You are on page 1of 10

Materials and Design 197 (2021) 109096

Contents lists available at ScienceDirect

Materials and Design

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

Increased efficiency gyroid structures by tailored material distribution


David Downing a,b, Alistair Jones a,c, Milan Brandt a,b, Martin Leary a,b,c,⁎
a
RMIT Centre for Additive Manufaturing, RMIT University, Melbourne, Australia
b
ARC Industrial Training Centre for Additive Biomanufacturing, Australia
c
ARC Training Centre for Lightweight Automotive Structures, Australia

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

• Identification of highly stressed helical


subregions within the solid-surface
gyroid structure under compressive
load.
• Quantification of the correlation be-
tween stress intensity with the dot
product of the load direction and local
surface normal orientation for the heli-
cal subregions.
• Local thickness adjustment of the helical
subregions, enables tuning of the me-
chanical response in multiple directions.
• Application of variable local thickness
distribution effected an increase in
compressive stiffness and strength of
approximately 20% at fixed relative
density.

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

Article history: The gyroid has been identified as a cellular topology suitable for engineering applications and, particularly in its
Received 27 July 2020 solid-network form, for biomedical applications. Its solid-surface form has received less attention but offers ad-
Received in revised form 21 August 2020 ditional benefits of a continuous surface that partitions space and provides large surface area for heat transfer
Accepted 23 August 2020
or cell attachment. Through numerical methods it is shown for the first time that under uniaxial loading the
Available online 26 August 2020
solid-surface gyroid distributes loads predominantly within helical substructures. Furthermore, by adjusting
Keywords:
the thickness of the load carrying helical regions, the mechanical response can be tuned in different directions.
Triply periodic minimal surfaces This novel discovery enables more efficient use of material distribution while taking advantage of the unique
Compression properties of the solid-surface gyroid.
Finite element analysis © 2020 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
Functional grading creativecommons.org/licenses/by-nc-nd/4.0/).
Relative stiffness

1. Introduction (TPMS) which was neither self-intersecting nor contained any straight
lines [1]. The gyroid has been recognised in biological and crystallo-
The gyroid was described by Schoen in 1970, as a surface topology graphic structures [2,3]; as photonic crystal structures in butterfly
which obeys the requirements of a triply periodic minimal surface wing scales [4], and mitochondria membranes on retinal cones in spe-
cies of tree shrews [5].
⁎ Corresponding author at: RMIT Centre for Additive Manufaturing, RMIT University,
Due to its smooth repeating surface with locally minimum area
Melbourne, Australia. which divides space into two distinct regions, the gyroid offers opportu-
E-mail address: martin.leary@rmit.edu.au (M. Leary). nities for efficiency in engineering applications. However, due to its

https://doi.org/10.1016/j.matdes.2020.109096
0264-1275/© 2020 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/).
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

