You are on page 1of 18

Materials and Design 183 (2019) 108137

Contents lists available at ScienceDirect

Materials and Design


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

SLM lattice structures: Properties, performance, applications and


challenges
Tobias Maconachie a, b, Martin Leary a, b, c, *, Bill Lozanovski a, c, Xuezhe Zhang a, Ma Qian a,
Omar Faruque b, d, Milan Brandt a, c
a
RMIT Centre for Additive Manufacture, RMIT University, Melbourne, Australia
b
ARC Training Centre for Lightweight Automotive Structures (ATLAS), Australian Research Council Grant IC160100032, Australia
c
ARC Training Centre in Additive Biomanufacturing, Australia
d
Ford Motor Company, Research Innovation Centre (RIC), Dearborn, MI, USA

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 A review of the design, fabrication


and mechanical performance of Se-
lective Laser Melting (SLM) lattice
structures.
 Reported data on the mechanical
performance and properties of a
broad range of SLM lattice structures
is compiled.
 Meta-analysis of reported data allows
insights into the applicability of the
Gibson-Ashby model to SLM lattice
structures.

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

Article history: Additive manufacturing (AM), particularly Selective Laser Melting (SLM) has enabled development of
Received 26 April 2019 lattice structures with unique properties. Through control of various parameters lattice structures can
Received in revised form produce unique mechanical, electrical, thermal and acoustic properties, and have received much research
12 August 2019
attention. Despite the increasing volume of published data on the mechanical response of specific SLM
Accepted 18 August 2019
Available online 19 August 2019
lattice structures, there exists no overarching analysis. This work addresses this identified deficiency by
providing a comprehensive summary of the experimental data reported on the mechanical response of
SLM lattice structures. The design, fabrication and performance of SLM lattice structures are reviewed
Keywords:
Additive manufacturing
and the quality of data reported is analysed to inform best-practice for future studies. This compre-
SLM hensive data summary enables meta-analysis of the reported mechanical performance of SLM lattice
Selective laser melting structures, providing insight into the bounds of their technical capabilities. Correlations were identified
Lattice structures between the relative density and mechanical properties of many unit cell topologies consistent with the
DFAM predictions of the Gibson-Ashby model, indicating its usefulness in describing and predicting the
Mechanical properties behaviour of SLM lattice structures. This review provides designers with a compiled resource of exper-
imental data and design for AM tools to inform future design applications of SLM lattice structures and
Data availability statement: facilitates their further commercial adoption.
The raw/processed data required to © 2019 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
reproduce these findings cannot be shared at creativecommons.org/licenses/by-nc-nd/4.0/).
this time due to technical or time limitations.

* Corresponding author at: RMIT Centre for Additive Manufacture, RMIT University, Melbourne, Australia.
E-mail address: martin.leary@rmit.edu.au (M. Leary).

https://doi.org/10.1016/j.matdes.2019.108137
0264-1275/© 2019 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
2 T. Maconachie et al. / Materials and Design 183 (2019) 108137

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Strut-based lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Triply periodic minimal surface lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.3. Shell lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. The SLM process for lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1. Processing parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4. Applications of SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.1. Biomedical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.2. Aerospace . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5. Mechanical properties of SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5.1. Compressive performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5.2. Fatigue performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.3. Effects of processing parameters on SLM lattice structure performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
6. Microstructure of SLM components and effect of heat treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
6.1. Effect of heat treatment on microstructure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
7. Potential geometric defects of SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
7.1. Manufacturability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
7.2. Dimensional accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
7.3. Surface roughness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
7.4. Effects of defects on the mechanical performance of SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
8. Numerical modelling of SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
9. The Gibson-Ashby model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
9.1. Gibson-Ashby model in the literature of SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
9.2. Compressive experimental data for SLM lattice structures reported in the literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
9.3. Comparison of reported experimental data with predictions of the Gibson-Ashby model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
9.4. Regression analysis of reported experimental data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
10. Reporting of experimental data on SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
11. Outstanding challenges and future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
11.1. Accounting for manufacturing defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
11.2. Efficient and precise predictive models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
11.3. Functionally-graded and locally tuned lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
11.4. Fatigue behaviour of SLM lattice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
11.5. Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
12. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
CRediT authorship contribution statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

1. Introduction structures can be used to reduce the stiffness of metallic medical


implants to be closer to that of bone [5] thereby avoiding stress
Additive manufacturing (AM) enables the fabrication of strong, shielding while allowing fluid flow due to their porosity [12] with a
lightweight structures with geometry that is unachievable by large surface area-to-volume ratio to facilitate osseointegration
traditional manufacturing methods, including complex lattice [13]. The high strength-to-weight ratio and thermal conduction
structures [1]. Lattice structures are topologically ordered, three- properties of lattice structures make them attractive for aerospace
dimensional open-celled structures composed of one or more applications [14]. The reliable control of the collapse response of
repeating unit cells [2,3]. These cells are defined by the dimensions lattice structures due to their structural uniformity [15] mean they
and connectivity of their constituent strut elements, which are may provide a technical advantage for energy-absorption when
connected at specific nodes. At the cellular level, lattice structures compared with alternative metallic foams [16].
can be considered to be structures as far as features and properties Selective Laser Melting (SLM) is a Powder Bed Fusion (PBF) form
are concerned, but behave like homogenised meta-materials when of AM whereby metallic components are fabricated from CAD data by
considered at the overall structural level [4]. the layer-wise fusion of metallic powders using a laser energy source
By tuning lattice structural parameters, such as cell topology [17]. For each layer, metallic powder is distributed across the previ-
(connectivity) or geometry (cell size and strut dimensions), the ous layer, which is then selectively scanned by the laser to locally fuse
physical response of these structures can be significantly altered to the powder. PBF is the preferred technique for metallic lattice
exhibit properties unachievable by their parent materials [5], manufacture as alternative metallic AM methods such as direct metal
including acoustic [6], dielectric [7] and mechanical [8] properties. deposition do not provide sufficient geometric accuracy [2].
AM lattice structures have been found to significantly outperform Many studies have reported on the mechanical performance of
cellular structures produced by alternative manufacturing methods particular SLM lattice structures. However, no overarching analysis
with equivalent porosity [9], particularly due to the greater geo- of their broader performance capabilities exists. To overcome this
metric control and predictability provided by AM fabrication [2]. identified deficit, this review provides a comprehensive summary of
The commercial usefulness of these lattice structures has the experimental literature concerning SLM lattice structures and
motivated much research, particularly for biomedical [10] and provides a meta-analysis of the associated mechanical performance.
aerospace [11] applications. For biomedical applications lattice The different means of lattice structure design, the details of the SLM
T. Maconachie et al. / Materials and Design 183 (2019) 108137 3

process for their fabrication and specific applications are detailed. properties of SLM lattice structures.
The mechanical behaviour of SLM lattice structures is described, and Lattice structures can generally be categorised based on their
the microstructure and potential geometric defects resulting from mechanical response as being either bending-dominated or
the fabrication process and their effects on performance are stretch-dominated. Bending-dominated structures experience
explored. Different means of predicting the behaviour of SLM lattice bending moments within their structure and so are compliant,
structures, including numerical modelling and the Gibson-Ashby whereas stretch-dominated structures experience axial loads,
model are discussed. Experimental data on the performance of meaning they are more stiff and strong than bending-dominated
SLM lattice structures was collected from the literature and a meta- structures [12]. A lattice structure’s cell topology defines whether
analysis of this data with comparison to predictions of the Gibson- it will be bending or stretch-dominated. A broad range of cell to-
Ashby model is presented. The quality of data reported regarding pologies have been investigated in the literature [3] which here are
SLM lattice structures is analysed. Finally, the technical and eco- further categorised as being either strut-based or triply periodic
nomic challenges facing the further implementation of SLM lattice minimal surfaces (TPMS).
structures are presented. This research provides a comprehensive
Design for AM (DFAM) resource for engineers seeking to design SLM 2.1. Strut-based lattice structures
lattice structures for technical applications as well as a roadmap for
further commercially relevant research activities. The most common strut-based cell topologies that have been
investigated are body-centred cubic (BCC) and face-centred-cubic
2. Lattice structures (FCC), or variations of these, such as the inclusion of z-struts
(BCCZ and FCCZ) (Fig. 1) [15], which are named after analogous
Natural cellular materials such as wood, cork and bone have crystalline structures. Other strut-based topologies also exist, such
been utilised for centuries, and their structure is mimicked in as the cubic, octet-truss and diamond [4].
modern technical materials such as manufactured honeycombs and These strut-based topologies are often chosen for their
foams [18]. Cellular structures are an attractive option for many simplicity of design [42], but strut-based topologies have also been
design applications, particularly light-weighting, due to the high generated from topological optimisation to maximise the efficiency
specific strength and stiffness provided by their porous structure of material distribution within the lattice structure and fully
[19]. The deformation behaviour of cellular structures also means embrace the opportunities presented by AM [42,43] (Fig. 2).
they are useful for energy absorption applications [20]. There are Strut-based topologies can be characterised by their Maxwell
many different types of manufactured cellular structures, and a number, M, which is dependent on the number of struts, s, and
variety of means of fabricating them. nodes, n (Eq. (1)) [44].
Metallic foams are a form of cellular structure that are manu-
factured by injecting gas or mixing a foaming agent into molten M ¼ s  3n þ 6 (1)
metal [21] and many studies have investigated their use for technical If M < 0, there are too few struts to equilibrate external forces
applications such as energy absorption [22e24] and light-weighting without equilibrating moments induced at the nodes, causing bending
[25e27]. While metallic foams are relatively inexpensive to manu- stresses to develop in struts and leading to bending-dominated
facture, they consist of a stochastic arrangement of open or closed behaviour. Whereas if M  0, external loads are equilibrated by axial
cells rather than a prescribed cellular arrangement. This stochastic tension and compression in struts meaning that no bending occurs at
structure results in an inconsistent mechanical response that is un- nodes, making these structures stretch-dominated [45]. Due to these
desirable as it necessitates overly conservative design [28]. phenomena, stretch-dominated structures are stiff and strong, espe-
Lattice structures are another type of cellular material that is cially considering their mass, whereas bending-dominated structures
differentiated from foams by the regular repeating structure of are compliant and deform more consistently [31].
their unit cells [29]. Gibson defines cellular materials as consisting
of “an interconnected network of struts or plates” [30]. Ashby
2.2. Triply periodic minimal surface lattice structures
further states that lattice structures, a form of cellular material,
differ from large scale engineered structures such as trusses or
Lattice structures with unit cells based on triply periodic minimal
frames in terms of their scale e the unit cells of a lattice structure
surfaces (TPMS) such as the Schoen gyroid, Schwartz diamond and
have a millimetre or micrometre scale [31]. This means that while
Neovius (Fig. 3) have also been investigated. These topologies are
the unit cells of lattice structures can be analysed as space frames
generated using mathematical formulae that define the U ¼ 0 iso-
using classical mechanics, a lattice structure should be considered a
surface boundary between solid and void sections of the structure
material with its own mechanical properties, which allows direct
(Table 1) [46]. Various parameters such as periodicity and relative
comparison between the properties of a lattice structure and those
density can be altered to tune their mechanical performance [47].
of its parent material.
Periodicities are defined by the k values (kx, ky, kz) which are
Lattice structures may be 2.5D1 [32] or 3D and can be fabricated
calculated using Eq. (2), where ni is the number of cell repetitions in
by a variety of means, including investment casting [33], a combi-
the x, y or z directions and Li is the absolute size of the structure in
nation of extrusion and electro discharge machining [34], or
that direction. The variable t can be used to alter relative density.
various composite fabrication methods including textile weaving
[35], interlacing, interlocking, hot-press [36] or filament winding ni
[37]. However, since the proliferation of AM in the late 1990s and ki ¼ 2p ðwithi ¼ x; y; zÞ (2)
Li
early 2000s [38], much research attention has been paid to the AM
fabrication of lattice structures [39]. The mechanical properties of TPMS lattice structures have potential advantages over strut-
SLM lattice structures were first reported in 2005 [40], and since, based topologies in terms of manufacturability and bone fixation.
due to the significant advantages of SLM for the fabrication of SLM constraints on inclination angle for strut-based lattice struc-
metallic lattice structures [41], many studies have reported on the tures may be avoided as the inclination angle of cell walls contin-
uously varies in TPMS structures, meaning previous layers support
subsequent layers, improving manufacturability [47]. Studies also
1
2.5D lattice structures consist of a two-dimensional shape extruded into the suggest the curvature of implant surfaces plays a critical role in
third dimension. promoting bone ingrowth [48], and due to both their geometric
4 T. Maconachie et al. / Materials and Design 183 (2019) 108137

