LAMPS) Software.

    Das, M., Balla, V.K., Basu, D., Bose, S., Bandyopadhyay, A., 2010. Laser processing of SiC-particle-reinforced coating on titanium. Scr. Mater. 63, 438–441.

    Gu, D.D., Meiners, W., Wissenbach, K., Poprawe, R., 2012a. Laser additive manufacturingof metallic components: materials, processes and mechanisms. Int. Mater. Rev. 57,133–164.

    Gu, D., Meng, G., Li, C., Meiners, W., Poprawe, R., 2012b. Selective laser melting of TiC/Ti bulk nanocomposites: influence of nanoscale reinforcement. Scr. Mater. 67,185–188.

    Fig. 15. SEM images showing typical morphologies of the worn surfaces of TiC/316L parts processed using (a) η=67 J/mm3, (b) η=125 J/mm3, (c) η=200 J/mm3, and (d) η=300 J/mm3.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    300

    http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0005http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0005http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0005http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0010http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0010http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0010http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0015http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0015http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0015http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0020http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0020http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0020http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0025http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0025http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0025http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0030http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0030http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0030http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0035http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0035http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0040http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0040http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0045http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0050http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0050http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0050http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0055http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0055http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0060http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0060http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0065http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0065http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0065http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0070http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0070http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0070

  • Gu, D., Shen, Y., 2009. Effects of processing parameters on consolidation and micro-structure of W–Cu components by DMLS. J. Alloys Compd. 473, 107–115.

    Gu, D., Shen, Y., Fang, S., Xiao, J., 2007. Metallurgical mechanisms in direct laser sin-tering of Cu–CuSn–CuP mixed powder. J. Alloys Compd. 438, 184–189.

    Gu, D., Wang, H., Chang, F., Dai, D., Yuan, P., Hagedorn, Y.C., 2014. Selective lasermelting additive manufacturing of TiC/AlSi10Mg bulk-form nanocomposites withtailored microstructures and properties. Phys. Proc. 56, 108–116.

    Hao, L., Dadbakhsh, S., Seaman, O., Felstead, M., 2009. Selective laser melting of astainless steel and hydroxyapatite composite for load-bearing implant development.J. Mater. Process. Technol. 209, 5793–5801.

    Jain, A., Basu, B., Kumar, B.M.J., Harshavardhan, Sarkar, J., 2010. Grain size–wear raterelationship for titanium in liquid nitrogen environment. Acta Mater. 58, 2313–2323.

    Kim, H.-C., Oh, D.-Y., Shon, I.-J., 2004. Sintering of nanophase WC–15 vol.%Co hardmetals by rapid sintering process. Int. J. Refract. Met. Hard Mater. 22, 197–203.

    Kruth, J.-P., Badrossamay, M., Yasa, E., Deckers, J., Thijs, L., Van Humbeeck, J., 2010.Part and material properties in selective laser melting of metals. In: Zhao, W., Ye, J.,Zhu, D. (Eds.), Proc. 16th Int. Symp. Electromach. Shanghai, China. pp. 1–12.

    Kruth, J.-P., Levy, G., Klocke, F., Childs, T., 2007. Consolidation phenomena in laser andpowder-bed based layered manufacturing. CIRP Ann. Manuf. Technol. 56, 730–759.

    Lee, G.-H., Kang, S., 2006. Sintering of nano-sized WC–Co powders produced by a gasreduction–carburization process. J. Alloys Compd. 419, 281–289.

    Liu, Y., Lan, F., Yang, G., Zhou, Y., 2004. Microstructural evolution of rapidly solidifiedTi–Al peritectic alloy. J. Cryst. Growth 271, 313–318.

    Ma, M., Wang, Z., Wang, D., Zeng, X., 2013. Control of shape and performance for directlaser fabrication of precision large-scale metal parts with 316L stainless steel. Opt.Laser Technol. 45, 209–216.

    Mills, K.C., 2002. Recommended Values of Thermophysical Properties for SelectedCommercial Alloys. Woodhead Publishing Ltd., Cambridge, UK, pp. 135–146.

    Moya, J.S., Lopez-Esteban, S., Pecharroman, C., 2007. The challenge of ceramic/metalmicrocomposites and nanocomposites. Prog. Mater. Sci. 52, 1017–1090.

    Niu, H.J., Chang, I.T.H., 1999. Selective laser sintering of gas and water atomized highspeed steel powders. Scr. Mater. 41, 25–30.

    Pagounis, E., Lindroos, V.K., 1998. Processing and properties of particulate reinforcedsteel matrix composite. Mater. Sci. Eng. A 246, 221–234.

    Pei, Y.T., Galvan, D., De Hosson, J.Th.M., 2005. Nanostructure and properties of TiC/a-C:H composite coatings. Acta Mater. 53, 4505–4521.

    Saheb, N., Iqbal, Z., Khalil, A., Hakeem, A.S., Al Aqeeli, N., Laoui, T., Al-Qutub, A.,Kirchner, R., 2012. Spark plasma sintering of metals and metal matrix nanocompo-sites: a review. J. Nanomater. 2012, 18. http://dx.doi.org/10.1155/2012/983470.