complex shape - difficult to produce with traditional manufacturing bending-dominated, displaying stiffness indices between 1.6 and 2.4;
methods - it did not receive much attention as a load bearing depending on respective orientation of the unit cell to the load [21].
engineering structure. Prior to the flexibility provided by additive On the other hand, the solid-surface gyroid stiffness response has
manufacturing (AM), foams represented some of the earliest been reported as closer to stretch-dominated with n ≈ 1.4 [13], al-
engineered forms of cellular structures. Attention to the gyroid grew though only primary loading directions 〈001〉 were presented. At the
with the emergence of industrial AM processes which could replicate same time, the conductivity of the solid-surface gyroid was described
its form in advanced engineering materials [6,7]. Compared to the as isotropic, however only the three primary directions were evaluated.
random pore architecture produced by non-AM methods, the controlled The repeatability of the solid-surface gyroid structure, and the ease
pore size of a stereolithographic produced gyroid demonstrated a with which its volume fraction can be managed with local thickness
10-fold increase in permeability [8]. Structures based on the gyroid has enabled the gyroid to be exploited as a building block for radio-
and other TPMS can now be found in aerospace, automotive and bio- graphic dosage phantoms, providing indistinguishable readings in
medical applications [9,10].The material efficiency of the gyroid pro- Hounsfeld units (a measure of radiodensity) between orientations as
vides structural performance benefits for light-weighting and stiffness opposed to alternate slit based phantoms [22]. The three-dimensional
modification as well as large surface area for heat transfer or biological repeating nature of the gyroid, and the symmetry of the terms in its
cell attachment [10–13]. representative level-set equations might lead to the presumption that
There are several structural forms derived from the gyroid surface the gyroid structures are isotropic in their mechanical response. How-
which are of engineering interest, including: the lattice or solid- ever, both the structure's topology and the AM process used to fabricate
network form (filling one side with material leaving the other as a them introduce directional differences. The AM process can influence
void) [14], the cellular, matrix or solid-surface form [15] (a thickened microstructural alignment and introduce orientation dependence in
surface, maintaining two distinct void regions) (Fig. 1). the mechanical response [17]. Numerical investigation of compressive
The solid-network form has received considerable attention in stiffness and strength for the solid-network form showed up to 30% var-
biomedical applications for use in patient specific implants. [17,18]. iation of strength and stiffness with load orientation [21,23]. Further-
However, the solid-surface gyroid consistently shows superior perfor- more, as commonly observed in AM parts, the as-manufactured
mance compared to the solid-network gyroid, particularly at low thickness can be dependent on build orientation; for instance, on a
volume fraction [15]. 316L stainless steel AM gyroid, the regions fabricated with surface nor-
Besides the gyroid, a variety of lattice and cellular structures are seen mal perpendicular to the build direction were undersized, while regions
in nature providing high material efficiency, such as light weight trabec- with surface normal parallel to the build direction were oversized [24].
ular and cortical bone structures, wood and honeycombs [19]. These The surface-based gyroid structures show improved fatigue perfor-
benefits are sought in engineering applications, where lattice and cellu- mance over strut based lattices [25]. Nevertheless, the surface gyroids
lar structures have been employed to efficiently attenuate the mechan- tend to suffer from crack initiation from lack-of-fusion micro-pores
ical, thermal and electrical response of their parent material. A useful under the surface and from increased surface roughness [26]. Process
metric is the ratio of the density of the cellular structure, ρ, to the den- parameters were more effective at enhancing fatigue performance on
sity of its parent material, ρs, known as the relative density or volume thicker walled structures (50–60% macro-porosity) than on thin walled
fraction. In relation to the properties of the solid material, these struc- structures (80–90% macro-porosity) [27]. As-built and stress relieved
tures tend to respond to an increase in their relative density with a Ti6Al4V gyroids tend to show brittle fracture at relative low cycle num-
power law behaviour [20]. A physical property of the cellular structure, ber, while hot isostatic pressing can extend fatigue life [26].
P, can be related to the property of the solid material, Ps, through Eq. (1). Significant variation in mechanical strength and stiffness between
The power index, n, and the constant, C, can be determined experimen- loading directions can also be introduced through design; with func-
tally. The value of n is dependent on the nature of the cellular structure tionally graduated changes in relative density or cell size along the prin-
(open-celled, closed-cell etc.) and the topology. cipal directions [28]. Compression loading along the direction of
 n gradient change can produce a controlled layer-by-layer crushing be-
P ρ haviour while maintaining a shear band crushing when loaded perpen-
¼C ð1Þ
Ps ρs dicular to the gradient direction [23].
To further investigate the mechanical behaviour of gyroid structures,
In the case of mechanical properties such as stiffness and strength, many researchers have used numerical modelling methods. These
lattice structures tend to be categorised by their topology into methods have been used to identify stress distribution, and to extend
bending-dominated and stretch-dominated structures. Gibson and experimental findings by investigating the effect of selected design pa-
Ashby identify foams as bending-dominated structures. The solid- rameters [16]. The structural performance of the gyroid compares well
network gyroid's uniaxial compression response is predominantly with other TPMS structures [29]. Under compression loading, the

Fig. 1. Gyroid forms: triply periodic minimal surface (a); solid-network or lattice form (b), solid-surface, solid-sheet, matrix or cellular form (c) [16].
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

            2
distribution of high stress throughout the unit cell has been identified as 2π 2π 2π 2π 2π 2π 2
sin x cos y þ sin y cos z þ sin z cos x ≤h
a sign of effective load transfer through the structure [29]. Often numer- a a a a a a
ical results are evaluated with Gibson and Ashby's power relationship ð4Þ
(Eq. (1)), highlighting the strong dependence on relative density.
            2
This research uses numerical methods to examine the compression 2π 2π 2π 2π 2π 2π 2
sin x cos y þ sin y cos z þ sin z cos x ≥h
response of the gyroid structure. Through analysis of high stress regions a a a a a a
within the solid-surface gyroid, previously unknown, orientation de- ð5Þ
pendent load carrying substructures are identified. Furthermore, the
paper provides a novel method for locally tuning the thickness of For the gyroid represented by implicit function G(x, y, z) as given by
these substructures to provide bulk anisotropic stiffness and strength Eq. (2), the local surface normal direction, n, is defined by combining the
properties which lead to superior performance through increased mate- partial derivatives of G, as shown in Eq. (6)-Eq. (9).
rial efficiency.
       