Fig. 1. Strut-based lattice structures: BCC (A), BCCZ (B), FCC (C), FCCZ (D), cubic (F), Octet-truss (G), and diamond (H).

similarity to natural trabecular bone as well as the three- elastic properties than strut-based open-cell structures with
dimensional curvature of their surfaces, TPMS lattice structures similar densities [51], though manufacture of these structures re-
potentially offer improved osseo-fixation over strut-based lattice mains problematic for powder-based AM systems due to the
structures [5]. Maskery et al. also found that gyroids have almost requirement of powder removal [50]. However, open-celled plate-
three times greater specific energy absorption (SEA) than BCC based lattice structures have been designed, manufactured and
structures with similar porosity [15]. However, these potential tested, and have been shown to exhibit superior strength and
advantages have not been conclusively demonstrated, and the stiffness with very low densities [52].
comparative performance of strut-based and TPMS lattice struc-
tures remains an area for further research.
3. The SLM process for lattice structures

2.3. Shell lattice structures SLM was developed by the Fraunhofer Institute for Laser Tech-
nology, Germany in the 1990s [53] and has since been adopted in
AM has enabled the design and manufacture of cellular struc- various industries due to its ability to produce fully functional and
tures whose unit cells are composed of plates rather than struts near fully dense components directly from metals without the use
[49]. These lattice structures have been described as TPMS-like of intermediate binders or additional processing techniques [54].
(though their surfaces do not necessarily have zero mean curva- During the SLM process, a layer of metallic powder is spread across
ture) and are referred to as “shell lattices” (Fig. 4) [50]. Closed-cell the surface of a build platen. A galvanometer directs a laser beam
plate-based lattice materials have been shown to have superior across this surface to melt powder where necessary, fusing it with

Fig. 2. Topologically optimised strut-based unit cells [42].


Strut connectivity is iteratively optimised based on node locations.
T. Maconachie et al. / Materials and Design 183 (2019) 108137 5

Fig. 3. Triply periodic minimal surface unit cells: Schoen gyroid (A), Schwarz diamond (B) and Neovius (C).

Table 1 reported in the published literature [66]. The SLM processing pa-
Example TPMS formulae for the generation of Schoen gyroids, Schwartz diamonds rameters used in published works reviewed in this research are
and Neovius topologies.
presented in Table 2.
Topology Isosurface formula There are inconsistencies in which processing parameters were
Schoen UG ¼ cos (kxx) sin (kyy) þ cos (kyy) sin (kzz) þ reported in the literature. Some parameters, such as laser power
gyroid cos (kzz) sin (kxx)  t were regularly reported (83%), though power used for both borders
Schwartz UD ¼ sin (kxx) sin (kyy) sin (kzz) þ sin (kxx) cos (kyy) cos (kzz) and hatching was not always clearly differentiated. Laser scan
diamond þ cos (kxx) sin (kyy) cos (kzz) þ cos (kxx) cos (kyy) sin (kzz)  t speed was the least commonly reported parameter (56%). For
Neovius UN ¼ 3[cos(x) þ cos (y) þ cos (z)] þ 4 cos (x) cos (y) cos (z) repeatability of experimental results data should be reported
wherever possible.
the layer below. The platen is then lowered, and this process is
repeated until fabrication is complete [55]. 4. Applications of SLM lattice structures
Advantages of SLM include the ability to produce near fully-
dense, complex components with high resolution [56] and supe- Rapid advances have led to interest in AM for a broad range of
rior properties compared with those produced by traditional applications including personal protective equipment [78], sports
methods such as casting [57]. However, certain disadvantages also equipment [79] and even for the future exploration of Mars [80].
exist such as the minimum feature size that is constrained by the Two industries which have taken particular interest in AM lattice
laser beam spot size [58], microstructural and metallurgical defects structures are the biomedical and aerospace industries [81].
inherited from the fabrication process [59] that can be difficult to
evaluate [60], and the potential for thermal distortion of compo- 4.1. Biomedical
nents due to residual stresses resulting from rapid cooling during
fabrication [61]. The requirements for powder removal are another AM lattice structures are used in the biomedical industry for
limitation of the SLM process [62] as enclosed internal voids cannot medical implants [56], the global market for which is expected to
be produced due to powder entrapment. There is also limited data grow to $116 billion by 2022 [82]. The ability to produce high-
available on the performance of SLM-produced materials [63], quality metallic components that conform to complex, patient-
which in part motivates this study. specific surfaces makes SLM perfect for the fabrication of medical
implants, and the ability to produce metallic components with
3.1. Processing parameters stiffness closer to that of bone makes AM lattice structures perfect
for biomedical applications [5]. These structures can be designed to
SLM processing parameters such as: laser parameters including produce optimal osseointegration and have been shown to sustain
speed power and beam size; powder properties, usually repre- excellent bone in-growth and achieve high performance in terms of
sented by mean particle size; and, layer thickness significantly implant fixation [74].
affect manufacturing outcomes [64]. Optimisation of processing Beyond the manufacture of metallic implants, potential innova-
parameters is key to improving manufacturing outcomes of the tive design applications of AM lattice structures have been sug-
SLM process [65]. gested, such as Burton et al. who have proposed an implant which
Appropriate processing parameters are highly dependent on the incorporates a reservoir that locally releases a therapeutic drug to
machine and material used, and a variety of parameters are achieve antimicrobial functionality by incorporating a reservoir [83].

Fig. 4. Sell lattice unit cells generated by placing plates on the closest-packed planes of cubic crystals combining simple cubic (SC), BCC and FCC: SC-BCC (A), SC-BCC-FCC (B) and SC-
FCC (C) [51].
6 T. Maconachie et al. / Materials and Design 183 (2019) 108137

Table 2
SLM processing parameters for reported data evaluated in this research. Dash (e) indicates an absence of data.

Reference Machine Manufacturer Material Spot size Border power Hatch power Scan speed Hatch spacing Layer thickness Mean powder
(mm) (W)a (W) (mm/s) (mm) (mm) size (mm)

[67] ProX DMP 300 3D Systems Ti-6Al-4V e e e e e e 8.64


[68] ProX-300 3D Systems SS 630 (17- 70 170 170 1600 0.05 40 e
4PH)
[11] Concept X-line Concept Laser AlSi10Mg e e 370 1500 0.19 30 31
1000R Company
[69] M2 Cusing® Concept Laser Ti-6Al-4V 60 150 150 1750 0.075 20 20e50
Company
[70] EOSINT-M270 EOS Ti-6Al-4V 100 58.5 117 225 0.18 30 45
[16] M280 EOS 316L SS 100 e e e e e 20e40
[71] DMLSEOSINT- EOS Ti-6Al-4V 100 170 170 1250 0.06 30 20
M270
[72] M 270 EOS Ti-6Al-4V 100 170 170 1250 0.1 30 29
[72] M 270 EOS Ti-6Al-4V 10 117 117 225 0.18 30 e
[73] M280 EOS AlSi10Mg e 370 370 1500 0.13 30 30
[64] Realizer II MCP 316L SS 90 80e160 80e160 e e e 16e38
[74] MCP Realizer 2, MCP Ti-6Al-4V 54 80 80 e e 50 45
250 SLM
[47] Realizer SLM MCP 316L SS 40 95 95 e 0.075 75 45
Workstation
[5] AM250 Renishaw AlSi10Mg 80 200 200 e 0.13 25 e
[15] AM250 Renishaw Ti-6Al-4V e 100 200 e 0.065 30 e
[75] AM250 Renishaw AlSi10Mg 70 e e e e 25 e
[76] SLM 125 SLM Solutions CP-Ti e 100 100 385 0.12 e 36.6
[77] SLM250HL SLM Solutions Ti-6Al-4V e 100 100 375 0.12 30 40
Number reported 12 14 15 10 13 14 13
Percentage reported (%) 67 78 83 56 72 78 72
a
Where no distinction was reported between border and hatch power, the same value was assumed to be reported for both.

4.2. Aerospace orientation of cells with respect to loading direction also signifi-
cantly affects the mechanical response of anisotropic lattice struc-
The light-weighting potential of AM lattice structures, such as tures [2].
the replacement of internal solid volumes with lattice structures Most mechanical testing of SLM lattice structures is performed
with a similar strength [55], means AM is of great interest to the in compression due to its greater simplicity than testing in tension.
aerospace industry [84]. Other properties of AM lattice structures, This is largely because special design of the sample-test rig inter-
such as the ability to produce conformal cooling channels also face is required for tensile testing, whereas compressive tests can
makes them an attractive option for aerospace applications [14]. be more simply performed by crushing a lattice between plates. For
Zhou et al. have developed a lightweight phase-change thermal example, when testing lattice structures in tension, Alsalla et al.
controller based on lattice cells [85]. Thermal controllers are an found that lattice structures failed near the interface between the
important component to manage the temperature of various elec- lattice structure and test apparatus attachment, suggesting that a
tronics in spacecraft, yet traditional designs considerably add to the critical stress concentration exists at the lattice interface [92]. The
spacecraft’s weight. Using an SLM-fabricated lattice sandwich tensile behaviour of lattice structures remains an area requiring
structure Zhou et al. were able to produce a thermal controller with further research.
50% increased thermal capacity compared to traditional alterna-
tives with similar mass.
5.1. Compressive performance

5. Mechanical properties of SLM lattice structures The general deformation behaviour of lattice structures can be
divided into three discreet stages (Fig. 5): linear elastic deforma-
Various terms such as mechanical properties, elastic modulus tion; plastic deformation; and, densification [86]. The component
and yield strength are applied to lattice structures, though they struts of lattice structures are susceptible to three collapse mech-
have a slightly different meaning than when applied to continuous anisms under compression e yield, buckling and fracture e that
bulk materials [2]. When referring to lattice structures, these compete until the mechanism with the lowest stress threshold is
properties are the “apparent macroscopic properties of structures reached [31].
that converge to certain values when the number of the unit cells is During elastic deformation, the material response is linear
large enough” [2]. Due to the geometric freedom afforded by AM, elastic with a modulus proportional to the structure material’s
lattice structures can achieve functional or mechanical properties elastic modulus. However, for the most accurate representation of
that cannot be achieved by bulk materials [86] such as auxetic the elastic modulus of metallic cellular structures, Ashby et al.
structures with negative Poisson’s ratio [87], negative stiffness [88], recommend measuring the unloading modulus as it better repre-
negative compressibility [89], negative thermal expansion coeffi- sents the structure’s performance [93].
cient [90] or very high stiffness with low mass [91]. Once the elastic limit is reached, plastic deformation begins as
Mechanical properties of lattice structures are usually expressed cells begin to yield or buckle. For bending-dominated structures
as a fraction of the mechanical properties of their parent material deformation continues with an almost constant stress, referred to
[2] and are dependent on the relative density of the lattice structure as the plateau stress, whereas the stress required for further
(r*/rs), which is the ratio of the apparent density of the cellular deformation oscillates in stretch-dominated structures. Once cell
structure (r*) to the density of the cellular structure’s material (rs). components deform enough that contact with other components
Regardless of topology, the mechanical properties of lattice struc- occurs, constraining further deformation, the densification strain is
tures are known to decrease with reduced relative density [69]. The reached, and densification begins as stress steeply increases.
T. Maconachie et al. / Materials and Design 183 (2019) 108137 7

This means that topologies that would technically be defined as


bending-dominated according to the Maxwell criterion, such as
BCCZ and FCCZ, are able to produce stretch-dominated behaviour
depending on the orientation of struts to the load direction [41].
Also, while lattice structures generally behave in the manner
described by Gibson and Ashby [95], as presented in Fig. 5, this may
not be the case for more unusual cellular topologies [96].

