Anda di halaman 1dari 14

Materials and Design 98 (2016) 344–357

Contents lists available at ScienceDirect

Materials and Design

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

Selective laser melting (SLM) of AlSi12Mg lattice structures


Martin Leary a,⁎, Maciej Mazur a, Joe Elambasseril a, Matthew McMillan a, Thomas Chirent b, Yingying Sun a,
Ma Qian a, Mark Easton a, Milan Brandt a
a
RMIT Centre for Additive Manufacture, RMIT University, Melbourne, Australia
b
Institut Français de Mécanique Avancée, Clermont-Ferrand, France

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

Article history: Additive manufacture (AM) enables innovative structural design, including the fabrication of complex lattice
Received 26 October 2015 structures with unique engineering characteristics. In particular, selective laser melting (SLM) is an AM process
Received in revised form 26 February 2016 that enables the manufacture of space filling lattice structures with exceptional load bearing efficiency and
Accepted 29 February 2016
customisable stiffness. However, to commercialise SLM lattice structures it is necessary to formally define the
Available online 6 March 2016
manufacturability of candidate lattice geometries, and characterise the associated mechanical response including
Keywords:
compressive strength and stiffness. This work provides an experimental investigation of the mechanical proper-
Additive manufacture ties of SLM AlSi12Mg lattice structures for optimised process parameters as well as the manufacturability of lat-
Lattice structures tice strut elements for a series of build inclinations and strut diameters. Based on identified manufacturability
Selective laser melting limits, lattice topologies of engineering relevance were fabricated, including both stretch-dominated and
Aluminium bending-dominated structures. Manufactured lattice morphology and surface roughness was quantified, as
AlSi12Mg were the associated mechanical properties and deformation and failure behaviour.
Optimisation © 2016 Elsevier Ltd. All rights reserved.
Design for additive manufacture

1. Additive manufacture and selective laser melting SLM material research is typically focused on titanium alloys due to
their high corrosion resistance, high specific strength [16–22] and bio-
Additive manufacturing (AM) refers to the process of fabricating compatibility [23]. These favourable properties have enabled a range
near-net components from unit materials [1] in a layerwise manner. of technical SLM titanium applications, including: customised load bear-
AM differs fundamentally from “subtractive manufacturing methodolo- ing medical implants [24,25], lattice structures with mechanical proper-
gies, such as traditional machining” [2], and has been labelled as a ties optimised for compatibility with bone [26], high-value aerospace
disruptive technology [3]. The disruptive nature of AM enables a series [27]; space [28], and automotive applications [29].
of technical and economic advantages, including: enhanced cost- Although aluminium alloys are not biocompatible, they exhibit a
competitiveness for low volume production [4]; reduced environmental number of properties that make them eminently suitable for commer-
impact of manufacture [5]; and, increased design complexity [3,6]. cial AM applications, including:
AM is based on a digital representation of product geometry. This
representation typically consists of tessellated triangular facets, known • Aluminium (2.7 g/cm3) has significantly lower density than titanium
as a Stereolithography (STL) file [7,8]. However, enhanced data struc- (4.5 g/cm3), thereby enabling commercial opportunities in scenarios
tures are also proposed that are more compatible with specific AM re- that benefit from mass reduction. The mass reduction advantage of
quirements, such as STL 2.0 [9], Additive Manufacturing Format (AMF) AM aluminium lattice structures is especially relevant for non-
[10] and Bezier STL [11]. stationary scenarios such as automotive and aerospace where the
Selective laser melting (SLM) is a powder-bed AM technology that fuel consumption associated with component mass is significant [30].
uses a scanning laser to sequentially melt layers of powdered metal • The high strength-to-weight ratio of aluminium enables optimisation
under an inert atmosphere [12]. The significant advantage of SLM tech- for structural applications, especially for structures subject to bending
nology includes high flexibility and achievable component complexity, loads such as beam and plate applications [31]. This outcome is partic-
enabling the fabrication of highly complex lattice structures that are ularly relevant to the optimisation of AM lattice structures, which are
not otherwise manufacturable [13]. Such lattice structures enable struc- subject to significant bending, either during elastic loading (in the
tural optimisation [14], while reducing associated manufacture time case of under-stiff structures) or plastic collapse (in the case of just-
and materials costs by minimising component volume [15]. stiff and over stiff structures) (Section 3).
• The fatigue-strength of aluminium, in combination with low material
⁎ Corresponding author. density, results in optimal material selection for weight limited design
E-mail address: martin.leary@rmit.edu.au (M. Leary). where the number of loading cycles is below, or close to the

http://dx.doi.org/10.1016/j.matdes.2016.02.127
0264-1275/© 2016 Elsevier Ltd. All rights reserved.
M. Leary et al. / Materials and Design 98 (2016) 344–357 345