∂G 2π 2π 2π 2π 2π 2π
2. Methods ¼ cos x cos y − sin x sin z ð6Þ
∂x a a a a a a

Finite element methods are used to investigate the compression re-        


∂G 2π 2π 2π 2π 2π 2π
sponse of the solid surface gyroid. All simulations were performed using ¼ cos y cos z − sin y sin x ð7Þ
∂y a a a a a a
the Abaqus implicit solver (Simulia). The following three subsections
describe: the representative equations, a finite element voxel model,        
∂G 2π 2π 2π 2π 2π 2π
and a more computationally efficient shell model. ¼ cos z cos x − sin z sin y ð8Þ
∂z a a a a a a
2.1. Gyroid representation
∂G ∂G ∂G
n ¼ ∇Gðx, y, zÞ ¼ iþ jþ k ð9Þ
The minimal surface of the gyroid can be approximated by the level ∂x ∂y ∂z
set equation (Eq. (2)), with the isovalue, h, set to h = 0 [30]. The param-
eter, a, is the unit cell length. By selecting alternative values for h be-
tween the limits (±1.413), the surface can be offset along its normal 2.2. Solid finite element model
direction; beyond those limits the surface becomes disconnected, and
ceases to exist for |h| > 1.5 [31]. Introducing an inequality, as in A voxel based finite element (FE) mesh with eight-node hexahedral
Eq. (3), enables the selection of the regions to either side of the shifted elements was generated to approximate the solid-surface gyroid within
surface, producing the solid-network form. The solid-surface form is a cubic bulk form (Fig. 2). Firstly, a triangular mesh was constructed
generated by selecting the region between two surfaces shifted along within Matlab (MathWorks) using the isosurface for h = 0 in Eq. (2).
the surface normal to either side from their gyroid mid-surface (or alter- This mesh was copied and translated along the local normal to create
natively using the inequality in Eq. (4)) [31]. The inequality in Eq. (5) the thickened solid-surface geometry and enclosed at the boundary
produces two interwoven non-connected solid-network forms. edges. A voxelization function was then used to produce the voxel
            mesh [32]. The voxel element size was set to target a minimum of 4 el-
2π 2π 2π 2π 2π 2π ements through the thickness direction.
sin x cos y þ sin y cos z þ sin z cos x ¼h
a a a a a a
Nodes on two opposite faces of the bulk cube were controlled
ð2Þ through kinematic couplings with reference points to approximate
        axial compression between two flat plates with no separation and no
2π 2π 2π 2π
sin x cos y þ sin y cos z sliding. The reference point controlling the Base Nodes was fixed,
a   a   a a while the reference point controlling the Top Nodes was displaced ver-
2π 2π
þ sin z cos x < h ð3Þ tically to reduce the distance between the two reference points, thereby
a a
compressing the voxel model. Furthermore, the reference points

Fig. 2. Voxel model of 3Z3Z3 gyroid. Nodes identified with kinematic coupling to controlled boundary conditions. Inset shows fine mesh scale.

3
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

1400 2.3. Shell based finite element model


1200
Engineering Stress (MPa)
Modelling of complex cellular structures through solid contin-
1000 uum elements becomes computationally expensive as the number
of unit cells increases. Methods used for simplification of models
800 for strut-based lattice structures include representing struts as
beams. The beams require cross-section information along its length
600 obtained from computer aided design (CAD) models or from mea-
surement of manufactured specimens [34,35]. An analogous simplifi-
400
cation method for surface-based structures would be to represent
200 the curvature with shell elements and define a uniform or local var-
iation in thickness.
0 To increase the efficiency of examining non-uniform thickness ap-
0 2 4 6 8 10 12
plied to the solid surface gyroid a shell model was developed in place
Engineering Strain (%) of the solid voxel model; again, kinematic coupling of the Top Nodes
and Base Nodes were used to control the compression. The mid-
Fig. 3. Tensile stress-strain response used for material model. Based on tensile test
surface of the gyroid (Eq. (2)) was used to generate a 3D FE model
specimens from 'conventional' SLM-fabricated Ti6Al4V, 30μm layer thickness, containing
non-equilibrium acicular α' martensite structure [33]. using conventional 4-node (quad), and 3-node (tri) linear shell ele-
ments. Through mesh convergence tests it was determined that shell el-
ement size of 0.024 unit cell length produced converged stiffness and
provide the displacement and reaction force data for determining bulk strength results within 2% of the converged solid voxel results.
compression properties. The ratio of the characteristic length, l, and the thickness, t, is impor-
The material properties were derived from published tensile testing tant for determining the appropriate use of thin shell or thick shell the-
data of Ti6Al4V specimens, additively manufactured via selective laser ory. Thin shell theory, which ignores transverse shear deformation, is
melting (SLM) [33]. The original tensile data (Fig. 3) was in the format recommended for l/t > 15. Thick shell theory, which includes transverse
of engineering stress (s) – engineering strain (e) and was converted to shear deformation, is recommended for l/t < 15. The general shell ele-
true stress (σ) – log plastic strain (εpl) for the purpose of simulation ments available in Abaqus apply thick shell theory as the shell thickness
using Eq. (10) and Eq. (11). The lack of local cross-section area measure- increases and become thin shell elements as the thickness decreases.
ments during necking implies that the accuracy of the model decreases The proportions of the gyroid structures considered in this study have
beyond the peak load in the material curve. unit cell size of 1 and, using the unit cell length as the characteristic
length, the ratio l/t varying between 10 and 100.
σ ¼ sð1 þ eÞ ð10Þ

σ
ε pl ¼ ln ð1 þ eÞ− ð11Þ 3. Results
E

Compression properties for the bulk material were obtained through The following subsections describe the high stress regions observed
simulation of a solid cube under the same boundary conditions as the in the voxel model, relate these regions to the local surface normal and
gyroid model (Fig. 2). These bulk material compression properties apply the findings to the shell model to explore tuneable performance
(Table 1) were then used to determine relative stiffness and strength. and material efficiency.