5.2. Fatigue performance

The fatigue performance of SLM lattice structures is critical for


many high-value technical applications. For example, both
biomedical and aerospace components are subject to stringent
limits associated with cyclic loading [12].
The fatigue behaviour of lattice structures subject to dynamic
loading can be broken into three stages: in Stage 1 strain increases
rapidly; in Stage 2 cumulative strain remains approximately con-
Fig. 5. General compressive behaviour of stretch and bending-dominated lattice
stant for around 104 to 106 cycles; and, in Stage 3 cumulative strain
structures during elastic deformation, plastic deformation and densification stages.
increases exponentially, eventuating in the rapid failure of the
During this process there are three compressive failure modes: specimen [12].
successive cell collapse; propagation of cracks through the lattice; The factors that most significantly affect the fatigue properties
and, diagonal shear [15]. Cells collapse due to buckling or crushing. of lattice structures are: the bulk material’s mechanical properties;
Crack propagation usually originates from a pre-existing defect, the relative density of the lattice; cell topology; and, the geometry
and diagonal shear results in an initial loss of around 50% strength, of the cell’s struts, which is defined by the distribution of material
followed by somewhat uniform strengthening during densification. within the structure [12]. Lattice fatigue life has been found to
Compressive strength decreases with increased unit cell size consistently increase with increased relative density, but cell to-
when porosity is kept constant, due to the greater tendency for pology significantly affects fatigue results. Cyclic ratchetting, which
buckling as strut slenderness increases [45]. Compressive strength refers to the progressive accumulation of strain during Stage 1 [97]
and stiffness are highly dependent on whether the unit cell exhibits has been found to be dependent on cell topology and local geom-
bending-dominated or stretch-dominated behaviour e bending- etry [3]. Yavari et al. found that while lattice structures with certain
dominated structures are compliant whereas stretch-dominated topologies (diamond and truncated cuboctahedron) had a fatigue
structures have greater strength [94]. Selected experimental data life between 1.5  105 and 2  105 cycles, lattice structures with
from the literature [45] on the compressive behaviour of certain cell cube unit cells did not fail after 106 cycles (the limit of their tests)
topologies which demonstrates this is presented in Fig. 6. [4].
BCC and FCC cell topologies behave as bending-dominated Both relative density and topology affect lattice structure fatigue
structures, with an initial stage of elastic deformation followed by performance. The fatigue behaviour of SLM lattice structures with
continuous deformation at an approximately constant plateau certain topologies and relative densities are compared in Fig. 7.
stress. However, when topologies are reinforced with z-struts Lattice structures with greater relative density could sustain greater
(BCCZ, FCCZ), the structures exhibit stretch-dominated behaviour e loads for a larger number of cycles than those with lower relative
they are stronger and stiffer, requiring greater loads to cause density (Fig. 7A), but truncated cuboctahedron lattice structures
yielding and deformation, and the post-yield deformation occurs could sustain greater loads for similar periods than diamond lattice
with an oscillating stress [86]. structures with similar relative densities (Fig. 7B).
Tensile loads cause bending stresses to develop in struts in

Fig. 6. Stress-strain relationships for different strut-based cell topologies [45]. Inclusion of z-struts in BCC and FCC lattice structures changes behaviour from bending-dominated to
stretch-dominated.
8 T. Maconachie et al. / Materials and Design 183 (2019) 108137

Fig. 7. A: Comparison of fatigue behaviour of SLM lattice structures with diamond (D) and truncated cuboctahedron (C) topologies and varying relative densities (%) [4]. B: Failure
stress after 50  103 cycles for different topologies and relative densities.

bending-dominated structures. These stresses can initiate cracks dislocation movement [76], but may also result in segregation
within the struts that propagate until final fracture. As more struts phenomena or non-equilibrium phases [105].
fail the load bearing capacity and stiffness of the fatigue-loaded High temperature gradients result in a hexagonally-packed acic-
structure decreases until complete catastrophic failure [12]. ular martensitic phase microstructure, a0 , for Tie6Ale4V [105,106].
Manufacturing defects such as the staircase effect can significantly Partial melting of previous layers during fabrication leads to elon-
affect the fatigue properties of lattice structures [12] and reducing gation of grains, which can improve mechanical properties such as
porous defects in struts by optimising processing parameters can strength [107], the direction of which is determined by laser scanning
significantly improve lattice fatigue performance [2]. path and component geometry [105]. Solidification occurs primarily
Although it is clear that both the relative density and cell to- through epitaxial growth in the direction of the heat flux [98].
pology of lattice structures affect their fatigue performance, the The microstructure of SLM lattice structures results from the
broader fatigue behaviour of SLM lattice structures remains an SLM fabrication process, but microstructure can be further altered
unexplored field [98]. with heat treatment.

5.3. Effects of processing parameters on SLM lattice structure 6.1. Effect of heat treatment on microstructure
performance
Heat treatments provide a viable means of improving the me-
As with any SLM-fabricated component, the selection of chanical properties of lattice structures [12] in various ways,
appropriate SLM processing parameters is vitally important for including the closure of internal pores, improvement of local fusion
producing lattice structures with optimal performance [41]. Many or microstructural modification [2]. For example, Wauthle et al.
factors, such as powder morphology, size and chemical composi- investigated the effects of stress relief heat treatment and Hot
tion [99], as well as particle size distribution [100], laser exposure Isostatic Pressing (HIP) on the mechanical properties of Ti6Al4V
strategy and power [101], laser scan speed, and layer thickness [66] lattice structures and found stress relief to improve yield strength
all affect the quality of SLM manufacturing outcomes. These factors 15e25% and HIP to improve ductility, though, as is often the case, at
affect the mechanical properties of SLM lattice structures such as the cost of strength [98]. A comparison of the microstructures of as
strength and modulus in different ways. For example, Sing et al. built, stress relief and HIP specimens are presented in Fig. 8.
[102] found the strength and modulus, as well as dimensional ac- Maskery et al. found that heat treatment involving solution
curacy of SLM lattice structures to be most sensitive to laser power treatment followed by water quenching and artificial ageing
compared to other processing parameters such as scan speed or significantly alters lattice deformation behaviour, resulting in stress-
layer thickness. Processing parameters also affect the microstruc- strain curves more closely resembling the ideal cellular solid
ture of SLM components [103], including lattice structures [101], deformation of the Gibson-Ashby model [15]. Heat treatment did not
which affects mechanical properties. However, further characteri- increase the specific energy absorption of the tested lattice struc-
sation of the relationships between processing parameters and the tures, but the same energy was absorbed with significantly lower
mechanical properties of SLM components is still required [104]. peak stresses. However, the lattice collapse strength was reduced by
approximately 25% compared to as-built lattice structures.
6. Microstructure of SLM components and effect of heat Pattanayak et al. found that post-processing chemical and heat
treatment treatment of titanium specimens significantly improves surface
roughness [70] which has significant implications for improving
SLM has many advantages as a manufacturing method, but the fatigue strength. Specimens were soaked in sodium hydroxide
manner of manufacture (melting of metallic powder) has implica- followed by hydrochloric acid for bioactivation, and then washed
tions for the microstructure of fabricated components. with pure water. They were then heated to 600  C for 1 h then
The rapid and directional solidification associated with SLM can allowed cooling naturally in the furnace. This was also found to
significantly affect local microstructure. For example directional improve biocompatibility as it produces a sodium-free titanium
solidification can result in a preferred crystallographic orientation oxide on the surface which has also been found to form a bone-like
that can influence mechanical properties [98]. Furthermore, rapid apatite layer in a simulated body fluid [70]. However, Koehnen
solidification can lead to grain refinement, which results in found annealing 316L stainless steel lattice had no beneficial effect
improved deformation resistance as it acts as a barrier against on tensile or fatigue strength [86].
T. Maconachie et al. / Materials and Design 183 (2019) 108137 9

minimum inclination angle for overhanging surfaces below which


structures cannot be manufactured without excessive geometric
distortion [110]. This minimum inclination angle is a function of
processing parameters, material type and powder characteristics,
and is often simplified as being 45 [111]. For example, Yan et al.
found that struts with an inclination angle <30 could not be
manufactured by SLM as this led to serious distortion [47].
There is also a minimum feature size that can be reliably man-
ufactured by SLM. For example Pattanayak found the fabrication of
wall thicknesses below 300 mm to be infeasible [70]. However, to a
certain extent these concerns can be overcome by careful control of
processing parameters. For instance Santorinaios et al. were able to
reduce the attachment of unmelted particles through optimisation
of processing parameters [96].