• Importantly, aluminium alloys often provide a lower cost to achieve a


Nomenclature
specified load bearing function in comparison to titanium [30], and
therefore provide an opportunity for the economical commercialisation
Term Definition
of AM technologies, especially in non-stationary applications.
AM Additive manufacture
CT Computed tomography
BCC Body centred cubic unit cell Despite the commercial opportunities associated with aluminium
BCCZ Body centred cubic unit cell with (vertical) Z-struts lattice structural design in SLM, there appears to be sparse available re-
DOE Design of experiments search in comparison with titanium. This lack of research may be in part
FBCCZ Face and body centred cubic unit cell with (vertical) Z- due technical challenges in SLM of aluminium powders, including poor
struts powder flowability, high laser reflectivity and oxidation of aluminium
FCC Face centred cubic unit cell during melting [40].
FCCZ Face centred cubic unit cell with (vertical) Z-struts
FEA Finite element analysis 2. SLM processing and properties
SLM Selective laser melting
SLM is a highly complex multiphysics process, due to: large thermal
Symbols gradients [41]; complex three-dimensional geometries [42]; local
Term Definition overheating [43,44]; local heat transfer paths that are temporally and spa-
Dv Volume weighted percentile particle diameter tially transient [45,46]; and, thermal powder bed resistance that is poorly
D [3,2] Surface area moment mean (Sauter Mean Diameter) understood and subject to significant experimental uncertainty [47,48].
D [4,3] Volume moment mean (De Brouckere Mean Diameter) SLM is highly dimensional, with nearly 130 variables of influence on
E Lattice specimen modulus [MPa] final component quality [49,50].
Es Solid material modulus [MPa] Processing parameters that determine the laser energy density
h Hatch spacing [μm] (Eq. (1)) are: laser power, P, laser scanning speed, v, hatch spacing, h,
j Roughness sample index and the layer thickness, t.
kR Relative stiffness
M Maxwell number Where :  
n Number of nodes ϒ : energy density J=mm3
N Number of roughness measures per sample, number of P Ρ : laser power ½W 
Υ¼ : ð1Þ
test specimens h  v  t ν : scan speed ½mm=s
ρ Lattice specimen density [kg/m3] t : layer thickness ½mm
ρc Lattice unit cell density [kg/m3] h : hatch spacing ½mm
ρs Lattice solid material density [kg/m3]
P Laser power [W] 2.1. SLM processing of aluminium alloys
Ra The average surface roughness [µm]
Rt The distance between the maximum peak and mini- Aluminium alloys provide significant technical and economic oppor-
mum valley [µm] tunities due to their favourable strength-to-weight and cost-to-weight
Rz The peak to peak roughness [µm] attributes. Aluminium-Silicon alloys are strong candidates for SLM due
s Number of struts to their robust welding characteristics and small difference between
σ Lattice specimen compressive strength [MPa] liquidus and solidus temperatures [51]. However, in-comparison with
σs Solid material compressive strength [MPa] titanium alloys, very little research exists on the processing of alumini-
t Layer thickness [mm] um by SLM [52]. A review of the available literature has identified the
v Scan speed [mm/s] following research of interest:
Υ Laser energy density [J/mm3]
• Kempen et al. [51] found that for AlSi10Mg (mean particle size: 16 and
W Volumetric energy absorption [MJ/m3]
48 μm), the manufactured part quality varies with powder morpholo-
gy, size and chemical composition.
endurance limit [30]. The robust design of fatigue-optimised AM lat- • Olakanmi et al. [53] assessed the effects of SLM process parameters on
tice structures is an active research domain [32,33] that will enable the degree and orientation of porosity in Al‐12Si (Particle size distri-
commercially relevant innovation in the design of fatigue-limited alu- bution: 45‐75 μm).
minium lattice structures. • Dadbakhsh and Hao [54] investigated the manufacturability of nu-
• The high thermal diffusivity of aluminium alloys combined with the merous aluminium-based powders, including commercially pure alu-
capability of AM for complex part geometry enables the fabrication minium, AlSi10Mg, Al Mg1SiCu, mixed with 15 wt% Fe2O3, and found
of high-efficiency thermal devices, including heat sinks and heat ex- AlSi10Mg to have the highest density due to low thermal conductivity
changers [34,35]. and oxide layer breakdown; a nano-scale dendritic structure was ob-
• Aluminium has excellent resistance to corrosion [36], enabling served leading to high hardness.
engineered aluminium lattice structures to be applied in corrosive en- • Dadbakhsh and Hao [55] investigated SLM of numerous aluminium-
vironments, for example in [37]: architectural and marine applica- based powders including: Al‐5FeO3, Al‐10FeO3, Al‐15FeO3, and found
tions; faying surfaces in automotive and aerospace applications, and that Hot Isostatic Pressing (HIP) increased density but did not eliminate
in the design of heat-exchangers in direct contact with corrosive all porosity due to the formation of oxide bands.
fluids. • Buchbinder et al. [56] found that for AlSi10Mg, high laser power and
• Aluminium has relatively high electrical conductivity [38], enabling its scan velocity enabled a build rate of 21mm3/s with 99.5% density
application in electrically conductive devices. The combined electrical, (420 MPa and 145 HV).
structural and lightweight properties of aluminium enable the fabri- • Brandl et al. [57] found that for AlSi10Mg with a particle size distri-
cation of high-value components that combine both electrical and bution of 25–40 μm, post manufacture heat treatment has the
structural function, for example in self-supporting lightweight electri- greatest effect on fatigue limit whilst orientation has a relatively
cal conductors [39]. smaller effect.
346 M. Leary et al. / Materials and Design 98 (2016) 344–357

Table 1
Chemical composition of AlSi12Mg powder and solid material using process parameters of Table 2.

Sample (wt.%) Al Si Cu Fe Mg Zn Cr Ni Mn Ti Sr Zr O

Powder Bal 11.7 b0.01 0.15 0.39 b0.01 b0.001 b0.005 0.053
Solid Bal 11.6 b0.01 0.15 0.37 b0.01 b0.001 b0.005 0.051

• Kempen et al. [58] identified SLM process parameters for AlSi10Mg excess of 200 W, part densities approaching 100% have been achieved
that obtain mechanical properties comparable to cast materials: with aluminium alloys including AlSi10Mg [40].
E = 68 GPa, UTS = 396 MPa, HV = 127.
• Ghosh and Saha [59] assessed crack density and wear performance 2.2. AlSi12Mg process parameter optimisation
for SLM processed aluminium Metal Matrix Composites (MMCs) to
fabricate wear resistant components. To characterise the effect of SLM process parameters on the quality
• Wong et al. [60] showed that aluminium alloy 6061 and stainless of the SLM manufactured AlSi12Mg, a design of experiments (DOE)
steel grade 316 L are viable materials for the fabrication of SLM was developed to study the effect of processing parameters on the po-
processed heat sinks. rosity of 4 mm cubes built with a 50 μm layer thickness.
• Ameli et al. [61] demonstrated the application of alloy 6061 for fab- The powder was provided by SLM Solutions GMBH and was
rication of porous wick structures. characterised by laser diffraction [70] to obtain a volume weighted par-
• Li et al. [62] investigated the mechanical properties of SLM proc- ticle size distribution and associated distribution statistics according to
essed Al\\12Si, and report as-manufactured yield and tensile ISO 9276–2:2014 [71] (Fig. 2). Powder chemical composition of powder
strengths of approximately 225 MPa and 360 MPa, with ductility and solid specimens was determined using inductively coupled plasma-
of around 5%. This work also reports on the effect of solution heat atomic emission spectroscopy (ICP-AES) (Table 1).
treatment on the trade-off between strength and ductility. Test cubes were manufactured with varying laser power and scan-
ning speed and cube porosity was measured with computed tomogra-
phy (CT) scanning. In summary:
The limited research data available for SLM processing of aluminium
may be in part due to technical challenges of processing aluminium in • The process parameters of Table 2 achieve a material density of 99.86%
comparison to titanium or steel powders [40,52,63]. Technical challenges with no cracks present.
associated with SLM processing include robust powder deposition and • Etched microstructures (Fig. 1) indicate the semi-elliptic structure of
achieving high fused density. solidified melt pool typical of SLM processing.
SLM process repeatability requires that the powder layer be consis- • The track boundary has been partially remelted and heat affected and
tently deposited; this is technically challenging with aluminium due to tends to be coarser than the structure produced from the melt pool.
low density and poor flowability, especially in the presence of humidity
or moisture [52]. This challenge is an active research area and can be ad-
Tensile specimens (N = 6, Specimen 3, E8M [72]) were subsequent-
dressed by appropriate machine design and powder specification.
ly manufactured with the following properties:
There exist several distinct technical challenges to the dense fusion of
aluminium in the SLM process, including challenges associated with • Yield strength is 236.1 ± 5.4 MPa which is consistent with literature
power absorption, heat diffusion and oxidisation effects. An aluminium reported data [56], and exceeds the requirements for die-cast material
powder layer has relatively high absorptivity due to internal reflections [73], and is slightly higher than the reported strength for SLM proc-
within the powder layer [64]; however, fused aluminium has relatively essed Al‐12Si [62].
high reflectivity, resulting in a relatively low effective heating for a • Tensile strength is 434.1 ± 14.0 MPa, which exceeds data reported in
given laser power. Relative to titanium-based alloys, aluminium alloys the literature for AlSi10Mg, i.e. 391 ± 6 (XY direction) to 396 ± 8 MPa
also have high thermal diffusivity [65], thereby resulting in more rapid (Z direction) [74], as well as for Al‐12Si [62].
heat transfer from the melt pool [66]. The combined effects of high reflec- • Strain to failure is 4.6 ± 0.5%, which is comparable with data reported
tivity and high thermal diffusivity thereby require commensurately for Al‐12Si [62], and significantly exceeds the minimum standard for
higher power to achieve powder fusion. die-cast AlSi12Mg [73].
A particular challenge to the fusion of aluminium powder is the pro-
pensity for aluminium to oxidise [67]. Particle surface oxidisation can
occur during powder production, handling, or during SLM processing; The identified parameters were subsequently used in the manufac-
and presents difficulties in robust SLM processing due to: increased resis- ture of lattice structures used in this work.
tance to heat diffusion from the particle surface to the core [68], reduced
wettability due to the presence of oxide films, and the potential for poros- 3. Structural lattice
ity and entrapments as the oxide mixes with molten aluminium alloy
[69]. The mechanical properties of lattice structures, in particular com-
Louvis et al. [52] proposes that oxide films are the technical basis for pressive strength and modulus is dependent on factors such as: cell to-
the relative difficulty of SLM processing aluminium and studied the ef- pology; structural boundary and loading conditions; as well as the
fect of oxide films on the processing of Al‐12Si and 6061 aluminium relative density (ρ/ρs) of the lattice density (ρ) to the solid material
powders. With a laser power of 100 W, Louvis et al. were able to achieve density (ρs) [75]. Mechanical properties are particularly influenced by
a solid-section Al‐12Si density approaching 90%. By using laser power in whether the cell topology results in a loading response of lattice struts