Fig. 4. Spiral or helix structure load path within solid-surface gyroid, observed in a finite element voxel mesh from alternate view directions.

4
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

Fig. 5. Helical substructures for gyroid surfaces aligned with the principal directions; x-direction [100] (a), y-direction [010] (b), and z-direction [001] (c). View from [010].

Fig. 6. Comparison of surface alignment to loading direction [001] and stress magnitude. Focussing on results from the central unit cell in a 3Z3Z3 solid-surface gyroid structure under
compressive loading. Surface alignment with load direction is indicated by the dot product b b (a), von Mises stress on voxel mesh (b), chart of element stress versus inclination-load dot
vn
product (c). Stresses are largest in regions where the surface normal is approximately perpendicular to the load direction (b b≈0), and decrease as the surface normal becomes aligned
vn
with the load direction (vbnb ≈1).

5
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

Uniform Thickness 0.025 Z - Helix 0.05 3.1. Structure within a 3D voxel model
X - Helix 0.05 Y - Helix 0.05
XYZ - Helix 0.05 Uniform Thickness 0.05 A 3Z3Z3 gyroid structure was simulated using a solid voxel finite
80 element mesh. The model was loaded by compressing two opposite
Engineering Stress (MPa)

faces, through the linear elastic region, past the yield point, and beyond
60 the compression strength. At a displacement within the linear elastic re-
gion of the response, the elements with von Mises stress greater than
40 half the materials yield stress were identified (Fig. 4). The majority of
these elements lie within several isolated regions, each of these regions
appear as a spiral or helix between the compressed faces.
20
Loading along any one of the three principal axes of the gyroid struc-
ture (eg. [001], [010], [100]) induces stresses predominantly within he-
0
0 0.02 0.04 0.06 0.08 lical substructures aligned with that loading direction. These helical
Engineering Strain (mm/mm) regions, aligned to loading along the principal axis, are evenly spaced
in a square grid and adjacent helices alternate in handedness (or chiral-
Fig. 7. Chart showing stress-strain response of a 3Z3Z3 gyroid nominal uniform thickness
ity). These substructures appear to provide the most direct load path
(0.025 unit length), and the effect of increased thickness (0.05 unit length) applied to the through the structure.
helices associated with the three principal axes (X, Y and Z).

(a)
Uniform (0.025U) Dual-Thick Dist. Thick Dist. Thick
Thickness (U)
Rao 1.0 Rao 2.52 Rao 2.5 Rao 5.0

(b)
Fig. 8. Reinforcing effect of redistributing material to the helical substructures. Normalised compression strength and stiffness at fixed relative density of 0.077 (a). Thickness distribution
contours for select thickness ratios (b). Distributed thickness based on stress-inclination relationship shows significantly greater improvement than the sharp transition of the dual-
thickness distribution across a ± 10° helical band.

6
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

Fig. 9. Gyroid substructures aligned with loading direction. Identifying alternating helices in square grid arrangement along primary axis [001] direction (a), Identifying alternating helices
in hexagonal grid arrangement along body diagonal [111] direction (b), and an intermediate connected structure along [110] direction (c). Structures are viewed along loading direction in
each case.

3.2. Load carrying sub-structures 3.3. Exploration of helix reinforcing effects using shell models

The regions of the gyroid surface which are parallel with the load are To explore the effects of thickening the helical subregions, the
expected to be more efficient at carrying the load. These regions can be modelling shifts to shell elements, where thickness variation can be
identified in the shell model. Shell elements are selected and grouped explored efficiently. Three sets of helices were identified by shell ele-
based on their alignment with the load direction. The selection is ments near perpendicular (80° ≤ α ≤ 100°) to the principal X, Y and Z di-
achieved by determining the angle, α, of the shells unit normal, n b , rela- rections. Compression loading is applied in the Z-direction. For a 3Z3Z3
tive to the load direction, b
v (Eq. (12)). Shell elements with normal ap- celled gyroid structure, with nominal thickness of 0.025 unit length, the
proximately perpendicular to the loading direction are expected to effect of doubling the thickness of the helical substructures for each of
dominate the stiffness response. the principal axes are displayed in Fig. 7. Doubling the thickness on
the Z-aligned helix, increases the peak load by approximately 25% com-
 
