You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/338736198

Performance of Triply Periodic Minimal Surface Lattice Structures Under


Compressive Loading for Tissue Engineering Applications

Conference Paper  in  AIP Conference Proceedings · January 2020


DOI: 10.1063/5.0001631

CITATION READS

1 1,719

3 authors:

Farouq Hisham Khogali Hui Leng Choo


Taylor's University Taylor’s University
1 PUBLICATION   1 CITATION    45 PUBLICATIONS   120 CITATIONS   

SEE PROFILE SEE PROFILE

Wei Hsum Yap


Taylor's University
48 PUBLICATIONS   423 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

A Technological Licensing Framework for 3D Printed Content: A Focus on China AHRC CENTRE FOR DIGITAL COPYRIGHT AND IP RESEARCH IN CHINA (Ningbo, China) View
project

3D-printing and characterisation of bone scaffold by Poly(vinyl) alcohol with natural hydroxyapatite obtained from whitefin wolf herring and Spanish mackerel fish-bones
View project

All content following this page was uploaded by Hui Leng Choo on 22 January 2020.

The user has requested enhancement of the downloaded file.


Performance of Triply Periodic Minimal Surface Lattice
Structures Under Compressive Loading for Tissue
Engineering Applications
Elfaroug Hosham Khogalia1, Hui Leng Choo1,a) and Wei Hsum Yap2
1
School of Engineering, Taylor’s University, No. 1 Jalan Taylor’s, 47500, Subang Jaya, Malaysia
2
School of Biosciences, Taylor’s University, No. 1 Jalan Taylor’s, 47500, Subang Jaya, Malaysia
a)
Corresponding author: HuiLeng.Choo@taylors.edu.my

Abstract. Triply Periodic Minimal Surfaces have been a major research topic in the field of tissue engineering. The interest
stems from their theorized ability to serve as functional tissue scaffolds. Tissue scaffolds need to fulfil certain criteria for
successful integration. It is important for the scaffold to possess similar mechanical properties to the surrounding tissue
where the scaffolds would be implanted to avoid stress shielding. This article discusses the evaluation of the compressive
mechanical properties of four types of TPMS lattice structures which are the Primitive, Diamond, Gyroid, and I-WP lattices.
Each of the lattices were tested under two relative densities. The first set of relative densities ranged between 29% - 30%.
The second set was comprised of relative densities ranging between 48% - 53%. The compressive properties studied were
the compressive modulus, compressive strength and yield strain to determine which of the lattices possess the most
favourable mechanical performance. The comparison of the lattices also featured a Cubic lattice structure to determine
whether TPMS lattices perform better under compression than the Cubic lattice structure. A comparison was then performed
between the obtained mechanical properties of the lattices and some of the known mechanical properties of several human
tissue types. The study revealed that under both the tested relative densities, the Cubic lattice illustrated the highest
mechanical performance while the I-WP lattice demonstrated the lowest mechanical properties for both relative densities.
Additionally, it was seen that the Primitive, Diamond, Gyroid, and I-WP lattices with lower relative density resemble the
mechanical properties of bone tendons and Cancellous bone. Lastly, an error analysis was performed on the relative density
of the manufactured lattices to determine the suitability of the Fused Deposition Modelling (FDM) method for manufacture
of functional tissue scaffold and the mechanical properties were compared to existing literature. The mechanical properties
of the lattices agreed with the published literature. FDM proved to be suitable for manufacturing the lattices though
improvements must be made for optimum results. Improvements can be integrated through usage of a smaller nozzle
diameter or by adjusting the printing settings such as the nozzle speed or temperature.