7.2. Dimensional accuracy

Limitations of the SLM process can lead to discrepancies between


the intended and as-fabricated lattice structure geometries. These
discrepancies may be due to shrinkage after melting, attachment of
unmelted particles or waviness and roughness of struts, and have
been shown to be detrimental to the strength and stiffness of these
structures [112]. Both the nominal and measured geometry of
fabricated components are often reported in the literature for this
reason (Table 7). For example, results of micro computed tomogra-
phy (mCT) analysis of SLM lattice structures performed by Delago
et al. showed significant discrepancies between CAD models and as
fabricated geometry, with build-up of additional material in some
areas and a lack of material in others [67]. The inclusion of these
geometric defects in finite element models of SLM lattice structures
was consistently found to reduce stiffness.
The “staircase effect” (Fig. 9) is a well-known phenomenon
common to many AM methods, including SLM which affects the
Fig. 8. The microstructural effects of stress relief heat treatment (B) and hot isostatic dimensional accuracy of fabricated components [113]. Due to the
pressing (HIP) (C) compared with as built Ti6Al4V (A) [97]. In the as built condition (A)
layer-wise fabrication process, angled surfaces are fabricated with
martensitic a0 -phase within prior b grains are observed to be mostly aligned with the
build direction, whereas the heat-treated microstructure (B) shows a transformed a0 -
layer thickness-sized ridges, reducing the conformity between
phase consisting of fine a platelets where the b-phase is present. After HIP treatment intended and fabricated geometries. Surface quality is reduced,
(C) no preferential grain orientation was observed resulting in an isotropic necessitating post-processing procedures.
microstructure. Dimensional errors can occur in SLM-fabricated components due
to discrepancies between the intended geometry and the size and
These findings suggest that heat treatments provide opportu-
position of the melt pool. To compensate for this laser scan vectors
nities for designers to tune the mechanical response of SLM lattice
are usually shifted half of the laser spot size inwards [47]. However,
structures. However, as with bulk materials, there is often a trade-
applying this method may still result in errors. For example, when
off between certain properties, such as strength and ductility [108],
using a spot size of 80 mm with scan vectors shifted inwards 40 mm,
and for precise control of response, these relationships must be
Van Bael et al. found the melt pool size to be approximately 180 mm,
further understood.
constraining the feasible minimum feature size [112]. It was noted
that decreasing laser power and increasing scan velocity can reduce
7. Potential geometric defects of SLM lattice structures melt pool size, but this limits porosity control and imposes a design
constraint by reducing the controllability of pore sizes below
SLM-fabricated components are potentially prone to geometric 400 mm. Mullen et al. proposed this could be due to diffusive heating
defects which can be caused by a variety of issues such as subop- of neighbouring struts due to momentary thermal diffusion through
timal processing parameters [109] or internal defects due to the powder during scanning [74].
incomplete filling by the laser trajectory [45]. Residual stresses
resulting from the significant thermal gradients experienced by 7.3. Surface roughness
components during manufacture can also cause geometric defects
and macroscopical deformation such as severe notches or distorted The surface roughness of SLM-fabricated components is affected
struts [67]. The staircase effect is another form of geometric error by a number of factors, including the stability, dimensions and
common to AM processes including SLM. These defects have im- behaviour of the melt pool during the SLM process [105], as well as the
plications for the mechanical and biological properties of fabricated orientation of a surface to the laser beam [111] (Fig. 10). When a sur-
components and affect the manufacturability, dimensional accu- face is facing upward during fabrication it experiences direct contact
racy and surface roughness of SLM lattice structures [2]. with the laser beam, and almost all particles in contact with the sur-
face are melted. However, unsupported downward-facing surfaces
7.1. Manufacturability are in contact with the powder bed, and when the melt pool solidifies
unmelted particles become attached. Increasing the inclination angle
The geometries that can feasibly be fabricated by SLM are sub- of downward facing surfaces reduces this effect [111].
ject to manufacturability constraints. For example, there exists a Surface roughness of SLM lattice structures has been found to be
10 T. Maconachie et al. / Materials and Design 183 (2019) 108137

Fig. 9. The staircase effect e discrepancies between intended and fabricated geometries result from the layer-wise fabrication method of AM. The difference between layer
thickness and laser melt depth enables fusion between layers at overlaps but also causes the adhesion of unmelted particles to downward-facing surfaces.

significantly affected by the inclination of its component struts due mechanisms of SLM lattice structures are not only defined by cell
to the adhesion of unmelted particles to the downward-facing topology and geometry, but that geometric defects induced by SLM,
surfaces [94]. During the SLM process downward-facing surfaces such as strut waviness, strut thickness variation and oversizing of
have the potential to overheat which can cause partially fused struts also play a significant role [116]. Microstructural defects
powder particles to adhere to them [114]. The area of downward- resulting from the SLM process also negatively affect the perfor-
facing surfaces is inversely proportional to inclination angle, mance of lattice structures [117].
meaning lower inclined struts have greater downward-facing sur- The presence of defects may be reduced by optimisation of
face area, increasing the potential for the adhesion of unmelted processing parameters [109], but due to the manner of fabrication
particles and reducing surface quality. and the complexity of the process, it seems SLM-fabricated com-
However, surface roughness is not necessarily a disadvantage ponents are inherently susceptible to certain levels of defects. For
for certain applications, as various biomedical criteria for implants this reason, much research has been conducted on accounting for
such as cell attachment are improved with increased surface these defects in the numerical modelling of SLM lattice structures
roughness [115]. For example, Sarker et al. found inclination angle to more accurately predict their behaviour.
can be utilised to modify the surface topography of metallic im-
plants for directed Staphylococcus aureus biofilm restriction [77].
8. Numerical modelling of SLM lattice structures

7.4. Effects of defects on the mechanical performance of SLM lattice Efforts to develop models to predict the mechanical behaviour
structures of cellular structures date back as far as the 1950s [118], but with
the modern broad adoption of Finite Element (FE) methods, various
Geometric imperfections associated with SLM lattice structures numerical models for predicting this behaviour have recently been
are understood to have a significant effect on mechanical perfor- developed.
mance and deformation behaviour [73], and can severely However, as stated above, defects resulting from the fabrication
compromise functionality, particularly the elastic and failure process significantly affect the mechanical behaviour of SLM lattice
response of lattice structures [75]. Studies suggest the failure structures, and so it is understood that these structures cannot be

Fig. 10. Comparison of upward and downward-facing surface qualities of specimens fabricated by SLM. Very few unmelted particles become attached to upward-facing surfaces (A),
whereas downward-facing surfaces are in contact with the powder bed and many unmelted particles are attached.
T. Maconachie et al. / Materials and Design 183 (2019) 108137 11

accurately modelled based on ideal geometries. Accurate modelling


E* ¼ The static modulus of the cellular structure.
of these structures requires that manufacturing defects be Es ¼ The static modulus of the cellular structure’s bulk material.
accounted for, and as such, the incorporation of defects into C ¼ Gibson-Ashby constant, the value of which is dependent on the unit cell
simplified models of SLM lattice structures has received consider- topology and geometry and is derived from experimental results.
able research attention: s* ¼ The strength of the cellular structure.
sy,s ¼ Yield stress of the cellular structure’s bulk material.

 Liu et al. [116] identified defects using mCT and incorporated


these into models where each strut was represented by four 9.1. Gibson-Ashby model in the literature of SLM lattice structures
beam elements, with nodal offset and radius assigned based on
respective Kernel density function for the waviness and thick- Many studies on the performance of SLM lattice structures have
ness variation. found experimental results to be consistent with predictions of the
 Lei et al. [73] used mCT to quantify strut diameter variation, and Gibson-Ashby model, such as:
generated struts where variation was reconstructed such that
defect distribution locations were maintained.  McKown et al. found the modulus of BCCZ lattice structured
 Lozanovski et al. [119] used mCT to study the variation in the increased with the square of the relative density, as predicted by
principal moments of inertia, principal axis orientations and the Gibson-Ashby model, but found the modulus of BCC lattice
centroid deviation of AM struts, and proposed a method for structures to fall below predicted values. This was explained by
modelling these with elliptical cross-sections that mimic the the lack of axial struts in BCC lattice structures making them
mCT-derived properties. more compliant under compressive loads, resulting in global
 Concli and Gilioli [120] proposed a lumped parameter model for deformation with minimal local failure [123].
a lattice-based sandwich panel, the elementary cells of which  Yan et al. observed increases in both modulus and strength of
were represented by a set of non-linear springs with stiffness SLM lattice structures with increased relative density to be
behaviour determined by the response of FEM models simulated consistent with expectations of the Gibson-Ashby model [47].
under compression, shear, torsion and tension.  Zargarian et al. were able to relate relative density, bulk material
 Amani et al. [121] compared homogenous and heterogeneous fatigue properties, cell geometry and solid distribution to the
models accounting for porosity through a Gurson-Tvergaard- fatigue performance of SLM lattice structures as predicted by the
Needleman model based on mCT scanning and found the ho- Gibson-Ashby model [12].
mogenous model overpredicted the response while the het-
erogeneous model was a better match to experimental results. However, discrepancies between experimental results and the
 Ravari et al. [117] generated FE models where strut diameter and model’s predications have also been observed, which have been
axial deviations were represented using spheres with varying attributed to several factors, including:
diameters and centroid locations. These models were able to
capture the effects of microstructural imperfections on lattice  Residual stresses resulting from the SLM fabrication process can
structure performance. lead to deformation at stress levels lower than those predicted [47].
 Campoli et al. [122] represented AM struts using beam elements  Lattice density can be overestimated by means such as Archi-
of varying diameters, the diameters were drawn from a medes’ method as SLM-fabricated components often have
Gaussian distribution fitted to diameters derived from scanning unmelted powder attached to their surface, which increase
electron microscopy. volume without contributing to strength, resulting in miscal-
culation of effective relative density and discrepancies with
Balancing the competing objectives of accurate modelling of predicted results [124].
SLM lattice structures while minimising computational expense  Zadpoor notes that analytical assessments of the relative density
remains an open field for research. Progress requires a multidisci- of SLM lattice structures for the application of the Gibson-Ashby
plinary approach incorporating simulation of both the PBF process model are prone to errors due to counting the mass contributed
and the mechanical response of lattice structures, as well as char- by struts multiple times at their junction [2].
acterisation of SLM processing and manufacturability. Due to the  The difficulty of accurate identification of strut geometry for
complexity of accurate numerical modelling of lattice structures, stress calculations is noted by Zhang et al. They also note that
other methods of predicting their performance are often used, such lattice structures can exhibit both linear and non-linear elastic
as the Gibson-Ashby model. behaviours when loaded meaning identification of the correct
modulus can be difficult [3].

9. The Gibson-Ashby model However, while certain discrepancies may exist between experi-
mental results and predictions of the Gibson-Ashby model, its
The Gibson-Ashby model is the most notable and commonly
accepted model for the prediction of the properties of cellular Table 3
structures, including lattice structures [95]. A range of mechanical, Gibson-Ashby model formulae [31].
thermal and electrical properties of cellular structures can be pre- Response type Mechanical property Formula Equation number
dicted based on the structure’s relative density and are expressed as   2
fractions of the properties of structure’s parent material. Formulae
Bending-dominated Modulus (E) E r 3
¼C
Es rs
are provided for various mechanical properties, including shear and
flexural moduli, endurance limit and hardness, but all are functions Strength (s)  3
s r 2 4
¼C
of the structure’s strength or modulus. These properties are sy;s rs
dependent on the type of response exhibited by the structure
 
(bending or stretch-dominated) and have a positive power rela- Modulus (E) E r 5
¼C
tionship with the structure’s relative density [31]. The formulae for Stretch-dominated Es rs
 
the modulus and strength of bending and stretch-dominated Strength (s) s r 6
¼C
cellular structures are presented in Table 3, where: sy;s rs
12 T. Maconachie et al. / Materials and Design 183 (2019) 108137

Table 4
Quasi-static compressive test data for SLM lattice data collected from literature, ordered by unit cell topology. Data represents ranges of values reported. Dash (e) indicates an
absence of data.