bn
α ¼ arccos v b ð12Þ pared to the original. The same increase on the X- and Y-aligned helices
increase the peak under compression loading aligned with the Z-axis by
Fig. 5 shows the shell elements that have been selected within the approximately 5%. Doubling the thickness of the helical regions for all
range 80° ≤ α ≤ 100° from each of the principal X, Y and Z directions, three directions in the same model increased the peak load by 37%.
for a gyroid defined by Eq. (2). These substructures appear as helices The reinforcing effect of the helices was also explored at a constant
similar to those identified by large von Mises stress in the voxel model. relative density. Balancing increased material at the load aligned helices
The stress distribution from the solid voxel model was extracted with material removed from the remaining structure highlights possible
along with the coordinates of the voxel centroid. The normal vector efficiency gains. Concentrating material within the helical band (de-
(n) for these voxels was then calculated based on the surface gradient fined by 80° ≤ α ≤ 100°), showed stiffness and strength improving to ap-
(Eq. (9)). The alignment with the loading direction is then found by proximately 6% and 10.5% respectively at a thickness ratio of 2.5. Further
increase in thickness ratio beyond 2.5 caused a reduction in perfor-
the dot product between the unit loading direction (bv) and the unit nor-
mance Fig. 8. The gains made by reinforcing the helices appears to be
mal vector (n b).
counteracted by reduced support from surrounding geometry and per-
Comparison of stress with surface inclination was made by isolating
haps stress concentrating effects caused by the increasing step change
the central unit cell within the 3Z3Z3 voxel model in order to reduce
in thickness at the edges of the helices.
boundary effects. The results, displayed in Fig. 6, clearly show higher
In addition to specifying distributions with two distinct thickness
local von Mises stress for the surfaces with normal perpendicular to
regions (Dual-thick), a continuous thickness distribution (Dist-thick),
the load direction, and lower local von Mises stress for surfaces with
based on the stress-inclination relation determined from the voxel
normal parallel to the load direction. Furthermore, the stress measured
model (Eq. (13)), was applied with increasing max/min thickness
at the helix is up to six times that measured elsewhere in the structure.
ratio. These results show a 22% increase in strength and 18% in-
The correlation between surface inclination and stress results are
crease in stiffness from the nominal uniform thickness and across
considered with a best fit curve based on a cosine relationship on the
a larger range of thickness ratio. A summary of these data appear
scalar product of surface normal and load direction vectors (Eq. (13)).
in Table 2.
This analysis was completed for a stress state within the linear-elastic
region corresponding to an effective strain of 0.005 and effective stress
of 27.2 MPa. 3.4. Alternative loading directions
   
StressVM b b ¼ A cos πb
v n b þB
v n
ð13Þ As the loading direction shifts from the principal X-, Y- and
A ¼ 258:6, B ¼ 361:6, R2 ¼ 0:8402, RMSE ¼ 69:24 Z-directions the substructures (defined by the shell elements normal
to the loading direction) take alternative forms, in particular:

Table 1
Experimental tensile properties and simulated bulk compression properties.

Experimental Tensile Test (from [33] 30\um PBF) Simulated Compression Test of Solid Cube (Relative Density = 1)

Elastic Modulus (E) 113,800 MPa Modulus (0.5% strain) 124,190 MPa
Ultimate Strength (true stress) 1,200 MPa Max Compressive Strength (engineering stress, first peak) 1,287 MPa

7
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

• The [001] loading direction (and other principal axes [100], [010]) has alternative directions will likely shed light on the orientation specific
a square arrangement of the spirals, with adjacent spirals rotating in differences in the structures response.
the opposite direction (Fig. 9a). The association of helices with the gyroid are not entirely new.
• The [111] loading direction (and other body diagonals [111],̅ [111], ̅ Previously, the solid-network gyroid has been approximated by
̅
[111]) has a hexagon arrangement of spirals with adjacent spirals ro- intersecting helices aligned with the three orthogonal principal di-
tating in opposite directions (Fig. 9b). rections (X,Y,Z). The intention, to simplify modelling the complex
• The [110] loading direction has an intermediate connected structure gyroid topology defined by the level set (Eq. (3)) [6]. There was no
(Fig. 9c). speculation of alternative substructure arrangements. By compari-
son, the current paper identifies different helical structures to
those of the solid-network model simplification. Here the helices
If the loading orientation is shifted continuously between the pri- are understood as load paths in the solid-surface gyroid, they alter-
mary axis and body diagonal the helices distort, join into a larger con- nate in chirality and their arrangement are particular to specific
nected structures and split again to a form a new set of helices aligned loading directions.
to different primary axis or body diagonal. The identification of the substructures representing the
Simulation of voxel models for the three cases identified in Fig. 9 re- dominant load path could also inform material location in multi-
produce the same substructures as the most highly stressed regions. The material additive manufacturing applications. The material provid-
simulations also indicate that the stiffness and strength along these al- ing the structural resistance would be concentrated at the load
ternate directions are different from those in the primary X, Y and Z di- path substructure and a second material could conform to the gyroid
rections. These results highlight that the solid-surface gyroid structure shape, providing, for instance, a fluid interface that takes advantage
is not isotropic in response to loading from alternative directions. of the gyroids superior permeability [8].
Fig. 8 showed limitations to improving efficiency by shifting material
4. Discussion from the bulk of the gyroid solid-surface to concentrate at the helical
substructures in two distinct thicknesses, while maintaining a fixed
In this research, it was shown that under macroscopic uniaxial load- mass. The overall structure weakens as the thickened helix receives
ing, the solid-surface gyroid distributes loads predominantly through less support from surrounding material. Advantages may come from
substructures which are load orientation dependent. These substruc- retaining thickness at other linking substructures to provide supporting
tures manifest as arrangements of helices for the loading directions as- interconnections between the main load carrying substructures rather
sociated with the principal axes and the cubic body diagonals. than isolating them.
Furthermore, by adjusting the thickness of the load carrying regions, The substructures, identified in this research for load path align-
the response can be tuned in specific directions. ment, could also be considered for build orientation manufacturability.
The compression response of the solid-network gyroid structure is The AM process allows flexibility to fabricate the complex cellular struc-
also orientation dependent [23]. Using numerical methods, the relative tures, but also has inherent behaviours which introduce non-uniformity
stiffness was observed to range between 0.0057 at [001] and 0.008 at to the fabricated part. Local geometry influences the thermal response,
[111], with the 〈110〉 family of directions providing an intermediate which in turn influences the surface. Inclined geometries are often pro-
stiffness. Across a volume fraction range of 5–15% the stiffness variation duced with increased thickness compared to geometries more aligned
between the three selected directions at the same volume fraction was with the build direction. The method used to identify the dominant
consistently measured at approximately 30%. By comparison, the solid- load carrying sub-structures can also identify regions that align pre-
surface gyroid examined here at lower relative density of 7.7%, showed dominantly with the build direction in AM, regions expected to be fab-
relative stiffness of 0.022 along the [001] loading direction, which in- ricated with tighter tolerances. Regions with surface inclined to the
creased by approximately 20% at the same relative density through build direction and tending to accumulate additional material could
the use of the locally distributed thickness. Further investigation of the also be identified by adjusting the selected angle range (Eq. (12)). The
substructures associated with loading the surface-based gyroid along geometry could then be tuned to accommodate the expected variation