Table 2
Optimal AlSi12Mg SLM processing parameters.

Laser energy Laser power, Hatch spacing, Focal Scan speed, Layer thickness, Relative
density, Υ (J/mm3) P (W) h (mm) offset (mm) v (mm/s) t (μm) density (%)

40 350 0.19 2 921.05 50 98.86


M. Leary et al. / Materials and Design 98 (2016) 344–357 347

Fig. 3. Examples of (i) under-stiff, (ii) just-stiff and (iii) over-stiff truss structures.

(b) • An under-stiff lattice structure (M b 0) occurs when the lattice has too
few struts to be statically determinant and will act as a mechanism
unless moments are transferred at nodes, for example, Fig. 2 (i). An
(a) under-stiff lattice shows bending-dominated behaviour and has high
compliance and relatively low strength.
• A just-stiff lattice structure (M = 0) occurs when the lattice has the
minimum number of struts required for the lattice to be statically
determinant, for example, Fig. 1 (ii). A just-stiff lattice does not re-
quire the transmission of moments at nodes to maintain structural
(b) integrity, and strut loads can be determined analytically by refer-
ence to the equations of equilibrium. A just-stiff lattice has rela-
tively low compliance (high stiffness), and relatively high
strength. Such a structure exhibits stretch-dominated behaviour
and has high structural efficiency.
• An over-stiff lattice structure (M N 0) occurs when the lattice has
an excess of struts required to achieve static equilibrium, for exam-
ple, Fig. 1 (iii). An over-stiff structure does not require the trans-
mission of moments at nodes to maintain structural integrity, and
(a) strut loads can only be calculated by assessing the effect of strut
stiffness. An over-stiff lattice has very low compliance (high stiff-
ness) and high strength due to the reinforcing effect of the redun-
Fig. 1. Optical micrographs of etched AlSi12Mg specimens at two different magnifications. dant strut elements [80]. Such structures exhibit stretch-
The arrow (a) indicates the direction normal to the platen. Sample of course dominated behaviour [81].
microstructure at track boundary identified by arrows (b).

that is either stretch-dominated or bending-dominated [76]. It is possi- In regards to the relative density of lattice cells, mechanical proper-
ble to gain insight into the expected loading response by considering the ties such as the relative strength σ/σs (where σ denotes the compres-
degree of strut connectivity and associated degrees of freedom present sive strength of the lattice structure and σs of the solid material) and
in a lattice cell [77]. These can be quantified by the Maxwell number (M) relative modulus E/Es typically exhibit a positive power relationship
(Eq. (2)) which quantifies the expected mechanical behaviour in a lat- with the relative density (ρc/ρs) [75]. Bending-dominated structures
tice consisting of a number of interacting struts, s, pin-jointed at a num- also typically exhibit proportionally lower values of specific compres-
ber of nodes, n, (Fig. 3) [78]: sive strength and modulus compared to stretch-dominated structures;
specific relationships are described in Section 6.
M ¼ s−2n þ 3 For a two−dimensional truss
ð2Þ
M ¼ s−3n þ 6 For a three−dimensional truss

The combination of struts and nodes (intersections) results in one of


the following mechanical behaviours [79]:

Fig. 4. For a cubic unit cell with body and face nodes, there are 3 possible strut inclination
Fig. 2. Volume weighted particle size distribution and associated distribution statistics. angles. Blue: 0°, Green: 35.3°, Yellow: 45° Red: 90°.
348 M. Leary et al. / Materials and Design 98 (2016) 344–357

4. Lattice manufacturability 4.1. Experimental investigation of strut manufacturability

SLM is a thermo-mechanical process with associated manufacturabil- This work responds to this identified deficiency by experimentally
ity limits. In particular, structures with acutely overhanging faces and geo- quantifying manufacturability of aluminium lattice strut elements. To
metrical features with excessive heat transfer resistance may not be establish the feasible build angle and feature size relevant for AlSi12Mg,
manufacturable [12,49]. SLM structures that violate manufacturability SLM cantilever strut test specimens were manufactured with a broad
limits require a support structure in order to achieve the required build design of experiments (DOE) analysis of manufacturability versus
geometry. Support structures offer significant design freedom [82]; how- strut diameter and associated inclination. For a unit cell with equal
ever the use of support material can increase surface roughness, compro- side length, there exist four possible strut inclination angles: 0, 35.26,
mise part functionality and is difficult to remove especially when 45, and 90° (Fig. 4).
constrained by limited access [83], as in lattice structures. To ensure ro- Sample specimens were assessed for manufacturability at these
bust lattice manufacture, it is necessary to experimentally determine angles with a range of strut diameters (Table 3). This work experi-
the minimum achievable strut diameter and strut inclination angles that mentally identified that horizontal struts (0° inclination) are not
can be manufactured without the need for support structures. manufacturable at any assessed diameter; for this case the strut
Data for Ti64 SLM lattice manufacturability is available, including: was not able to remain in contact with part and fused material was
wiped away by the recoater. The assessed strut diameters are manu-
• Successful production of 0.2 mm diameter struts is commonly report- facturable for angles trialled other than 0°, although for these scenar-
ed, VanBael et al. [84] were able to produce a strut diameter of 0.1 mm ios there is variation in geometric fidelity of the manufactured struts.
and Hao and Raymond [85] state that 0.05 diameter mm struts are In this work the strut length was defined to be significantly larger
possible than the strut diameter, in order to assess worst-case manufactur-
• A reported difference between virtual and as built geometry in terms ability; however research suggests that manufacturability can be im-
of geometric accuracy and surface quality. This variation is related to proved for low inclination struts by reducing the spanned distance
the limitations of SLM including meltpool characteristics, powder [50]. Further work is required to extend this broad manufacturability
size [84,86], material type [52] and process parameters [87]. Micro- DOE in order to fully characterise AlSi12Mg SLM manufacturability
CT and optical inspection have shown this variation and identified it for all geometric configurations.
as a function of build angle [84,88].
5. Lattice specimen manufacture