Topology Material Relative Cell size Strut Geometry (N ¼ nominal, Compressive Modulus Reference Data
density (%) (mm) diameter M ¼ measured) strength (MPa) (MPa) points
(mm)

BCC 316L SS 3.5e13.8 1.25 0.19e0.22 N 0.36e5.89 17.89e378 [125] 21


BCC 316L SS 2.3e4.3 e 0.162 N 0.2e1 8.68e57.56 [64] 18
e0.23
BCC, BCCZ 316L SS 5.3e16.6 1.5e2.5 0.25 N 0.92e15 50e2700 [123] 8
BCC, BCCZ AlSi10Mg 0.7e22.2 10 1 N, M 0.46e4.36 21.71 [73] 8
e490.22
BCC, BCCZ, FCC, FCCZ Inconel 625 2.5e13.8 2e4 0.3 N, M 0.8e10.9 22.1e1246 [45] 13
BCC, BCCZ, FCC, FCCZ, FBCCZ, Ti6Al4V, 7.5e39.7 2e7.5 0.3e1 N, M 4e124 110e2780 [126] 18
FBCCXYZ AlSi12Mg
BCC, Octet-truss SS 630 (17- 43 e e N, M e 9710 [68] 5
4PH) e13,960
Diamond Ti6Al4V 3.6e26.5 e e N 8.2e99.64 370e4240 [13] 4
Dodecahedron Ti6Al4V 15.78 e e N, M 19.4e117.2 550e3490 [127] 4
e314.55
Gyroid CP titanium 26.7e31.3 2e3 e N 44.9e54.5 1465e2676 [76] 3
Gyroid, Schwartz diamond Ti6Al4V 5e95 3e7 e N 47e1559 120 [71] 4
e17,190
Octet-truss, AlSi10Mg 10.4e14.7 e e N, M 4.7e9.1 690e1250 [75] 2
Rhombicuboctahedron
Schwartz primitive, cylinder grid Ti6Al4V 70e90 1.5 0.37e0.58 N, M e 920e2420 [69] 4
Simple cubic Ti6Al4V 19.4e36.2 e 0.54e0.64 N, M 108e170 5360e8730 [72] 5
Truncated cuboctahedron, simple Ti6Al4V 20e34 e e N, M 31.7e112.6 2180e4578 [4] 12
cubic, diamond
Total 129

fundamental insight e that there is a positive power relationship   n


between the properties of a lattice structure and its relative density E r
¼C (7b)
e is generally accepted, and as such, the model is regularly refer- Es rs
enced in the literature of AM lattice structures. The Gibson-Ashby model defines exponent values depending on
whether the structure exhibits bending-dominated or stretch-
9.2. Compressive experimental data for SLM lattice structures dominated behaviour (Table 3), and coefficients are derived from
reported in the literature experimental results for given unit cell topologies. However, the
exponent may also be derived from experimental results, and co-
Experimental data from compressive testing as reported in the efficients defined based on these derived exponents.
literature was collected and is presented in Table 4. The data is or- Many studies on the compressive behaviour of SLM lattice
dered by the unit cell topologies investigated and summarised by the structures compare their experimental results with predictions of
ranges of reported values. The mechanical performance of lattice the Gibson-Ashby model, but broader insight regarding the pre-
structures is usually quantified in terms of the structure’s strength dictive capabilities of the Gibson-Ashby model for SLM lattice
and modulus meaning these are the most commonly reported values. structures is limited by the scope of these studies. To expand this
Lattice density may be measured or predicted by various insight, collected experimental data is compared with the predicted
methods, each of which has a certain level of error. For example, range of properties for bending-dominated open-celled cellular
when the density is measured from CAD data (nominal geometry), structures (Fig. 11). The Gibson-Ashby model predicts the co-
significant errors are introduced as the SLM-manufactured geom- efficients for metallic open-celled cellular structures to be in the
etry varies from that of the input CAD data. Alternately, density may range of [0.1e4] and [0.1e1] for modulus and strength respectively
be measured directly (by Archimedes’ method) or inferred from based on analytical modelling, extensive testing on polymeric
mCT data (measured geometry). These experimental methods in- foams, and empirical fits to experimental data [93].
crease the confidence in the reported relative density but incur Reported strength and modulus values were normalised against
experimental costs and effort. To allow insight into the effect of standard reported material properties as presented in Table 5.
these density quantification methods on the reported data, the From this comparison (Fig. 11) the following is observed:
prediction method is reported in Table 4.
 For both modulus and strength data, most results were found to
fall within the predicted range.
9.3. Comparison of reported experimental data with predictions of
 Although some data points do not sit within the predicted range,
the Gibson-Ashby model
this range was generated based on the behaviour of open-celled
metallic foams, and it is not unexpected that non-stochastic
As stated above, the Gibson-Ashby model relates the relative
lattice structures with repeating unit cells may exhibit
strength (s*/ss) or modulus (E*/Es) of a cellular structure with its
different behaviour.
relative density (r*/rs) in the form of a positive power relationship
 Topologies seem to congregate within certain regions of the
with a coefficient (C) and exponent (n), as presented in Eqs. (7a) and
graphs, particularly BCC lattice structures, which is consistent
(7b).
with the understanding that unit cell topology is very significant
 n in defining the behaviour of an SLM lattice structure.
s r
¼C (7a)  Although most of the topologies presented would technically be
ss rs
categorised as bending-dominated structures according to the
Maxwell criterion, some, such as the octet-truss and FBCCXYZ,
T. Maconachie et al. / Materials and Design 183 (2019) 108137 13

Fig. 11. Comparison of reported experimental compressive strength (A) and modulus (B) data with predictions of the Gibson-Ashby model.

are technically stretch-dominated structures. Despite this these predicting their behaviour. Although precise prediction of these
structures still fell within the predicted range for bending- properties may be infeasible, the model is intended to be “useful in
dominated structures. However, there are overlaps in the per- the early stages of design when approximate analysis of compo-
formance of bending and stretch-dominated structures, and no nents and structures is needed” [93].
equivalent range of coefficients is predicted for stretch-
dominated cellular structures [93]. 9.4. Regression analysis of reported experimental data
 Consistent with the findings of McKown et al. [123], the moduli
of BCC lattice structures were found to be in the lower range or To investigate the relationship between the relative density and
below the predicted values. mechanical properties (strength and modulus) of SLM lattice
structures, power regression2 analysis was performed on the
These findings suggest the Gibson-Ashby model describes the
relationships between the relative density and the properties of
2
SLM lattice structures and demonstrates it to be a useful means of Power regression identifies trends in the form as presented in Eqs. (7a) and
(7b), where “C” is the coefficient and “n” is the exponent.
14 T. Maconachie et al. / Materials and Design 183 (2019) 108137

Table 5 10. Reporting of experimental data on SLM lattice structures


Material properties used for normalisation of experimental results.

Material Property Value Reference Within the literature of SLM lattice structures there are in-
Inconel 625 Density 8.44 g/cm3 [128]
consistencies in the reporting of experimental parameters and
Yield strength 460 MPa processes. To provide a robust basis for engineering design, the
Modulus 205.8 GPa [129] reporting of SLM lattice mechanical response should be accompa-
Density 4.43 g/cm3 nied by sufficient experimental documentation such that these
TiAl64V Yield strength 880 MPa [130]
results can be understood and are repeatable. The fundamental
Modulus 193 GPa
Density 8 g/cm3 data reported to document the SLM experimental data summarised
Stainless steel Yield strength 205 MPa [131] in this work is presented in Table 7.
Modulus 193 GPa Consideration of the reported data provides the following
Density 2.67 g/cm3
insights:
AlSi10Mg Yield strength 240 MPa [132]
Modulus 70 GPa
 The geometry of the lattice structures investigated was
commonly reported (85%), but usually only the nominal ge-
experimental data collected from the literature. This analysis was
ometry (derived from analytical formula of from CAD model),
performed on different topologies discreetly, but also on the whole
whereas fewer reported the measured geometry (50%).
data set (Table 6). These identified relationships provide a resource
 Reporting of chemical composition was quite limited (25%)
for designers wishing to identify topologies appropriate to their
though those who did report composition usually reported both
needs.
the nominal composition and the actual composition of the
From this regression analysis the following is observed:
material used.
 Load rate and strain rate describe the same experimental
 There was significant variance in the correlations (R2) found for
parameter e the rate at which specimens were loaded during
different topologies. While some topologies, such as FCC had
experiments e but a minority (30%) reported this in normalised
very high correlations for both strength (99.79%) and modulus
terms (strain rate) while most (65%) reported the actual speed of
(99.79%), others, such as FBCCZ, had high correlations for one
the testing rig (load rate). Very few (10%) reported both.
property (strength ¼ 95.31%) but low correlations for the other
(modulus ¼ 46.6%). Some, such as the Gyroid, had very low
The absence of certain data is in some cases explained by the
correlations for both strength (5.64%) and modulus (1.85%).
focus of the research. For example, as Hao et al. were concerned
 The correlation for the whole data set (“all topologies”) was
with the thermal behaviour of SLM lattice structures, samples were
relatively low for both strength (70.35%) and modulus (60.25%).
not loaded, meaning there was no reportable load or strain rate.
However, this is to be expected as both coefficient and exponent
Others, such as Burton et al. [69] were concerned with modelling of
appear to be dependent on topology considering the generally
SLM lattice structures, using mechanical testing only for validation
high correlations found for individual topologies.
of these models, and so were not so concerned with certain data
 Correlations were generally higher for strength than they were
such as composition or process parameters.
for modulus, though only marginally.
However, for repeatability and robustness, the methods and
 Comparisons between the derived exponents and those pre-
materials used to produce reported data should be presented
dicted by the Gibson-Ashby model for bending and stretch-
wherever possible.
dominated structures provide insight into the general behav-
iour of these topologies. For example, the exponent for the
strength of FCCZ (n ¼ 1.59) is very close to that predicted for
stretch-dominated structures (n ¼ 1.5) suggesting this topology 11. Outstanding challenges and future directions
behaves more like a stretch-dominated structure despite being
bending-dominated according to the Maxwell criterion. Although great interest has been taken in SLM lattice structures
 Consistent with the insights of the Gibson-Ashby model, these for a variety of applications, challenges remain for their broader
findings suggest that a positive power relationship can be drawn commercial adoption, such as accounting for the manufacturing
between the relative density of an SLM lattice structure and its defects which result from the fabrication process, models for effi-
strength or modulus. This further demonstrates the relevance of ciently and accurately predicting the behaviour of lattice structures,
the Gibson-Ashby model for predicting and describing the implementation of functionally-graded lattice structures and
behaviour of SLM lattice structures. greater characterisation of fatigue behaviour.

Table 6
Regression analysis of reported lattice structure experimental data.

Unit cell topology Strength (MPa) Modulus (MPa)


2
Coefficient (C) Exponent (n) Correlation (R ) (%) Coefficient (C) Exponent (n) Correlation (R2) (%)

All topologies 0.24 1.16 68.02 0.15 1.67 60.27


BCC 0.14 1.14 66.56 0.07 1.71 66.74
BCCZ 0.06 0.48 38.84 0.01 0.22 12.42
Diamond 0.08 0.55 9.43 0.04 0.86 21.99
Dodecahedron 2.21 2.51 99.04 0.63 2.64 99.93
FBCCXYZ 1.20 1.99 98.34 0.10 1.28 87.72
FBCCZ 1.04 1.72 95.31 0.03 0.54 46.60
FCC 1.18 1.75 99.79 0.37 1.94 99.97
FCCZ 1.32 1.59 99.43 0.35 1.50 91.93
Gyroid 0.06 0.17 5.64 0.01 0.22 1.85
Schwartz diamond 0.77 1.36 99.76 0.05 0.76 96.61
Simple cubic 0.64 1.23 64.09 0.15 0.93 37.02
Average 70.35 Average 60.25
T. Maconachie et al. / Materials and Design 183 (2019) 108137 15

Table 7
Summary of fundamental data reported to document experimental parameters for reported SLM lattice test data.
Tick (✓) indicates the presence of data. Dash () indicates an absence of data.