Table 2
Relative Density, Stiffness and Strength of simulated gyroid structures as compared to the compression simulation of the bulk material properties. (U represents unit cell length).

Description Unit Cell Repetitions Helix/Web Thickness Element Size Relative Density Relative Stiffness Relative Strength

Solid voxel model 3Z3Z3 0.05 U nominal 0.03125 (1/32)U 0.155 0.043 0.050
Shell uniform 0.05 U thickness 3Z3Z3 0.05 U 0.024 U 0.155 0.044 0.050
Shell uniform 0.05 U thickness 3Z3Z3 0.05 U 0.048 U 0.155 0.044 0.051
Shell uniform 0.05 U thickness 3Z3Z3 0.05 U 0.096 U 0.156 0.044 0.052
Shell uniform 0.025 U thickness 3Z3Z3 0.025 U 0.024 U 0.0773 0.0204 0.0227
Shell Z-helix, Dual-thick 2.0 3Z3Z3 0.05 U / 0.025 U 0.024 U 0.0863 0.0243 0.0286
Shell X-helix, Dual-thick 2.0 3Z3Z3 0.05 U / 0.025 U 0.024 U 0.0863 0.0218 0.0240
Shell Y-helix, Dual-thick 2.0 3Z3Z3 0.05 U / 0.025 U 0.024 U 0.0863 0.0219 0.0238
Shell XYZ-helices, Dual-thick 2.0 3Z3Z3 0.05 U / 0.025 U 0.024 U 0.1016 0.0271 0.0312
Shell Z-helix, Dual-thick 1.45 3Z3Z3 0.03454 U / 0.02375 U 0.024 U 0.0773 0.0212 0.0243
Shell Z-helix, Dual-thick 1.96 3Z3Z3 0.04409 U / 0.0225 U 0.024 U 0.0773 0.0216 0.0249
Shell Z-helix, Dual-thick 2.52 3Z3Z3 0.05363 U / 0.02125 U 0.024 U 0.0773 0.0216 0.0251
Shell Z-helix, Dual-thick 3.16 3Z3Z3 0.06317 U / 0.02 U 0.024 U 0.0773 0.0214 0.0250
Shell Z-helix,Dual-thick 3.88 3Z3Z3 0.07271 U / 0.01875 U 0.024 U 0.0773 0.0210 0.0245
Shell Z-helix, Dual-thick 4.70 3Z3Z3 0.08825 U / 0.0175 U 0.024 U 0.0773 0.0204 0.0234
Shell Z-helix, Dist-thick 1.5 3Z3Z3 0.0304 U / 0.0203 U 0.024 U 0.0773 0.0220 0.0253
Shell Z-helix, Dist-thick 2.0 3Z3Z3 0.0341 U / 0.0170 U 0.024 U 0.0773 0.0228 0.0265
Shell Z-helix, Dist-thick 2.5 3Z3Z3 0.0368 U / 0.0147 U 0.024 U 0.0773 0.0233 0.0271
Shell Z-helix, Dist-thick 3.0 3Z3Z3 0.0388 U / 0.0129 U 0.024 U 0.0773 0.0237 0.0274
Shell Z-helix, Dist-thick 3.5 3Z3Z3 0.0404 U / 0.00115 U 0.024 U 0.0773 0.0239 0.0275
Shell Z-helix, Dist-thick 4.0 3Z3Z3 0.0417 U / 0.0104 U 0.024 U 0.0773 0.0240 0.0277
Shell Z-helix, Dist-thick 4.5 3Z3Z3 0.0427 U / 0.0095 U 0.024 U 0.0773 0.0241 0.0278
Shell Z-helix, Dist-thick 5.0 3Z3Z3 0.0436 U / 0.0087 U 0.024 U 0.0773 0.0242 0.0278