Despite the technical opportunities associated with aluminium lat- Based on the experimental study of aluminium lattice manufactur-
tice structures, the authors are aware of no formal experimental data ability, a series of manufacturable lattice structures with varying param-
on aluminium lattice manufacturability. eters (such as topology, Maxwell Number and strut alignment to the

Fig. 5. As manufactured specimens. (i) (BCC) body centred cubic (ii) (BCCZ) body centred cubic with Z struts (iii) (FCC) face centred cubic (iv) (FCCZ) face centred cubic with Z struts
(v) (FBCCZ) face and body centred cubic with Z struts (vi) Inset describes typical cell dimensions.
M. Leary et al. / Materials and Design 98 (2016) 344–357 349

Fig. 6. μCT scan data of a 2 × 2 × 2 array subset of the manufactured lattice specimen array.

loading direction) were selected for experimental testing in order to as- • Face centred cubic (FCC): High compliance.
sess mechanical performance under compressive loading (Table 4). Lat- • Body centred cubic with vertical struts (BCCZ) and Face centred cubic
tice nomenclature is based on conventions in crystallography with with vertical struts (FCCZ): Theoretically under-stiff but high compres-
additional designation for vertical (Z) struts [89]. Specific lattice struc- sive strength and stiffness along z-direction struts.
tures were selected due to the following range of expected properties: • Face and body centred cubic with vertical struts (FBCCZ): Theoretically
under-stiff but high strength and stiffness along z-direction struts, with
• Body centred cubic (BCC): High compliance and low relative density body centred struts resisting loads in x and y-directions. Comparatively
(Table 5). high relative density.

Fig. 7. Representative optical microscope image of the lower and upper faces of 35.26°, 45°and 90° struts.
350 M. Leary et al. / Materials and Design 98 (2016) 344–357

Fig. 10. Volumetric energy absorption (W) of tested lattice specimens.

based on a 2 × 2 × 2 cell array. In particular μCT analysis indicates that


Fig. 8. Summary surface roughness data (error bars indicate ±1 SD). (Fig. 6):

• Downward facing surfaces consistently display higher than average


surface roughness; specific analysis of strut surface roughness was
completed to quantify the effect of geometry and inclination on
The selected lattice structures were successfully manufactured (Fig. 4)
strut roughness (Section 5.2)
with the previously identified processing parameters (Table 2). A strut di-
• Visible geometric defects occur when struts intersect at nodes. These
ameter of 1.0 mm was selected due to manufacturability constraints; 10
defects appear to be asymmetric even for lattice structures with nom-
cells were specified with a cell size of 7.5 mm such that the overall lattice
inally symmetric geometry. The effect of node geometry on structural
structure size (75 mm cubed) allowed all specimens to be manufactured
integrity has been assessed by computational analysis of the μCT data
concurrently on one 250 mm square build plate.
(Section 6.2).
The complex transient thermal effects, layerwise geometry and sto-
chastic effects characteristic of SLM manufacture of aluminium, result in
variation between the intended CAD geometry (Table 5) and the as-
manufactured geometry. To provide insight into variation of lattice 5.2. Surface roughness evaluation
geometry, destructive and non-destructive methods were used to qual-
itatively assess the as-manufactured lattice morphology, and the associ- The struts of cell topologies assessed in this work exhibit build an-
ated strut roughness. gles of 35.26°, 45° and 90° to the platen (Table 5). To assess the associ-
ated surface roughness, three struts were mechanically removed from
the manufactured lattice array for each build angle. An optical micro-
5.1. Geometric fidelity of as-manufactured lattice geometry scope (5× magnification, 200 ms exposure time, transmission settings)
was used to capture a shadow profile as an image with a nominal sur-
Manufactured specimens were inspected using micro Computed To- face length of 1.41 mm (Fig. 7). Four angular orientations of interest
mography (μCT) to provide quantitative insight into cell morphology were assessed: upper and lower surfaces have surface normal oriented

Fig. 9. Uniaxial compression stress-strain responses of tested lattice specimens.


M. Leary et al. / Materials and Design 98 (2016) 344–357 351

Inspection of the surface roughness data (Fig. 8) indicates:

• Lower strut surfaces have an order of magnitude higher surface


roughness than the other sampled surfaces. The increased roughness
is due to preferential particle adhesion on downward facing surfaces
on account of limited heat transfer between the powder bed and
solid material [12,84,88].
• The average, Ra, and peak-to-peak, Rz, roughness of lower surfaces is
highest for 35.26° struts. For example (Table 6), for 35.26° struts,
Ra = 120.2 ± 39.7 μm and Rz = 456.5 ± 113.6 μm; for 45° struts,
Ra = 43.52 ± 20.56 μm and Rz = 235.5 ± 118.2 μm.
• Side and upper surfaces have average and peak-to-peak surface rough-
ness that is statistically indistinguishable.

6. Experimental results
Fig. 11. Loading response for a BCCZ specimen identifying Young's modulus as well as
average of loading and unloading moduli for 1% and 2% strain.
The manufactured lattice specimens (Fig. 5) were tested in quasi-
static uniaxial compression in a calibrated universal testing machine
away from, and toward the platen, respectively; side surfaces have sur- to quantify compressive strength and modulus as well as failure
face normal parallel to the platen. Note the 90° struts do not exhibit modes. A strain rate of 1 × 10−3 s−1 was used according to applicable
upper or lower faces and the reported result is the average for all four testing standards [91]. The strain was determined from machine cross-
measured perpendicular faces. Average roughness, Ra, and peak-to- head displacement and the compressive stress was calculated based on
peak roughness, Rz, was assessed for each orientation. Two positions applied compression force and the initial XY plane cross-sectional base
were measured for 3 struts, resulting in 6 samples for every combina- area of each lattice specimen. Tests were conducted on two repetitions
tion of angle and orientation (Table 6). of each lattice specimen:

1X N • Test #1: Compression to 15% strain. The data was used to characterise
Ra ¼ jy j ð3Þ the compressive strength, and volumetric energy absorption.
N i¼1 i
• Test #2: Compression to the onset of collapse (approximately 1–2%
strain beyond peak stress). Test results were used to characterise the
where yi is the distance between the local position of the surface and Young's, 1% strain and 2% strain moduli and to obtain images of failure
the nominal surface. mode associated with collapse (Fig. 9).

Rt ¼ maxðyi Þ− minðyi Þ ð4Þ

6.1. Test results


1X
j
Rz ¼ R ð5Þ
j i¼1 ti The uniaxial compression stress-strain responses for the tested 103
cell lattice specimens (corresponding to Test #1, Section 6) are shown
where Rti is each unique value of Rt samples. in Fig. 9. Fig. 10 shows the corresponding volumetric energy absorption

Fig. 12. Mechanical properties of tested lattice specimens. (i) Absolute compressive strength and modulus (ii) specific compressive strength and specific modulus.
352 M. Leary et al. / Materials and Design 98 (2016) 344–357

Fig. 13. Failure summary of tested lattice specimens, including: initial state, 2% strain, 4% strain, prior to collapse, collapse and progressed failure.

Fig. 14. Detail of observed deformation and failure modes of tested lattice specimens at collapse.
M. Leary et al. / Materials and Design 98 (2016) 344–357 353

Fig. 15. Relative modulus and relative compressive strength versus relative density for tested lattice cell specimens, including typical trends for bending and stretch dominated cellular
structures. (i) Relative Young's modulus (ii) relative modulus (1% strain) (ii) relative modulus (2% strain) (iv) Relative compressive strength.