Publication Reference Geometry Geometry Density Density Composition Composition SLM Heat Load rate Strain rate
scope (nominal) (actual) (nominal) (actual) (nominal) (actual) process treatment (mm/min) (s1)
parameters

Aerospace [14] ✓ e e e e e e e e e
Aerospace [133] e e e e e e e e 5 e
Aerospace [11] ✓ e ✓ e e e e e e e
Biomedical [71] ✓ ✓ ✓ ✓ e e ✓ ✓ 0.4 e
Biomedical [76] ✓ ✓ ✓ ✓ ✓ ✓ ✓ e e 0.001
Biomedical [5] e e ✓ e e e ✓ e 36 0.05
Biomedical [83] ✓ ✓ ✓ ✓ e e ✓ e 4.5 0.1
Biomedical [4] ✓ ✓ e ✓ e e e e e e
Biomedical [112] ✓ e e ✓ ✓ e ✓ ✓ 1 e
Energy [16] ✓ ✓ ✓ ✓ e e e e e 0.001
absorption
Energy [15] ✓ e ✓ ✓ e e ✓ ✓ 0.54 e
absorption
Energy [123] ✓ e ✓ e e e e e 0.5 e
absorption
Mechanical [65] e e ✓ e e e e e 1 e
properties
Mechanical [45] ✓ ✓ ✓ ✓ ✓ ✓ ✓ e 1.8e3.6 e
properties
Mechanical [86] ✓ e ✓ e ✓ ✓ ✓ ✓ e 0.001
properties
Mechanical [94] ✓ ✓ ✓ ✓ e e ✓ e e 0.001
properties
Mechanical [9] ✓ ✓ ✓ ✓ e e ✓ ✓ 0.4 e
properties
Mechanical [47] ✓ ✓ ✓ ✓ e e ✓ e 0.4 e
properties
Mechanical [85] ✓ e ✓ e ✓ ✓ e e 0.6 e
properties
Mechanical [134] ✓ ✓ ✓ ✓ e e ✓ e 0.4 e
properties
Number reported 17 10 16 12 5 4 12 5 13 6
Total reported (%) 85 50 80 60 25 20 60 25 65 30

11.1. Accounting for manufacturing defects capture deformation and the high number of degrees of freedom
necessary to do this, they are very computationally expensive [135].
As discussed above, SLM fabrication of lattice structures is un- As discussed, the inclusion of defects in FE models to accurately
derstood to result in manufacturing defects which affect the per- predict the behaviour of SLM lattice structures has received much
formance of SLM lattice structures. Optimisation of processing research attention, which is further exacerbated by the re-
parameters has been demonstrated to reduce these defects, how- quirements for minimising computational expense.
ever it seems a certain level of defect is inherent to the SLM
fabrication process. While certain defects such as geometric irreg-
11.3. Functionally-graded and locally tuned lattice structures
ularities or surface roughness may be overcome by post-processing,
the complexity of SLM lattice structures mean these defects must
While this review has been focused on lattice structures with a
be accounted for as they cannot be entirely eliminated. One way in
constant density, functionally-graded lattice (FGL) are another form
which these defects are being accounted for is by the inclusion of
of lattice enabled by AM. FGL have a varying density across their
defects into predictive finite element models.
structure and have several potential advantages over uniform
density lattice structures, such as greater material efficiency [136]
and superior energy absorption potential [137].
11.2. Efficient and precise predictive models
Topologically optimised structures are another form of structure
enabled by AM that have been used for similar design purposes as
To avoid the costs of fabrication and experimentation of lattice
lattice structures [138,139]. FGL are useful for the application of
structures, models which can precisely and efficiently predict the
multi-objective topology optimisation, as they can be used to
behaviour of lattice structures are necessary. While FE modelling
match density gradients within structures designed by these
provides opportunities in this space, managing the inherent trade-
methods [140]. Some studies have been conducted to quantify the
off between the computational expense and accuracy of these
properties of FGL and compare them to uniform lattice structures,
models for predicting the behaviour of SLM lattice structures re-
but these usually involve simple, linear density profiles. These
mains the focus of much research.
structures have potential for highly-tuned performance for specific
Some models have been developed that use reduced order el-
applications [141], meaning opportunities for further investigation
ements such as beam elements. These models are computationally
into applications of FGL remain.
efficient, but struggle to account for the complex interactions be-
tween struts at their intersections [135] which significantly affect
mechanical behaviour. Continuum element models that more 11.4. Fatigue behaviour of SLM lattice structures
precisely capture the behaviour of strut interactions have also been
developed. These models more accurately represent as-fabricated This paper has explored some of the reported fatigue behaviour
geometries, but due to the dense meshes necessary to accurately of lattice structures (Section 5.2) but as stated above, this behaviour
16 T. Maconachie et al. / Materials and Design 183 (2019) 108137

is still relatively undefined. Some specifics have been observed, suggests that topology is significant in determining the me-
such as the general SeN behaviour of lattice structures under fa- chanical response of SLM lattice structures.
tigue and the effects of relative density and certain topologies on  Regression analysis of the reported data demonstrated positive
lattice fatigue life, but a broader understanding of the fatigue per- power relationships with high correlations could be drawn be-
formance of AM lattice structures is yet to be achieved. tween the relative density and strength of modulus for most
topologies, consistent with the insights of the Gibson-Ashby
11.5. Future work model.

Other areas of study which would facilitate the greater charac-


terisation and commercial adoption of SLM lattice structures CRediT authorship contribution statement
include:
Tobias Maconachie: Formal analysis, Investigation, Data cura-
 The performance of SLM lattice structures have been charac- tion, Writing - original draft, Writing - review & editing, Visuali-
terised in their as fabricated state, but knowledge of their per- zation. Martin Leary: Conceptualization, Methodology, Writing -
formance during operation is still limited. The thermal original draft, Writing - review & editing, Project administration,
behaviour of SLM lattice structures during operation, especially Funding acquisition. Bill Lozanovski: Writing - original draft.
the effect of atmospheric phenomena for aerospace applications Xuezhe Zhang: Conceptualization, Methodology. Ma Qian: Writing
[14], and the effect of biodegradation on the fatigue strength of - review & editing, Supervision. Omar Faruque: Resources, Su-
SLM lattice implants for biomedical applications [2] are yet to be pervision, Project administration, Funding acquisition. Milan
characterised. Brandt: Resources, Supervision, Project administration.
 Most research on SLM lattice structures focuses on either strut-
based or TPMS lattice structures discretely. Research on their Acknowledgements
comparative performance is still needed to identify the appro-
priateness of these design approaches for different applications This research was conducted by the Australian Research Council
and performance requirements. Industrial Transformation Training Centre in Additive Bio-
 The tensile performance of SLM lattice structures is still not well manufacturing (IC160100026).
defined as most research uses compressive experiments for The authors would like to acknowledge the financial support
identification of their mechanical properties largely due to the from the members of the ARC Training Centre for Lightweight
greater difficulty of tensile lattice experimental design. Automotive Structures and from the Australian Research Council
(Grant Reference IC160100032).
Knowledge of the behaviour of lattice structures is quite The authors acknowledge the facilities, and the scientific and
disparate, as researchers have investigated isolated areas of the technical assistance of the RMIT Advanced Manufacturing Precinct.
broad field of lattice structures. Greater homogenisation and
standardisation of this field could lead to new insights and enhance
References
the applicability of SLM lattice structures for all applications [96].
[1] L. Hao, et al., Design and additive manufacturing of cellular lattice structures,
12. Concluding remarks in: The International Conference on Advanced Research in Virtual and Rapid
Prototyping (VRAP), Taylor & Francis Group, Leiria, 2011.
[2] A.A. Zadpoor, Mechanical performance of additively manufactured meta-
SLM allows the fabrication of complex geometries with high biomaterials, Acta Biomater. (2018).
resolution, which is perfect for the manufacture of lattice struc- [3] X.Z. Zhang, et al., Selective electron beam manufactured Ti-6Al-4V lattice
tures. Using this technology, lattice structures with highly tuned structures for orthopedic implant applications: Current status and
outstanding challenges, Curr. Opinion Solid State Mater. Sci. 22 (3) (2018)
geometries and topologies can be fabricated to produce a wide 75e99.
variety of properties unachievable by their bulk materials. This has [4] S.A. Yavari, et al., Relationship between unit cell type and porosity and the
led to research into these structures, particularly for biomedical, fatigue behavior of selective laser melted meta-biomaterials, J. Mech. Behav.
Biomed. Mater. 43 (2015) 91e100.
light-weighting and energy-absorption applications. However,
[5] E. Alabort, D. Barba, R.C. Reed, Design of metallic bone by additive
there exists no overarching analysis of this data. This work has manufacturing, Scr. Mater. 164 (2019) 110e114.
sought to overcome this limitation by providing an overview of the [6] J. Christensen, F.J.G. de Abajo, Anisotropic metamaterials for full control of
literature regarding SLM lattice structures, with meta-analysis on acoustic waves, Phys. Rev. Lett. 108 (12) (2012) 124301.
[7] U. Levy, et al., Inhomogenous dielectric metamaterials with space-variant
the reported data. Key findings of this analysis include: polarizability, Phys. Rev. Lett. 98 (24) (2007) 243901.
[8] N. Fang, et al., Ultrasonic metamaterials with negative modulus, Nat. Mater. 5
 Most of the strength and modulus data analysed fell within the (6) (2006) 452.
[9] C. Yan, et al., Microstructure and mechanical properties of aluminium alloy
range of values predicted by the Gibson-Ashby model, sug- cellular lattice structures manufactured by direct metal laser sintering,
gesting positive power relationships between the relative den- Mater. Sci. Eng. A 628 (2015) 238e246.
sity and mechanical properties of SLM lattice structures, and [10] D. Shidid, et al., Just-in-time Design and Additive Manufacture of Patient-
specific Medical Implants, Phys. Procedia 83 (2016) 4e14.
indicating the model to be a useful tool for describing and [11] Z. Hao, et al., Lightweight structure of a phase-change thermal controller
predicting the mechanical behaviour of SLM lattice structures. based on lattice cells manufactured by SLM, Chin. J. Aeronaut. 32 (7) (2019)
 Certain topologies, such as the octet-truss and FBCXYZ, which 1727e1732.
[12] A. Zargarian, et al., On the Fatigue Behavior of Additive Manufactured Lattice
are technically stretch-dominated according to the Maxwell
Structures, Theor. Appl. Fract. Mech. 100 (2019) 225e232.
criterion fell within the range of values predicted for bending- [13] S.M. Ahmadi, et al., Mechanical behavior of regular open-cell porous bio-
dominated structures. Although there are overlaps in these materials made of diamond lattice unit cells, J. Mech. Behav. Biomed. Mater.
34 (2014) 106e115.
behaviours, this does suggest that in terms of strength and
[14] M. Bici, et al., Development of a multifunctional panel for aerospace use
modulus bending-dominated predictions may be appropriate to through SLM Additive Manufacturing, Procedia CIRP 67 (1) (2018) 215e220.
stretch-dominated structures. [15] I. Maskery, et al., Compressive failure modes and energy absorption in
 Topologies tended to congregate in certain areas of the strength additively manufactured double gyroid lattices, Addit. Manuf. 16 (2017)
24e29.
and modulus graphs e for example BCC lattice structures tended [16] J.A. Harris, R.E. Winter, G.J. McShane, Impact response of additively manu-
to congregate around the lower end of predicted range e which factured metallic hybrid lattice materials, Int. J. Impact Eng. 104 (2017)
T. Maconachie et al. / Materials and Design 183 (2019) 108137 17

177e191. (45) (2018), 1803334.