INTRODUCTION
Human tissue possesses the ability to naturally regenerate and heal in the case of injury. However, in some instances
medical intervention is needed to assist in the healing process and to prevent further damage to the injured tissue. The
field is relatively new but promises great potential in the near future. Tissue engineering is identified as a
multidisciplinary effort to create effective alternatives to rehabilitate or enhance regeneration of damaged tissue [1].
Adopting tissue engineering techniques as a means of treatment would be considered as an upgrade from currently
adopted methods such as allografts and autografts due to their limitations. Autografts involve tissue transfer in one
region to another from the same person, while allografts involve transplanting tissue from another individual. For
instance, autografts are expensive and can result in great pain to the patient. While allografts are limited by the body’s
immune system’s potential rejection of the donated tissue, imposing the risk of infection which leads to deterioration
of the injured tissue [2]. Due to such reasons, new methods should be adopted to avert these adverse symptoms.
One of the more recent applications in the field of tissue engineering involves utilisation of three-dimensional
scaffolds with controllable physical properties. For instance, metallic scaffolds can be used in prosthetic devices. In
addition, they hold potential to function as constructs that assist the body in regeneration of tissues and organs [3]. For
tissue scaffolds to be successfully employed in tissue regeneration applications, they need to fulfil certain requirements.
Ideally, a tissue scaffold should possess a porous topology to allow for the transport of necessary nutrients for cell
growth. Additionally, the scaffold must be biocompatible to avoid the risk of rejection by the bod’s defence
mechanisms. The scaffold must be biodegradable such that when the tissue heals, the scaffolds would degrade in
accordance to the healing rate of the body without leaving any toxic by-products to avoid infection. The scaffold should
also possess similar mechanical properties as well as internal architecture to the surrounding tissue where the scaffold
will be implanted. The strength of the scaffolds should ideally be enough to withstand handling during the implantation
process [4 , 5]. One factor that makes TPMS lattices candidates for tissue scaffolds is the fact that they exist in nature
[6]. Their topology is more likely to resemble the complex internal architecture of human tissue.
In addition to their architecture, several studies that investigate the mechanical properties of Triply Periodic
Minimal Surface (TPMS) lattice structures indicate that TPMS structures may possess the necessary mechanical
properties to serve as tissue scaffolds. The most important mechanical property for the scaffolds to resemble in human
tissue is the compressive modulus. Studies have utilized several manufacturing methods to produces the lattices such
as Stereolithography, Selective Laser Melting (SLM), Power Bed Fusion, and Selective Laser Sintering (SLS). Table
1 summarizes the reported compressive modulus of several TPMS lattices obtained under compression testing. The
table also summarizes the additive manufacturing technologies and materials utilized to produce the lattices.

TABLE 1. Summary of reported findings from five publications describing the compressive modulus, relative density,
manufacturing method, and material of the lattices

Compressive Relative Manufacturing


Reference
Structure Modulus (MPa) Density (%) Method Material
257 30 Stereolithography Photopolymer resin [7]
Primitive 3100 25 Powder Bed Fusion Maraging Steel [8]
192 30 SLS Polyamide PA2200 [9]
355 22 FDM Ceramic Paste [10]

83 29 Stereolithography Photopolymer resin [7]


Diamond 2178 28 SLM Porous Titanium [5]
3400 25 Powder Bed Fusion Maraging Steel [8]
93 30 SLS Polyamide PA2200 [9]
283 22 FDM Ceramic Paste [10]

Gyroid 112 30 Stereolithography Photopolymer resin [7]


3500 25 Powder Bed Fusion Maraging Steel [8]
95 30 SLS Polyamide PA2200 [9]
51.7 19 FDM Ceramic Paste [10]

I-WP 72 29 Stereolithography Photopolymer resin [7]


3200 25 Powder Bed Fusion Maraging Steel [8]

Cubic 3691 27 SLM Porous Titanium [5]

According to the data in table 1, metallic scaffolds possess a much higher compressive modulus than polymeric
scaffolds with the same relative density as expected. For instance, the photopolymer resin Gyroid scaffold with a
relative density of 30% demonstrated a compressive modulus of 112 MPa in comparison to a compressive modulus of
3.5 GPa with the Maraging steel Gyroid scaffold. This illustrates that even though relative density is a key factor in
determining the lattices’ mechanical properties, the material plays a vital part in contributing to the resulting
mechanical properties. Furthermore, since the Stereolithography created lattices possessed virtually the same relative
density, a fair comparison of their compressive modulus can be performed. The comparison of the created lattices
shows that the Primitive lattice recorded the highest stiffness (257 MPa), while the I-WP lattice recorded the least
stiffness (72 MPa).