(W) of the lattice specimens, defined as the energy absorption per lat- 6.2. Numerical analysis
tice specimen solid material volume (Eq. (6)). The energy absorption
is plotted up to a strain integration limit, ϵs, of 10% as well the strain cor- Numerical analysis of μCT data was completed to provide insight
responding to the compressive strength achieved by the test specimen into: the fundamental deformation and failure mechanisms; associated
(Fig. 8). mechanical property trends; and, to allow comparison with observed
experimental data.
Where :  
Cell boundary conditions vary significantly depending on boundary
W : Volumetric energy absorption MJ=m3 conditions imposed on the external lattice boundary as well as cell po-
ϵs   ϵ : Strain ½m=m sition within the bulk cellular array. For example, a unit cell within a

ρ
ρs ∫ 0
σdϵ
MJ
m3
ϵs : Strain integration limit ½m=m
σ : Stress ½MPa
ρ
: Lattice specimen relative density ½%
ρs
ð6Þ

Fig. 11 shows a representative stress-strain curve obtained from a


Test #2 (Section 6) BCCZ specimen which identifies Young's Modulus,
and the average of loading and unloading moduli at 1% and 2% strain.
The response is typical of the other tested lattice specimens. Fig. 12 sum-
marises compressive strength and modulus results (including specific
values) for all tested lattice specimens; associated failure modes are
photographically documented in Figs. 13 and 14.
Fig. 15 (i–iii) compares the relative Young's, 1% and 2% strain moduli

ðE Es Þ against the theoretical relative density ðρρ Þ for the tested lattice
s

specimens. The relative compressive strength ðσ =σ s Þ is compared


against the measured relative density ðρρ Þ in Fig. 15 (iv). Expected theo-
s

retical trends are presented for bending and stretch dominated cellular Fig. 16. Relative compressive stress versus relative density for experimental (EXP) and
structures in which mechanical properties show a positive power rela- numerical (FEA) results. Numerical results include constrained (CON.) and unconstrained
tionship with the relative density [75]. (UNCON.) boundary conditions.
354 M. Leary et al. / Materials and Design 98 (2016) 344–357

bulk cellular array will be constrained by interaction with neighbouring exhibited by FCC specimens (Fig. 12). These results align with the ex-
cells, while a unit cell on an lattice boundary may be unconstrained de- pected high structural compliance of the BCC and FCC lattice topolo-
pending on its specific location [89]. To gain insight into the feasible cell gies due to their: under-stiff behaviour, lack of struts aligned in
response, the following extreme in boundary conditions were assessed loading direction and low relative density (Table 4).
individually: • The low compliance of BCC and FCC specimens corresponds to a stress
concentration at the strut node intersection in the associated numer-
• Constrained boundary condition, whereby nodes on the lateral lat-
ical results (Table 7). These specimens are not capable of absorbing
tice boundary are constrained to disallow lateral translation. The
useful amounts of energy during deformation (Fig. 8). Of the highly
constrained scenario represents lattice cells fully within bulk lat-
compliant lattice topologies, FCC specimens show notably higher spe-
tice structures as well as external lattice structures that are subject
cific compressive strength and modulus values compared to BCC spec-
to rigid external constraints.
imens, indicating superior structural efficiency, however with lower
• Unconstrained boundary condition, whereby nodes on the lateral
overall ductility (Fig. 9).
lattice boundary are free to translate. The unconstrained scenario
• The highest overall compressive strength and observed modulus was
represents lattice cells on the external lattice surface that are not
achieved by FBCCZ specimens (Fig. 12), attributable to their increased
subject to external constraints.
strut connectivity, presence of Z-axis struts aligned with loading di-
rection and correspondingly high density. Numerical analysis indi-
For both cases, linear-elastic analysis was completed with solid ele- cates that peak stresses occur in the Z-struts in the FBCCZ lattice;
ments loaded to a strain of 5% of total cell size, based on μCT data however, it is apparent that the body, and face centred elements with-
(Section 5.1); VonMises stress and relative stiffness, kR, are shown in the FBCCZ lattice assist in supporting the Z-struts as the associated
(Table 7) normalised to results for BCC, which displays the lowest stiff- stress concentration is lower than for either FCZ or BCZ specimens
ness of all assessed topologies. Numerical and experimental data are (Table 7). Of all experimental results, FBCCZ is the only lattice struc-
compared in terms of relative compressive strength and relative density ture that displays a crushing failure mode – all other specimens fail
in Fig. 16. by shearing of lattice cells (Fig. 11).
The numerical analysis results indicate that: • When observing specific strength and moduli values (Fig. 11 (ii)),
FCCZ specimens show the highest specific strength and modulus rela-
• Higher stiffness is associated with lattices including vertical struts tive to other specimens, despite having lower absolute values com-
aligned with the loading direction confirming the associated high re- pared to FBCCZ specimens (Fig. 12 (i)). This outcome indicates that
sistance to compressive loads observed in Section 6. FCCZ provides superior performance for compressive load scenarios
• Stress concentrations are present at the intersection of diagonal struts aimed at maximizing strength- and stiffness-to-weight ratio. Similar
due to the presence of small fillet radii. Increased filleting may im- performance has been noted for titanium lattice structures [89].
prove strength by reducing stress concentrations, particularly in lat- • The ductility of the tested specimens was approximately 6–8% and all
tice structures without vertical struts where stress concentrations specimens showed a large ratio of collapse stress to the plateau stress
are particularly noticeable due to pronounced bending at intersecting (average stress post initial peak) where high initial strength is follow-
strut nodes. ed by drastic collapse after reaching maximum compressive strength
• Specimens with constrained boundary constraints exhibit expectedly (Fig. 7).
higher stiffness values than unconstrained specimens. The results for • Young's modulus was found to be always lower than the Moduli at 1%
constrained boundary conditions are in general closer to the experi- and 2% strain (Fig. 12) indicating that localised plasticity is occurring
mental results, suggesting the constrained boundary conditions are in the lattice specimens at stresses below the compressive strength.
more representative of experimental conditions. Similar observations have been made for metallic foams, for example
• The loading response trends in Fig. 16 generally align with theoretically [75], and recently in titanium alloy SLM lattice structures [89]. This out-
expected trends for stretch and bending dominated cellular structures come is particularly important as a number of publications relevant to
in which compressive strength increases with relative density [75]. additively manufactured lattices associate the Young's modulus with
the expected in-situ lattice stiffness, for example [92–94] despite the
observation demonstrated in this work that the Young's modulus may
6.3. Discussion of experimental results underestimate stiffness for repeated elastic loading of lattice structures.
• Of the investigated lattice, the volumetric energy absorption is highest
The following characteristics are observed for the experimental and for FBCCZ specimens (Fig. 10). However, FBCCZ is also associated with
numerical lattice specimen test results: high modulus and compressive strength (Fig. 12) and if applied to ab-
• BCC specimens exhibit the lowest absolute compressive strength and sorb crushing stress, the force transferred to the protected structure
Moduli of all candidate lattice cell structures. Similarly low values are by the lattice will also be the highest of the tested specimens.
• The response of tested lattice specimens exhibits trends in: relative

Table 3
Table 4
Manufacturability of cantilever strut inclination specimens.
Cell topology, Maxwell number, and load-aligned struts for candidate cubic unit cell lattice
Ranking indices: red: failure, green: manufacturable.
structures.