[17] H. Gong, et al., Analysis of defect generation in Ti-6Al-4V parts made using [52] S.C. Han, J.W. Lee, K. Kang, A New Type of Low Density Material: Shellular,
powder bed fusion additive manufacturing processes, Addit. Manuf. 1 (2014) Adv. Mater. 27 (37) (2015) 5506e5511.
87e98. [53] I. Gibson, D.W. Rosen, B. Stucker, Additive manufacturing technologies Vol.
[18] L.J. Gibson, M.F. Ashby, B.A. Harley, Cellular materials in nature and medicine, 17, Springer, 2014.
Cambridge University Press, 2010. [54] I. Yadroitsev, P. Bertrand, I. Smurov, Parametric analysis of the selective laser
[19] J. Banhart, H.W. Seeliger, Aluminium Foam Sandwich Panels: Manufacture, melting process, Appl. Surf. Sci. 253 (19) (2007) 8064e8069.
Metallurgy and Applications, Adv. Eng. Mater. 10 (9) (2008) 793e802. [55] O. Rehme, C. Emmelmann, Rapid manufacturing of lattice structures with
[20] F. Zhu, et al., Plastic Deformation, Failure and Energy Absorption of Sandwich selective laser melting, in: Laser-based Micropackaging, International Society
Structures with Metallic Cellular Cores, Int. J. Protective Struct. 1 (4) (2010) for Optics and Photonics, 2006.
507e541. [56] L. Yuan, S. Ding, C. Wen, Additive manufacturing technology for porous metal
[21] J. Banhart, Manufacturing routes for metallic foams, Jom 52 (12) (2000) implant applications and triple minimal surface structures: A review,
22e27. Bioactive Mater. 4 (1) (2019) 56e70.
[22] S. Santosa, T. Wierzbicki, Crash behavior of box columns filled with [57] K.G. Prashanth, et al., Microstructure and mechanical properties of Ale12Si
aluminum honeycomb or foam, Comput. Struct. 68 (4) (1998) 343e367. produced by selective laser melting: Effect of heat treatment, Mater. Sci. Eng.
[23] M. Seitzberger, et al., Crushing of axially compressed steel tubes filled with A 590 (2014) 153e160.
aluminium foam, Acta Mech. 125 (1-4) (1997) 93e105. [58] J.-P. Kruth, et al., Selective laser melting of iron-based powder, J. Mater.
[24] A. Hanssen, M. Langseth, O. Hopperstad, Static crushing of square aluminium Process. Technol. 149 (1-3) (2004) 616e622.
extrusions with aluminium foam filler, Int. J. Mech. Sci. 41 (8) (1999) [59] J.N. Domfang Ngnekou, et al., Influence of defect size on the fatigue resis-
967e993. tance of AlSi10Mg alloy elaborated by selective laser melting (SLM), Procedia
[25] F. Simancik, et al., Aluminium foam-a new light-weight structural material, Struct. Integrity 7 (2017) 75e83.
Met. Mater. 35 (1997) 187e194. [60] M. Hirsch, et al., Meso-scale defect evaluation of selective laser melting using
[26] J. Banhart, C. Schmoll, U. Neumann, Light-weight aluminium foam structures spatially resolved acoustic spectroscopy, Proc. R. Soc. A Math. Phys. Eng. Sci.
for ships, in: Proceedings of the Conference on Materials in Oceanic Envi- 473 (2205) (2017), 20170194.
ronment, 1998. [61] E. Liverani, et al., Effect of selective laser melting (SLM) process parameters
[27] T. Mukai, et al., Experimental study of energy absorption in a close-celled on microstructure and mechanical properties of 316L austenitic stainless
aluminum foam under dynamic loading, Scr. Mater. 40 (8) (1999). steel, J. Mater. Process. Technol. 249 (2017) 255e263.
[28] M. Smith, Z. Guan, W.J. Cantwell, Finite element modelling of the compres- [62] A. Armillotta, R. Baraggi, S. Fasoli, SLM tooling for die casting with conformal
sive response of lattice structures manufactured using the selective laser cooling channels, Int. J. Adv. Manuf. Technol. 71 (1) (2014) 573e583.
melting technique, Int. J. Mech. Sci. 67 (2013) 28e41. [63] T. Hermann Becker, The achievable mechanical properties of SLM produced
[29] C. Yan, et al., Evaluation of light-weight AlSi10Mg periodic cellular lattice Maraging Steel 300 components, Rapid Prototyp. J. 22 (3) (2016) 487e494.
structures fabricated via direct metal laser sintering, J. Mater. Process. [64] S. Tsopanos, et al., The Influence of Processing Parameters on the Mechanical
Technol. 214 (4) (2014) 856e864. Properties of Selectively Laser Melted Stainless Steel Microlattice Structures,
[30] L.J. Gibson, Modelling the mechanical behavior of cellular materials, Mater. J. Manuf. Sci. Eng. 132 (4) (2010) 041011. -041011-12.
Sci. Eng. A 110 (1989) 1e36. [65] R.R. Rashid, et al., A comparative study of flexural properties of additively
[31] M. Ashby, The properties of foams and lattices, Philos. Trans. R. Soc. A Math. manufactured aluminium lattice structures, Mater. Today Proc. 4 (8) (2017)
Phys. Eng. Sci. 364 (1838) (2005) 15e30. 8597e8604.
[32] H.L. Fan, F.H. Meng, W. Yang, Sandwich panels with Kagome lattice cores [66] P. Hanzl, et al., The Influence of Processing Parameters on the Mechanical
reinforced by carbon fibers, Compos. Struct. 81 (4) (2007) 533e539. Properties of SLM Parts, Procedia Eng. 100 (2015) 1405e1413.
[33] G.W. Kooistra, V.S. Deshpande, H.N.G. Wadley, Compressive behavior of age [67] M. Dallago, et al., Effect of the geometrical defectiveness on the mechanical
hardenable tetrahedral lattice truss structures made from aluminium, Acta properties of SLM biomedical Ti6Al4V lattices, Procedia Struct. Integrity 13
Mater. 52 (14) (2004) 4229e4237. (2018) 161e167.
[34] D.T. Queheillalt, Y. Murty, H.N.G. Wadley, Mechanical properties of an [68] D. Kang, et al., Multi-lattice inner structures for high-strength and light-
extruded pyramidal lattice truss sandwich structure, Scr. Mater. 58 (1) weight in metal selective laser melting process, Mater. Des. 175 (2019),
(2008) 76e79. 107786.
[35] H. Fan, et al., Interlocked hierarchical lattice materials reinforced by woven [69] H.E. Burton, et al., The design of additively manufactured lattices to increase
textile sandwich composites, Compos. Sci. Technol. 87 (2013) 142e148. the functionality of medical implants, Mater. Sci. Eng. C 94 (2019) 901e908.
[36] H.-L. Fan, et al., Mechanics of advanced fiber reinforced lattice composites, [70] D.K. Pattanayak, et al., Bioactive Ti metal analogous to human cancellous
Acta Mech. Sinica 26 (6) (2010) 825e835. bone: fabrication by selective laser melting and chemical treatments, Acta
[37] W. Li, et al., A novel carbon fiber reinforced lattice truss sandwich cylinder: Biomater. 7 (3) (2011) 1398e1406.
Fabrication and experiments, Compos. A: Appl. Sci. Manuf. 81 (2016) [71] C. Yan, et al., Tie6Ale4V triply periodic minimal surface structures for bone
313e322. implants fabricated via selective laser melting, J. Mech. Behav. Biomed.
[38] T. Wohlers, T. Gornet, History of additive manufacturing, in: Wohlers report, Mater. 51 (2015) 61e73.
2014 24, 2014, p. 118. [72] E. Sallica-Leva, A.L. Jardini, J.B. Fogagnolo, Microstructure and mechanical
[39] G. Dong, Y. Tang, Y.F. Zhao, A Survey of Modeling of Lattice Structures behavior of porous Tie6Ale4V parts obtained by selective laser melting,
Fabricated by Additive Manufacturing, J. Mech. Des. 139 (10) (2017) 100906. J. Mech. Behav. Biomed. Mater. 26 (2013) 98e108.
-100906-13. [73] H. Lei, et al., Evaluation of compressive properties of SLM-fabricated multi-
[40] W. Brooks, et al., Rapid design and manufacture of ultralight cellular mate- layer lattice structures by experimental test and m-CT-based finite element
rials, in: Proceedings of the Solid Freeform Fabrication Symposium, Austin, analysis, Mater. Des. 169 (2019), 107685.
TX, 2005. [74] L. Mullen, et al., Selective Laser Melting: A regular unit cell approach for the
[41] M. Leary, et al., Selective laser melting (SLM) of AlSi12Mg lattice structures, manufacture of porous, titanium, bone in-growth constructs, suitable for
Mater. Des. 98 (2016) 344e357. orthopedic applications, J Biomed Mater Res B Appl Biomater 89 (2) (2009)
[42] Z. Xiao, et al., Evaluation of topology-optimized lattice structures manufac- 325e334.
tured via selective laser melting, Mater. Des. 143 (2018) 27e37. [75] S. Arabnejad, et al., High-strength porous biomaterials for bone replacement:
[43] S. Xu, et al., Design of lattice structures with controlled anisotropy, Mater. A strategy to assess the interplay between cell morphology, mechanical
Des. 93 (2016) 443e447. properties, bone ingrowth and manufacturing constraints, Acta Biomater. 30
[44] V.S. Deshpande, N.A. Fleck, M.F. Ashby, Effective properties of the octet-truss (2016) 345e356.
lattice material, J. Mech. Phys. Solids 49 (8) (2001) 1747e1769. [76] A. Ataee, et al., Ultrahigh-strength titanium gyroid scaffolds manufactured by
[45] M. Leary, et al., Inconel 625 lattice structures manufactured by selective laser selective laser melting (SLM) for bone implant applications, Acta Mater. 158
melting (SLM): Mechanical properties, deformation and failure modes, (2018) 354e368.
Mater. Des. 157 (2018) 179e199. [77] A. Sarker, et al., Rational Design of Additively Manufactured Ti6Al4V Im-
[46] I. Maskery, et al., Insights into the mechanical properties of several triply plants to Control Staphylococcus aureus Biofilm Formation, Materialia
periodic minimal surface lattice structures made by polymer additive (2019), 100250.
manufacturing, Polymer 152 (2018) 62e71. [78] J. Brennan-Craddock, et al., The design of impact absorbing structures for
[47] C. Yan, et al., Advanced lightweight 316L stainless steel cellular lattice additive manufacture, J. Phys. Conf. Ser. 382 (2012), 012042.
structures fabricated via selective laser melting, Mater. Des. 55 (2014) [79] A. Subic, Materials in sports equipment, Woodhead Publishing, 2019.
533e541. [80] A. Owens, et al., Benefits of additive manufacturing for human exploration of
[48] A.A. Zadpoor, Bone tissue regeneration: the role of scaffold geometry, Bio- mars, in: 45th International Conference on Environmental Systems, 2015.
mater. Sci. 3 (2) (2015) 231e245. [81] N. Guo, M.C. Leu, Additive manufacturing: technology, applications and
[49] J.B. Berger, H.N.G. Wadley, R.M. McMeeking, Mechanical metamaterials at research needs, Front. Mech. Eng. 8 (3) (2013) 215e243.
the theoretical limit of isotropic elastic stiffness, Nature 543 (2017) 533. [82] Allied Market Research, Medical Implants Market Size, Share | Industry
[50] C. Bonatti, D. Mohr, Mechanical performance of additively-manufactured Analysis 2022, Available from, https://www.alliedmarketresearch.com/
anisotropic and isotropic smooth shell-lattice materials: Simulations & ex- medical-implants-market, 2019.
periments, J. Mech. Phys. Solids 122 (2019) 1e26. [83] H.E. Burton, et al., The design of additively manufactured lattices to increase
[51] T. Tancogne-Dejean, et al., 3D Plate-Lattices: An Emerging Class of Low- the functionality of medical implants, Mater. Sci. Eng. C 94 (2019) 901e908.
Density Metamaterial Exhibiting Optimal Isotropic Stiffness, Adv. Mater. 30 [84] L. Nickels, AM and aerospace: an ideal combination, Met. Powder Rep. 70 (6)
18 T. Maconachie et al. / Materials and Design 183 (2019) 108137