8
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

with build inclination angle, to produce a more uniform thickness or to Acknowledgements


influence mechanical properties.
The helical substructures were seen in solid-surface gyroid struc- The authors acknowledge use of facilities within the RMIT
tures at relative densities of 15%, 30% and 60%. As the solid-surface is Advanced Manufacturing Precinct. This research was conducted by the
thickened, the pores shrink and the load path through the material is Australian Research Council Industrial Transformation Training
less contorted. The regions with highest von Mises stresses tend to Centre in Additive Biomanufacturing (IC160100026). http://www.
join and lose their resemblance of isolated helices. Rather than additivebiomanufacturing.org.
considering von Mises stress, the stress field can be analysed for
the highest compressive stresses and the highest tensile stresses. References
The highest compressive stresses tend to follow the inner side
[1] A.H. Schoen, Infinite periodic surfaces without self-intersections. NASA TN D-5541,
of the helix, while the highest tensile stresses tend to occur at the Federal Scientific and Technical Information, Springfield, VA, 1970.
outer side of the helix as well as alternating between the upper [2] S. Rajagopalan, R.A. Robb, Schwarz meets Schwann: design and fabrication of bio-
and lower surface of the horizontal regions. Further work could ex- morphic and durataxic tissue engineering scaffolds, Med. Image Anal. 10 (5)
(2006) 693–712.
plore these observations in more detail.
[3] S.T. Hyde, M. O’Keeffe, D.M. Proserpio, A short history of an elusive yet ubiquitous
5. Conclusions structure in chemistry, materials, and mathematics, Angew. Chem. Int. Ed. 47 (42)
(2008) 7996–8000.
[4] K. Michielsen, D.G. Stavenga, Gyroid cuticular structures in butterfly wing scales:
Gyroid structures are typically deployed with an assumed constant biological photonic crystals, J. R. Soc. Interface 5 (18) (2008) 85–94.
wall thickness or with a thickness gradient along a primary spatial di- [5] Z. Almsherqi, F. Margadant, Y. Deng, A look through ‘lens’ cubic mitochondria, Inter-
rection. However, this numerical investigation of peak structural face Focus 2 (5) (2012) 539–545.
[6] K. Monkova, et al., Three Approaches to the Gyroid Structure Modelling as a Base of
stresses identifies previously unknown substructural regions of local Lightweight Component Produced by Additive Technology, DEStech Transactions on
stress concentration. The shape of these regions and their arrangement Computer Science and Engineering, 2017 cmsam.
within the solid-surface gyroid structure morph with the orientation of [7] V.J. Challis, et al., High specific strength and stiffness structures produced using
selective laser melting, Mater. Des. 63 (Supplement C) (2014) 783–788.
the loading direction. In particular the substructures take the form of
[8] F.P.W. Melchels, et al., Effects of the architecture of tissue engineering scaffolds on
helices arranged with square grid spacing and alternating chirality in cell seeding and culturing, Acta Biomater. 6 (11) (2010) 4208–4217.
the three primary cubic directions; while for the four cubic diagonal di- [9] C. Yan, et al., Ti–6Al–4V triply periodic minimal surface structures for bone implants
rections they appear as helices with hexagon grid spacing, again with al- fabricated via selective laser melting, J. Mech. Behav. Biomed. Mater. 51 (Supple-
ment C) (2015) 61–73.
ternating chirality between adjacent helices. This initial research [10] A.A. Zadpoor, Additively manufactured porous metallic biomaterials, J. Mater. Chem.
demonstrates the feasibility of enhancing the mechanical efficiency of B 7 (26) (2019) 4088–4117.
solid-surface gyroid structures by locally increasing wall thickness at [11] S. Ehrig, et al., Surface tension determines tissue shape and growth kinetics, Sci. Adv.
5 (9) (2019) p. eaav9394.
these identified regions using either a dual-thickness approach or a con-
[12] C.M. Bidan, et al., Geometry as a factor for tissue growth: towards shape optimiza-
tinuous thickness distribution. tion of tissue engineering scaffolds, Adv. Healthcare Mater. 2 (1) (2013) 186–194.
Reduced order numerical simulations, for the dual-thickness ap- [13] D.W. Abueidda, et al., Effective conductivities and elastic moduli of novel foams with
proach, show that adding material at the helical substructures increases triply periodic minimal surfaces, Mech. Mater. 95 (2016) 102–115.
[14] S. Khaderi, V. Deshpande, N. Fleck, The stiffness and strength of the gyroid lattice,
the absolute stiffness and strength of the gyroid structure. However an Int. J. Solids Struct. 51 (23–24) (2014) 3866–3877.
optimal point can be found, when redistributing material to the sub- [15] S.C. Kapfer, et al., Minimal surface scaffold designs for tissue engineering, Biomate-
structures whilst maintaining the same overall relative density (or vol- rials 32 (29) (2011) 6875–6882.