Strut diameter (mm) BCC BCCZ FCC FCCZ FBCCZ


Strut angle
(degrees)
0.5 1.0 2.0 3.0 Cell type

0
Struts, s 8 12 16 20 28
35.3 Struts, s 9 9 12 12 13
Maxwell
45 number, M -13 -9 -14 -10 -5
Strut aligned to
90 NO YES NO YES YES
load direction
M. Leary et al. / Materials and Design 98 (2016) 344–357 355

Table 5 Table 7
Test specimen geometric parameters. Numerical analysis of candidate lattice structures. Image shows Von Mises stress and rel-
ative stiffness, kR, both normalised to data for unconstrained BCC.
Cell Unit cell Specimen Cell Strut Number
Topology theoretical relative theoretical relative Size diameter of cells Unconstrained Constrained
density, ρc [%] density, ρ [%] [mm] [mm]

BCC 8.5% 8.5% 7.5 1.0 103


BCCZ 10.8% 9.7% BCC
FCC 11.6% 7.5%
FCCZ 14.7% 8.7%
FBCCZ 20.6% 16.1%

density, relative modulus and compressive strength which are associat-


ed with the presence of vertical Z-struts aligned with the compressive kR 1.0 3.5
loading direction (Fig. 15). Lattice topologies which include vertical
struts (BCCZ, FCCZ, and FBCCZ) correspond with the expected behav- BCZ
iour of stretch dominated structures (suggesting that the topologies
are more subject to stretch dominated deformation mechanisms),
whereas lattice topologies without vertical axis struts (BCC, FCC) do
not adhere to this trend.

kR 8.1 11.1

7. Concluding remarks FBCZ

Although some researchers have reported difficulty in processing


SLM aluminium, the processing variables optimised for AlSi12Mg
allowed robust manufacture of structural lattice. Limitations of manu-
facturability were documented, including minimum strut diameter
and allowable inclination angle for the proposed processing parameters. kR 9.6 15.3
Based on these identified processing limits, a number of relevant lattice
structures were defined, including structures with low Maxwell num- FCZ
ber (intending to provide high compliance) as well as structures with
higher Maxwell number and struts aligned in the loading direction
(these lattice elements provide high energy absorption and stiffness).
Sample specimens were successfully manufactured for all identified
lattice structures - the specimens displayed ductility of approximately
6–8%, and all specimens display classic lattice structural behaviour, in- kR 8.0 10.9
cluding a large ratio of collapse stress to plateau stress (Fig. 7). Based
on this work, the manufacturability of SLM AlSi12Mg lattice structures
FC
was theoretically and experimentally assessed for varying cell topology,
including surface morphology and roughness; failure mechanisms and
energy absorption characteristics. These novel outcomes assist to enable
commercial application of SLM AlSi12Mg lattice structures.
In particular, the energy absorption capacity of the lattice structures
of interest displayed considerable variation. The associated data allows kR 1.5 6.2
the prediction of energy absorption forces of structural AlSi12Mg lattice
structures – in particular for the FBCCZ structure, which was
demonstrated to provide stable crushing behaviour and excellent ener-
gy absorption characteristics.
Table 6 Experimental observation indicates that Young's Modulus is always
Average, Ra, and peak-to-peak, Rz, roughness of strut elements (Fig. 8). SD denotes stan- lower than the Moduli at 1% and 2% strain. This effect is due to local plas-
dard deviation. Measurements in μm.
ticity not observed in bulk structures, and although typically reported
Build angle for metallic foams, does not appear to be commonly reported for AM
(degrees) structures. This observation is therefore particularly important for AM
35.26 45 90 lattice designers and researchers as the reported Young's modulus
Ra lower 120.2 43.5 –
may underestimate stiffness for actual in-service loading.
SD 39.7 20.6 – Although all tested lattices are classified as bending dominated
Ra sides 8.1 9. 2 8.2 (Table 4) according to Maxwell's stability criteria (Section 3), the results
SD 3.7 1.9 3.0 highlight that the criteria is a necessary but not sufficient condition for
Ra upper 8.8 8.9 –
predicting mechanical response as the significant effect of loading direc-
SD 2.6 1.5 –
Rz lower 456.5 235.5 – tion on cell rigidity is not directly accommodated by the Maxwell criteria.
SD 113.6 118.2 –
Rz sides 56.2 57.2 48.9
SD 22.1 15.3 11.8 References
Rz upper 53.0 60.5 –
[1] Y. Zhai, D.A. Lados, J.L. LaGoy, Additive manufacturing: making imagination the
SD 10.3 16.0 –
major limitation, JOM 66 (5) (2014) 808–816.
356 M. Leary et al. / Materials and Design 98 (2016) 344–357