(2015) 300e303. method for metal additive manufacturing: A reduced-order DFAM tool
[85] J. Zhou, P. Shrotriya, W. Soboyejo, On the deformation of aluminum lattice applied to SLM, Mater. Des. 132 (2017) 226e243.
block structures: from struts to structures, Mech. Mater. 36 (8) (2004) [115] G. Pyka, et al., Surface Roughness and Morphology Customization of Additive
723e737. Manufactured Open Porous Ti6Al4V Structures, Materials 6 (10) (2013)
[86] P. Koehnen, et al., Mechanical properties and deformation behavior of 4737e4757.
additively manufactured lattice structures of stainless steel, Mater. Des. 145 [116] L. Liu, et al., Elastic and failure response of imperfect three-dimensional
(2018) 205e217. metallic lattices: the role of geometric defects induced by Selective Laser
[87] S. Babaee, et al., 3D soft metamaterials with negative Poisson’s ratio, Adv. Melting, J. Mech. Phys. Solids 107 (2017) 160e184.
Mater. 25 (36) (2013) 5044e5049. [117] M.R.K. Ravari, et al., On the effects of geometry, defects, and material
[88] E.B. Duoss, et al., Three-dimensional printing of elastomeric, cellular archi- asymmetry on the mechanical response of shape memory alloy cellular
tectures with negative stiffness, Adv. Funct. Mater. 24 (31) (2014) lattice structures, Smart Mater. Struct. 25 (2) (2016), 025008.
4905e4913. [118] M.R.K. Ravari, et al., Numerical investigation on mechanical properties of
[89] J.N. Grima, et al., Negative linear compressibility of hexagonal honeycombs cellular lattice structures fabricated by fused deposition modeling, Int. J.
and related systems, Scr. Mater. 65 (7) (2011) 565e568. Mech. Sci. 88 (2014) 154e161.
[90] Q. Wang, et al., Lightweight mechanical metamaterials with tunable negative [119] B. Lozanovski, et al., Computational modelling of strut defects in SLM man-
thermal expansion, Phys. Rev. Lett. 117 (17) (2016) 175901. ufactured lattice structures, Mater. Des. 171 (2019), 107671.
[91] X. Zheng, et al., Ultralight, ultrastiff mechanical metamaterials, Science 344 [120] F. Concli, A. Gilioli, Numerical and experimental assessment of the me-
(6190) (2014) 1373e1377. chanical properties of 3D printed 18-Ni300 steel trabecular structures pro-
[92] H. Alsalla, L. Hao, C. Smith, Fracture toughness and tensile strength of 316L duced by Selective Laser Melting e a lean design approach, Virtual Phys.
stainless steel cellular lattice structures manufactured using the selective Prototyping 14 (3) (2019) 267e276.
laser melting technique, Mater. Sci. Eng. A 669 (2016) 1e6. [121] Y. Amani, et al., Compression behavior of lattice structures produced by
[93] M.F. Ashby, et al., Metal foams: a design guide, Elsevier, 2000. selective laser melting: X-ray tomography based experimental and finite
[94] M. Mazur, et al., Mechanical properties of Ti6Al4V and AlSi12Mg lattice element approaches, Acta Mater. 159 (2018) 395e407.
structures manufactured by selective laser melting (SLM), in: Laser Additive [122] G. Campoli, et al., Mechanical properties of open-cell metallic biomaterials
Manufacturing: Materials, Design, Technologies, and Applications, 2016, manufactured using additive manufacturing, Mater. Des. 49 (2013) 957e965.
pp. 119e161. [123] S. McKown, et al., The quasi-static and blast loading response of lattice
[95] L.J. Gibson, M.F. Ashby, Cellular solids: structure and properties, Cambridge structures, Int. J. Impact Eng. 35 (8) (2008) 795e810.
university press, 1999. [124] M. Yakout, M.A. Elbestawi, S.C. Veldhuis, Density and mechanical properties
[96] M. Santorinaios, et al., Crush behaviour of open cellular lattice structures in selective laser melting of Invar 36 and stainless steel 316L, J. Mater.
manufactured using selective laser melting, WIT Trans. Built Environ. 85 Process. Technol. 266 (2019) 397e420.
(2006). [125] R. Gümrük, R.A.W. Mines, Compressive behaviour of stainless steel micro-
[97] S. Zhao, et al., The influence of cell morphology on the compressive fatigue lattice structures, Int. J. Mech. Sci. 68 (2013) 125e139.
behavior of Ti-6Al-4V meshes fabricated by electron beam melting 59, 2016, [126] M. Mazur, et al., 5 - Mechanical properties of Ti6Al4V and AlSi12Mg lattice
pp. 251e264. structures manufactured by Selective Laser Melting (SLM), in: M. Brandt
[98] R. Wauthle, et al., Effects of build orientation and heat treatment on the (Ed.), Laser Additive Manufacturing, Woodhead Publishing, 2017,
microstructure and mechanical properties of selective laser melted Ti6Al4V pp. 119e161.
lattice structures, Addit. Manuf. 5 (2015) 77e84. [127] S. Amin Yavari, et al., Fatigue behavior of porous biomaterials manufactured
[99] Kempen, K., et al. Process optimization and microstructural analysis for selec- using selective laser melting, Mater. Sci. Eng. C 33 (8) (2013) 4849e4858.
tive laser melting of AlSi10Mg. [128] ASM Aerospace Specification Metals Inc, ASM Material Data Sheet - Inconel
[100] E. Brandl, et al., Additive manufactured AlSi10Mg samples using Selective 625, Available from, http://asm.matweb.com/search/SpecificMaterial.asp?
Laser Melting (SLM): Microstructure, high cycle fatigue, and fracture bassnum¼NINC33, 2019.
behavior, Mater. Des. 34 (2012) 159e169. [129] Alloy Wire International, Inconel® 625 - Alloy Wire International, Available
[101] T. Niendorf, F. Brenne, M. Schaper, Lattice Structures Manufactured by SLM: from, https://www.alloywire-au.com/products/inconel-625/, 2019.
On the Effect of Geometrical Dimensions on Microstructure Evolution During [130] ASM Aerospace Specification Metals Inc, ASM Material Data Sheet - Ti6Al4V,
Processing, Metall. Mater. Trans. B 45 (4) (2014) 1181e1185. Available from, http://asm.matweb.com/search/SpecificMaterial.asp?
[102] S.L. Sing, F.E. Wiria, W.Y. Yeong, Selective laser melting of lattice structures: bassnum¼mtp641, 2019.
A statistical approach to manufacturability and mechanical behavior, Robot. [131] ASM Aerospace Specification Metals Inc, ASM Material Data Sheet - 316L
Comput. Integr. Manuf. 49 (2018) 170e180. Stainless Steel, Available from, http://asm.matweb.com/search/
[103] P. Delroisse, et al., Effect of strut orientation on the microstructure hetero- SpecificMaterial.asp?bassnum¼mq316q, 2019.
geneities in AlSi10Mg lattices processed by selective laser melting, Scr. [132] GPI Prototype & Manufacturing Services, AlSi10Mg Material Data Sheet,
Mater. 141 (2017) 32e35. Available from: https://gpiprototype.com/pdf/EOS_Aluminium_AlSi10Mg_
[104] T.M. Mower, M.J. Long, Mechanical behavior of additive manufactured, en.pdf, 2019.
powder-bed laser-fused materials, Mater. Sci. Eng. A 651 (2016) 198e213. [133] C.G. Ferro, et al., Lattice structured impact absorber with embedded anti-
[105] L. Thijs, et al., A study of the microstructural evolution during selective laser icing system for aircraft wings fabricated with additive SLM process,
melting of Tie6Ale4V, Acta Mater. 58 (9) (2010) 3303e3312. Mater. Today Commun. 15 (2018) 185e189.
[106] L. Murr, et al., Microstructure and mechanical behavior of Tie6Ale4V pro- [134] C. Yan, et al., Evaluations of cellular lattice structures manufactured using
duced by rapid-layer manufacturing, for biomedical applications, J. Mech. selective laser melting, Int. J. Mach. Tools Manuf. 62 (2012) 32e38.
Behav. Biomed. Mater. 2 (1) (2009) 20e32. [135] M.H. Luxner, J. Stampfl, H.E. Pettermann, Finite element modeling concepts
[107] S. Li, G. Cui, Dependence of strength, elongation, and toughness on grain size and linear analyses of 3D regular open cell structures, J. Mater. Sci. 40 (22)
in metallic structural materials, J. Appl. Phys. 101 (8) (2007), 083525. (2005) 5859e5866.
[108] Z. Li, et al., Metastable high-entropy dual-phase alloys overcome the [136] I. Maskery, et al., An investigation into reinforced and functionally graded
strength-ductility trade-off, Nature 534 (2016) 227. lattice structures, J. Cell. Plast. 53 (2) (2016) 151e165.
[109] H. Gong, et al., Influence of defects on mechanical properties of Tie6Ale4 V [137] S.Y. Choy, et al., Compressive properties of functionally graded lattice
components produced by selective laser melting and electron beam melting, structures manufactured by selective laser melting, Mater. Des. 131 (2017)
Mater. Des. 86 (2015) 545e554. 112e120.
[110] B. Vandenbroucke, J.-P. Kruth, Selective laser melting of biocompatible [138] K.-U. Bletzinger, E. Ramm, Structural optimization and form finding of light
metals for rapid manufacturing of medical parts, Rapid Prototyp. J. 13 (4) weight structures, Comput. Struct. 79 (22-25) (2001) 2053e2062.
(2007) 196e203. [139] C.B. Pedersen, Topology optimization design of crushed 2D-frames for
[111] A. Sarker, et al., Angle defines attachment: Switching the biological response desired energy absorption history, Struct. Multidiscip. Optim. 25 (5-6) (2003)
to titanium interfaces by modifying the inclination angle during selective 368e382.
laser melting, Mater. Des. 154 (2018) 326e339. [140] I. Maskery, et al., A mechanical property evaluation of graded density Al-
[112] S. Van Bael, et al., Micro-CT-based improvement of geometrical and me- Si10-Mg lattice structures manufactured by selective laser melting, Mater.
chanical controllability of selective laser melted Ti6Al4V porous structures, Sci. Eng. A 670 (2016) 264e274.
Mater. Sci. Eng. A 528 (24) (2011) 7423e7431. [141] A. Panesar, et al., Strategies for functionally graded lattice structures derived
[113] M.A. Isa, I. Lazoglu, Five-axis additive manufacturing of freeform models using topology optimisation for Additive Manufacturing, Addit. Manuf. 19
through buildup of transition layers, J. Manuf. Syst. 50 (2019) 69e80. (2018) 81e94.
[114] M. McMillan, M. Leary, M. Brandt, Computationally efficient finite difference

You might also like