ume fraction), limiting stiffness and strength. [16] E. Yang, et al., Effect of geometry on the mechanical properties of Ti-6Al-4V Gyroid
structures fabricated via SLM: A numerical study, Mater. Des. (2019) 108165.
The relationship between local von Mises stress and solid-surface in- [17] A. Ataee, et al., Anisotropic Ti-6Al-4V gyroid scaffolds manufactured by electron
clination to loading along the primary axis [001] direction was evalu- beam melting (EBM) for bone implant applications, Mater. Des. 137 (Supplement
ated and attributed a fit based on the cosine function. Scaling this fit C) (2018) 345–354.
[18] A. Yanez, et al., Compressive behaviour of gyroid lattice structures for human can-
function as a thickness distribution to the solid-surface gyroid resulted
cellous bone implant applications, Mater. Sci. Eng. C Mater. Biol. Appl. 68 (2016)
in approximately 20% increase in stiffness and strength at a fixed rela- 445–448.
tive density of 7.7%. [19] L.J. Gibson, M.F. Ashby, B.A. Harley, Cellular Materials in Nature and Medicine,
The specification of thickness distribution through local surface orien- Cambridge University Press, 2010.
[20] L.J. Gibson, M.F. Ashby, Cellular Solids: Structure and Properties, Cambridge univer-
tation represents a novel method for tuning the performance of the sity press, 1999.
gyroid and other TPMS cellular structures to the expected load conditions. [21] I. Maskery, et al., Effective design and simulation of surface-based lattice structures
This developed methodology provides a basis for further work, including: featuring volume fraction and cell type grading, Mater. Des. 155 (2018) 220–232.
[22] R. Tino, et al., Gyroid structures for 3D-printed heterogeneous radiotherapy phan-
1. Future investigations could refine the material distribution around toms, Phys. Med. Biol. 64 (21) (2019) 21NT05.
the helical substructures to tune the stiffness response in multiple [23] L. Yang, et al., Investigation on the orientation dependence of elastic response in
Gyroid cellular structures, J. Mech. Behav. Biomed. Mater. 90 (2019) 73–85.
directions. [24] L. Zhang, et al., Energy absorption characteristics of metallic triply periodic minimal
2. Integration of thermal field prediction methods to accommodate the surface sheet structures under compressive loading, Add. Manufact. 23 (2018)
influence of manufacturing process on local thicknesses – thereby 505–515.
[25] F.S.L. Bobbert, et al., Additively manufactured metallic porous biomaterials based on
allowing optimisation of as-manufactured geometry.
minimal surfaces: a unique combination of topological, mechanical, and mass trans-
port properties, Acta Biomater. 53 (2017) 572–584.
Data availability [26] A. du Plessis, S.M.J. Razavi, F. Berto, The effects of microporosity in struts of gyroid
lattice structures produced by laser powder bed fusion, Mater. Des. 194 (2020)
108899.
The raw/processed data required to reproduce these findings cannot [27] C.N. Kelly, et al., Fatigue behavior of as-built selective laser melted titanium scaffolds
be shared at this time as the data also forms part of an ongoing study. with sheet-based gyroid microarchitecture for bone tissue engineering, Acta
Biomater. 94 (2019) 610–626.
[28] M. Zhao, et al., Mechanical and energy absorption characteristics of additively
manufactured functionally graded sheet lattice structures with minimal surfaces,
Declaration of Competing Interest Int. J. Mech. Sci. 167 (2020) 105262.
[29] D.W. Abueidda, et al., Mechanical properties of 3D printed polymeric Gyroid cellular
The authors declare that they have no known competing financial structures: experimental and finite element study, Mater. Des. 165 (2019) 107597.
[30] C.A. Lambert, L.H. Radzilowski, E.L. Thomas, Triply periodic level surfaces as models
interests or personal relationships that could have appeared to influ- for cubic tricontinuous block copolymer morphologies, Philos. Trans. 354 (1715)
ence the work reported in this paper. (1996) 2009–2023.

9
D. Downing, A. Jones, M. Brandt et al. Materials and Design 197 (2021) 109096

[31] M.R.J. Scherer, Double-Gyroid-Structured Functional Materials, Springer Theses, 1 [34] B. Lozanovski, et al., A Monte Carlo simulation-based approach to realistic modelling
ed, XXI, Springer International Publishing 2013, p. 198. of additively manufactured lattice structures, Add. Manufact. 32 (2020) 1–21.
[32] A.H. Aitkenhead, Mesh Voxelisation, MATLAB Central File Exchange https://www. [35] L. Liu, et al., Elastic and failure response of imperfect three-dimensional metallic lat-
mathworks.com/matlabcentral/fileexchange/27390-mesh-voxelisation 2020. tices: the role of geometric defects induced by selective laser melting, J. Mech. Phys.
[33] W. Xu, et al., Additive manufacturing of strong and ductile Ti-6Al-4V by selective Solids 107 (2017) 160–184.
laser melting via in situ martensite decomposition, Acta Mater. 85 (2015) 74–84.

10

You might also like