[2] ASTM standard F2792-12a, F2792-12a, Standard Terminology for Additive [37] C. Vargel, Corrosion of Aluminium, Elsevier, 2004.
Manufacturing Technologies, ASTM International, West Conshohocken, Pa, 2012. [38] E.A. Avallone, T. Baumeister, A. Sadegh, Marks' Standard Handbook For Mechanical
[3] I. Gibson, D.W. Rosen, B. Stucker, Additive Manufacturing Technologies, Springer, Engineers, Mcgraw-Hill, 2006.
New York, 2010. [39] Ishibashi, K., Kazumi Mochizuki, and Yasunobu Kondo, Electric Wire or Cable. 2013:
[4] D.E. Cooper, M. Stanford, K.A. Kibble, G.J. Gibbons, Additive manufacturing for prod- U.S. Patent Application 13/867,275.
uct improvement at Red Bull Technology, Mater. Des. (41) (2012) 226–230. [40] L. Thijs, et al., Fine-structured aluminium products with controllable texture by se-
[5] A. Drizo, J. Pegna, Environmental impacts of rapid prototyping: an overview of re- lective laser melting of pre-alloyed AlSi10Mg powder, Acta Mater. 61 (5) (2013)
search to date, Rapid Prototyp. J. 12 (2) (2006) 64–71. 1809–1819.
[6] G. Guan, et al., Evaluation of selective laser sintering processes by optical coherence [41] L. Facchini, Microstructure and Mechanical Properties of Biomedical Alloys Produced
tomography, Mater. Des. 88 (2015) 837–846. by Rapid Manufacturing Techniques, University of Trento, 2010.
[7] R. Gibson, Stucker, Additive Manufacturing Technologies — Rapid Prototyping to Di- [42] J.P. Kruth, P. Mercelis, J. Van Vaerenbergh, T. Craeghs, Feedback Control of Selective
rect Digital Manufacturing2010. Laser Melting. In 3rd International Conference on Advanced Research in Virtual and
[8] M. McMillan, et al., Programmatic lattice generation for additive manufacture, Proc. Rapid Prototyping, 2007 (Leiria, Portugal).
Technol. 20 (2015) 178–184. [43] E. Foroozmehr, R. Kovacevic, Effect of path planning on the laser powder deposition
[9] J.D. Hiller, H. Lipson, STL 2.0: a Proposal for a Universal Multi-Material Additive process: thermal and structural evaluation, Int. J. Adv. Manuf. Technol. 51 (5–8)
Manufacturing File Format, In Proceedings of the Solid Freeform Fabrication Sympo- (2010) 659–669.
sium, 2009. [44] N. Karapatis, et al., Thermal Behavior of Parts Made by Direct Metal Laser Sintering, 9
[10] W.E. Frazier, Metal additive manufacturing: a review, J. Mater. Eng. Perform. 23 (6) Th SFF Symposium, Austin, 1998.
(2014) 1917–1928. [45] T. Craeghs, et al., Detection of process failures in layerwise laser melting with optical
[11] R. Paul, Modeling and Optimization of Powder Based Additive Manufacturing (AM) process monitoring, Phys. Procedia 39 (2012) 753–759.
Processes, University of Cincinnati, 2013. [46] B.A. Szost, et al., A comparative study of additive manufacturing techniques: Resid-
[12] J.P. Kruth, L. Froyen, J. Van Vaerenbergh, P. Mercelis, M. Rombouts, B. Lauwers, Selec- ual stress and microstructural analysis of CLAD and WAAM printed Ti–6Al–4 V com-
tive laser melting of iron-based powder, J. Mater. Process. Technol. 149 (1) (2004) ponents, Mater. Des. 89 (2016) 559–567.
616–622. [47] M.R. Alkahari, T. Furumoto, T. Ueda, A. Hosokawa, R. Tanaka, A. Aziz, M. Sanusi, Ther-
[13] C. Chu, G. Graf, D.W. Rosen, Design for additive manufacturing of cellular structures, mal conductivity of metal powder and consolidated material fabricated via selective
Comput.-Aided Des. Applic. 5 (5) (2008) 686–696. laser melting, Key Eng. Mater. 523 (2012) 244–249.
[14] M. Leary, L. Merli, F. Torti, M. Mazur, M. Brandt, Optimal topology for additive man- [48] A.V. Gusarov, et al., Contact thermal conductivity of a powder bed in selective laser
ufacture: a method for enabling additive manufacture of support-free optimal struc- sintering, Int. J. Heat Mass Transf. 46 (6) (2003) 1103–1109.
tures, Mater. Des. 63 (2014) 678–690. [49] R.O.a.E. C., Rapid Manufacturing of Lattice Structures with Selective Laser Melting,
[15] S. Mellor, L. Hao, D. Zhang, Additive manufacturing: a framework for implementa- Proc. of SPIE 6107, Laser-Based Micropackaging, 2005.
tion, Int. J. Prod. Econ. 149 (2014) 194–201. [50] F.G. Bachmann, et al., Rapid Manufacturing of Lattice Structures with Selective Laser
[16] J. Zhang, et al., Fatigue crack propagation behaviour in wire + arc additive Melting, Lasers and Applications in Science and Engineering, 2006 (6107: p.
manufactured Ti–6Al–4V: effects of microstructure and residual stress, Mater. Des. 61070K-61070K-12).
90 (2016) 551–561. [51] K. Kempen, L. Thijs, E. Yasa, M. Badrossamay, W. Verheecke, J.P. Kruth, Process Op-
[17] H. Gong, et al., Influence of defects on mechanical properties of Ti–6Al–4V compo- timization and Microstructural Analysis for Selective Laser Melting of AlSi10Mg. In
nents produced by selective laser melting and electron beam melting, Mater. Des. Solid Freeform Fabrication Symposium, 2011.
86 (2015) 545–554. [52] E. Louvis, P. Fox, C.J. Sutcliffe, Selective laser melting of aluminium components, J.
[18] P. Li, et al., Critical assessment of the fatigue performance of additively Mater. Process. Technol. 211 (2) (2011) 275–284.
manufactured Ti–6Al–4V and perspective for future research, Int. J. Fatigue 85 [53] E.O. Olakanmi, R.F. Cochrane, K.W. Dalgarno, Densification mechanism and micro-
(2016) 130–143. structural evolution in selective laser sintering of Al–12Si powders, J. Mater. Process.
[19] T. Tancogne-Dejean, et al., Probabilistic fracture of Ti–6Al–4V made through additive Technol. 211 (1) (2011) 113–121.
layer manufacturing, Int. J. Plast. 78 (2016) 145–172. [54] S. Dadbakhsh, L. Hao, Effect of Al alloys on selective laser melting behaviour and mi-
[20] A.J. Sterling, et al., Fatigue behavior and failure mechanisms of direct laser deposited crostructure of in situ formed particle reinforced composites, J. Alloys Compd. 541
Ti–6Al–4V, Mater. Sci. Eng. A 655 (2016) 100–112. (2012) 328–334.
[21] Q.C. Liu, J. Elambasseril, S.J. Sun, M. Leary, M. Brandt, P.K. Sharp, The effect of [55] S. Dadbakhsh, L. Hao, Effect of hot isostatic pressing (HIP) on Al composite parts
manufacturing defects on the fatigue behaviour of Ti–6Al–4V specimens fabricated made from laser consolidated Al/Fe2O3 powder mixtures, J. Mater. Process. Technol.
using selective laser melting, Adv. Mater. Res. 891 (2014) 1519–1524. 212 (11) (2012) 2474–2483.
[22] Y.Y. Sun, et al., The influence of As-built surface conditions on mechanical properties [56] D. Buchbinder, et al., High power selective laser melting (HP SLM) of aluminum
of Ti–6Al–4V additively manufactured by selective electron beam melting, JOM parts, Phys. Procedia 12 (2011) 271–278 (Part A(0)).
(2016) 1–8. [57] E. Brandl, et al., Additive manufactured AlSi10Mg samples using Selective Laser
[23] P.A. Kobryn, S.L. Semiatin, The laser additive manufacture of Ti–6Al–4V, JOM 53 (9) Melting (SLM): Microstructure, high cycle fatigue, and fracture behavior, Mater.
(2001) 40–42. Des. 34 (2012) 159–169.
[24] J. Imanishi, P.F. Choong, Three-dimensional printed calcaneal prosthesis following [58] K. Kempen, et al., Mechanical properties of AlSi10Mg produced by selective laser
total calcanectomy, Int. J. Surg. Case Rep. 10 (83–87) (2015). melting, Phys. Procedia 39 (2012) 439–446.
[25] X. Wang, et al., Topological design and additive manufacturing of porous metals for [59] S.K. Ghosh, P. Saha, Crack and wear behavior of SiC particulate reinforced aluminium
bone scaffolds and orthopaedic implants: a review, Biomaterials 83 (2016) based metal matrix composite fabricated by direct metal laser sintering process,
127–141. Mater. Des. 32 (1) (2011) 139–145.
[26] J. Parthasarathy, B. Starly, S. Raman, A design for the additive manufacture of func- [60] M. Wong, et al., Selective laser melting of heat transfer devices, Rapid Prototyp. J. 13
tionally graded porous structures with tailored mechanical properties for biomedi- (5) (2007) 291–297.
cal applications, J. Manuf. Process. 13 (2) (2011) 160–170. [61] M. Ameli, et al., A novel method for manufacturing sintered aluminium heat pipes
[27] M. Brandt, S.S., M. Leary, S. Feih, J. Elambasseril, Q. Liu, High-Value SLM Aerospace (SAHP), Appl. Therm. Eng. 52 (2) (2013) 498–504.
Components: From Design to Manufacture, Adv Mater Res, Trans Tech Publications, [62] X.P. Li, et al., A selective laser melting and solution heat treatment refined Al–12Si
Switzerland, Vol. 633 2013, pp. 135–147. alloy with a controllable ultrafine eutectic microstructure and 25% tensile ductility,
[28] J.M. Waller, B.H. Parker, K.L. Hodges, E.R. Burke, J.L. Walker, E.R. Generazio, (), Non- Acta Mater. 95 (2015) 74–82.
destructive Evaluation of Additive Manufacturing, NASA, 2014. [63] K. Bartkowiak, et al., New developments of laser processing aluminium alloys via
[29] D. Cooper, J. Thornby, N. Blundell, R. Henrys, M.A. Williams, G. Gibbons, Design and additive manufacturing technique, Phys. Procedia 12 (2011) 393–401.
manufacture of high performance hollow engine valves by Additive Layer [64] A.V. Gusarov, J.P. Kruth, Modelling of radiation transfer in metallic powders at laser
Manufacturing, Mater. Des. 69 (2015) 44–55. treatment, Int. J. Heat Mass Transf. 48 (16) (2005) 3423–3434.
[30] M. Leary, in: R.E. Lumley (Ed.), Materials Selection and Substitution Using Alumin- [65] F. Incropera, D. DeWitt, Introduction to Heat Transfer, Wiley, 1985.
ium Alloys, in Fundamentals of Aluminium Metallurgy: Production, Processing and [66] P. Fischer, et al., A model for the interaction of near-infrared laser pulses with metal
Applications, Woodhead publishing, 2011. powders in selective laser sintering, Appl. Phys. A Mater. Sci. Process. 74 (4) (2002)
[31] M.F. Ashby, Multi-objective optimization in material design and selection, Acta 467–474.
Mater. 48 (1) (2000) 359–369. [67] A. Rai, et al., Understanding the mechanism of aluminium nanoparticle oxidation,
[32] E. Masoumi Khali Abad, S. Arabnejad Khanoki, D. Pasini, Fatigue design of lattice ma- Combust. Theor. Model. 10 (5) (2006) 843–859.
terials via computational mechanics: application to lattices with smooth transitions [68] Z.A. Munir, Analytical treatment of the role of surface oxide layers in the sintering of
in cell geometry, Int. J. Fatigue 47 (2013) 126–136. metals, J. Mater. Sci. 14 (11) (1979) 2733–2740.
[33] A. Foroozmehr, M. Badrossamay, E. Foroozmehr, Finite element simulation of selec- [69] J. Campbell, Castings, Butterworth-Heinemann, 2003.
tive laser melting process considering optical penetration depth of laser in powder [70] Standardization, T.I.O.F., ISO 13320:2009 Particle Size Analysis – Laser Diffraction
bed, Mater. Des. 89 (2016) 255–263. Methods, 2009.
[34] K.J. Maloney, et al., Multifunctional heat exchangers derived from three-dimensional [71] Standardization, T.I.O.F., ISO 9276–2:2014 Representation of Results of Particle Size
micro-lattice structures, Int. J. Heat Mass Transf. 55 (9–10) (2012) 2486–2493. Analysis – Part 2: Calculation of Average Particle Sizes/Diameters and Moments
[35] C.S. Roper, Multiobjective optimization for design of multifunctional sandwich panel from Particle Size Distributions, 2014.
heat pipes with micro-architected truss cores, Int. J. Heat Fluid Flow 32 (1) (2011) [72] ASTM, E8/E8M - 11 standard test methods for tension testing of metallic materials.
239–248. [73] W. Hesse, Aluminium-Schlüssel, Beuth Verlag, 2012.
[36] J.R. Davis, in: J.R. Davis, A.S.M. international (Eds.),Aluminum and Aluminum Alloys, [74] K. Kempen, T.L., J. Van Humbeeck, J.P. Kruth, Mechanical properties of AlSi10Mg pro-
1993. duced by selective laser melting, Phys. Procedia 39 (2012) 439–446.
M. Leary et al. / Materials and Design 98 (2016) 344–357 357