    Schwarz, M., Arnold, C.B., Aziz, M.J., Herlach, D.M., 1997. Dendritic growth velocity anddiffusive speed in solidification of undercooled dilute Ni-Zr melts. Mater. Sci. Eng. A

    226, 420–424.Sheikhzadeh, M., Sanjabi, S., 2012. Structural characterization of stainless steel/TiC na-

    nocomposites produced by high-energy ball-milling method at different millingtimes. Mater. Des. 39, 366–372.

    Simchi, A., Petzoldt, F., Pohl, H., 2001. Direct metal laser sintering: Material considera-tions and mechanisms of particle: Rand tooling of powdered metal parts. Int. J.Powder Metall. 37, 49.

    Simchi, A., 2006. Direct laser sintering of metal powders: mechanism, kinetics and mi-crostructural features. Mater. Sci. Eng. A 428, 148–158.

    Simchi, A., 2008. Effect of C and Cu addition on the densification and microstructure ofiron powder in direct laser sintering process. Mater. Lett. 62, 2840–2843.

    Simchi, A., Asgharzadeh, H., 2004. Densification and microstructural evaluation duringlaser sintering of M2 high speed steel powder. Mater. Sci. Technol. 20, 1462–1468.

    Smallman, R.E., Ngan, A.H.W., 2013. Modern Physical Metallurgy, third ed. Elsevier,Oxford.

    Song, B., Dong, S., Coddet, C., 2014a. Rapid in situ fabrication of Fe/SiC bulk nano-composites by selective laser melting directly from a mixed powder of microsized Feand SiC. Scr. Mater. 75, 90–93.

    Song, B., Dong, S., Deng, S., Liao, H., Coddet, C., 2014b. Microstructure and tensileproperties of iron parts fabricated by selective laser melting. Opt. Laser Technol. 56,451–460.

    Tjong, S.C., 2007. Novel nanoparticle-reinforced metal matrix composites with enhancedmechanical properties. Adv. Eng. Mater. 9, 639–652.

    Viswanathan, V., Laha, T., Balani, K., Agarwal, A., Seal, S., 2006. Challenges and ad-vances in nanocomposite processing techniques. Mater. Sci. Eng. R Rep. 54, 121–285.

    Vrancken, B., Thijs, L., Kruth, J.-P., Van Humbeeck, J., 2014. Microstructure and me-chanical properties of a novel β titanium metallic composite by selective lasermelting. Acta Mater. 68, 150–158.

    Wei, Q., Li, S., Han, C., Li, W., Cheng, L., Hao, L., Shi, Y., 2015. Selective laser melting ofstainless-steel/nano-hydroxyapatite composites for medical applications:Microstructure, element distribution, crack and mechanical properties. J. Mater.Process. Technol. 222, 444–453.

    Yuan, P., Gu, D., 2015. Molten pool behaviour and its physical mechanism during se-lective laser melting of TiC/AlSi10Mg nanocomposites: simulation and experiments.J. Phys. D. Appl. Phys. 48, 035303.

    Zhang, D.Q., Liu, Z.H., Cai, Q.Z., Liu, J.H., Chua, C.K., 2014. Influence of Ni content onmicrostructure of W–Ni alloy produced by selective laser melting. Int. J. Refract. Met.Hard Mater. 45, 15–22.

    Zhu, Y., Liu, D., Tian, X., Tang, H., Wang, H., 2014. Characterization of microstructureand mechanical properties of laser melting deposited Ti–6.5 Al–3.5 Mo–1.5 Zr–0.3 Sititanium alloy. Mater. Des. 56, 445–453.

    B. AlMangour et al. Journal of Materials Processing Tech. 257 (2018) 288–301

    301

    http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0075http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0075http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0080http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0080http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0085http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0085http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0085http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0090http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0090http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0090http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0095http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0095http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0100http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0100http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0105http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0105http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0105http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0110http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0110http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0115http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0115http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0120http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0120http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0125http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0125http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0125http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0130http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0130http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0135http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0135http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0140http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0140http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0145http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0145http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0150http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0150http://dx.doi.org/10.1155/2012/983470http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0160http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0160http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0160http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0165http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0165http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0165http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0170http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0170http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0170http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0175http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0175http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0180http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0180http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0185http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0185http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0190http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0190http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0195http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0195http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0195http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0200http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0200http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0200http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0205http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0205http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0210http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0210http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0215http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0215http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0215http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0220http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0220http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0220http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0220http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0225http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0225http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0225http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0230http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0230http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0230http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0235http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0235http://refhub.elsevier.com/S0924-0136(18)30029-3/sbref0235

    Thermal behavior of the molten pool, microstructural evolution, and tribological performance during selective laser melting of TiC/316L stainless steel nanocomposites: Experimental and simulation methodsIntroductionExperimental procedures and simulation setupPowder preparation and SLM processingMicrostructural characterizationModeling approach and numerical simulationPhysical model description and governing equationsPhysical propertiesSimulation parameters

    Hardness and wear testing

    Results and discussionEffects of applied η on the distribution state of TiCEffects of applied η on TiC particle sizeEffects of applied η on grain sizeThermal behavior predictionMicrohardness and tribological performance

    ConclusionAcknowledgmentsReferences