[75] L.J. Gibson, M.F. Ashby, Cellular Solids: Structure and Properties, Cambridge univer- [87] G. Strano, et al., Surface roughness analysis, modelling and prediction in selective
sity press, 1997. laser melting, J. Mater. Process. Technol. 213 (4) (2013) 589–597.
[76] M.F. Ashby, D. Cebon, Materials selection in mechanical design, J. Phys. IV 3 (C7) [88] G. Pyka, et al., Surface roughness and morphology customization of additive
(1993) (p. C7-1). manufactured open porous Ti6Al4V structures, Materials 6 (10) (2013) 4737–4757.
[77] S. Xu, et al., Design of lattice structures with controlled anisotropy, Mater. Des. 93 [89] M. Mazur, M. Leary, S. Sun, D. Shidid, M. Brandt, Deformation and Failure Behaviour
(2016) 443–447. of Ti–6Al–4V Structural Lattices Processed by Selective Laser Melting (SLM)Journal
[78] V.S. Deshpande, N.A. Fleck, M.F. Ashby, Effective properties of the octet-truss lattice of Advanced Manufacturing Technologies 2015.
material, J. Mech. Phys. Solids 49 (8) (2001) 1747–1769. [91] ASTM Standard E9–09, E9–09 standard test method compression testing of metallic
[79] V.S. Deshpande, M.F. Ashby, N.A. Fleck, Foam topology: bending versus stretching materials at room temperature. ASTM International: West Conshohocken, Pa.
dominated architectures, Acta Mater. 49 (06) (2001) 1035–1040. [92] J. Wieding, A. Wolf, R. Bader, Numerical optimization of open-porous bone scaffold
[80] D.L. Schodek, Structures, Prentice-Hall, Englewood Cliffs, NJ, 1980. structures to match the elastic properties of human cortical bone, J. Mech. Behav.
[81] M.F. Ashby, et al., Metal Foams: a Design Guide: a Design Guide, Elsevier, 2000. Biomed. Mater. 37 (2014) 56–68.
[82] M. Leary, et al., Feasible Build Orientations for Self-Supporting Fused Deposi- [93] O. Cansizoglu, et al., Properties of Ti–6Al–4V non-stochastic lattice structures fabri-
tion Manufacture: a Novel Approach to Space-Filling Tesselated Geometries, cated via electron beam melting, Mater. Sci. Eng. A 492 (1) (2008) 468–474.
Advanced Materials Research, Trans Tech Publ., 2013. [94] C. Yan, et al., Evaluations of cellular lattice structures manufactured using selective
[83] N. Vanderesse, et al., Image analysis characterization of periodic porous materials laser melting, Int. J. Mach. Tools Manuf. 62 (2012) 32–38.
produced by additive manufacturing, Mater. Des. 92 (2016) 767–778.
[84] S. Van Bael, et al., Micro-CT-based improvement of geometrical and mechanical con-
trollability of selective laser melted Ti6Al4V porous structures, Mater. Sci. Eng. A 528
(24) (2011) 7423–7431.
[85] L. Hao, D. Raymond, Design and Additive Manufacturing of Cellular Lattice Struc-
tures. … on Advanced Research in Virtual and …, 2011 249–254.
[86] A.B. Spierings, M. Schneider, R. Eggenberger, Comparison of density measurement
techniques for additive manufactured metallic parts, Rapid Prototyp. J. 17 (5)
(2011) 380–386.

Anda mungkin juga menyukai