OBJECTIVES OF STUDY
This study aimed to examine and compare four different TPMS lattice structures, namely, the Primitive, Diamond,
Gyroid and I-WP Surface lattice structures, in terms of their mechanical properties. Even though a great number of
TPMS exist, the shape defining mathematical equations of only Primitive, Diamond, Gyroid, and I-WP structures are
well described in literature. Several studies investigated and compared the Primitive, Diamond, Gyroid, and I-WP
lattices however, no study compared the mechanical performance of all four TPMS lattices in addition to the cubic
lattice. Therefore, these four TPMS lattices were selected for this study and compared to the cubic lattice. The
comparison was performed to determine which topology of the four TPMS lattices leads to the highest stiffness,
compressive strength, and compressive strain of the lattices in addition to determine whether the TPMS lattices will
outperform the strut-based lattice (cubic lattice). The obtained data was then compared to the known mechanical
properties of some human tissues to determine if any of the lattices’ mechanical properties resemble that of the tissues.
The mechanical properties were determined through compression testing.
Another aim of this study was to evaluate the accuracy of the additive manufacturing technology Fused Deposition
Method (FDM) which was used to produce the lattices. The evaluation was conducted to determine the suitability of
FDM as a cheaper method to other additive manufacturing technologies utilised in other studies such as
Stereolithography, SLS and SLM. The TPMS lattices were additively manufactured from biocompatible PLA as a trial
material. It is worth noting that this study is purely structural and does not consider factors like the material of which
the lattices are made permeability of the structures.

METHODOLOGY
TPMS structures are defined by mathematical equations – referred to as triply periodic level sets equations – that
govern their architecture. Each of the structures (Primitive, Diamond, Gyroid, and I-WP surface) possesses its own
unique equation which can be seen from equations 1-4 [6]:

Schwartz Primitive (P):


…‘•ሺ‫ݔ‬ሻ ൅ …‘•ሺ‫ݕ‬ሻ ൅ …‘•ሺ‫ݖ‬ሻ  ൌ ‫ݐ‬ (1)
Schwartz Diamond (D):
•‹ሺ‫ݔ‬ሻǤ •‹ሺ‫ݕ‬ሻǤ •‹ሺ‫ݖ‬ሻ ൅ •‹ሺ‫ݔ‬ሻǤ …‘•ሺ‫ݕ‬ሻǤ …‘•ሺ‫ݖ‬ሻ ൅ …‘•ሺ‫ݔ‬ሻǤ •‹ሺ‫ݕ‬ሻǤ …‘•ሺ‫ݖ‬ሻ ൅ …‘•ሺ‫ݔ‬ሻǤ …‘•ሺ‫ݕ‬ሻǤ •‹ሺ‫ݖ‬ሻ  ൌ ‫ݐ‬ (2)
Schoen’s Gyroid (G):
…‘•ሺ‫ݔ‬ሻǤ •‹ሺ‫ݕ‬ሻ ൅ …‘•ሺ‫ݕ‬ሻǤ •‹ሺ‫ݖ‬ሻ ൅ …‘•ሺ‫ݖ‬ሻǤ •‹ሺ‫ݔ‬ሻ  ൌ ‫ݐ‬ (3)
I-WP Surface (IW):
…‘•ሺ‫ݔ‬ሻǤ …‘•ሺ‫ݕ‬ሻ ൅ …‘•ሺ‫ݕ‬ሻǤ …‘•ሺ‫ݖ‬ሻ ൅ …‘•ሺ‫ݖ‬ሻǤ …‘•ሺ‫ݔ‬ሻ  ൌ ‫ݐ‬ (4)
where x, y, and z define the spatial coordinates of the structures, while the variable t directly influences the relative
density of the unit cells.
Relative density represents the ratio of the solid regions in the structure to the total volume occupied by unit
cells. The volume fraction is an important factor to be considered as they dictate the mechanical properties of the
TPMS structures. When a unit cell is replicated to form a lattice structure, the relative density of the unit cell will be
conserved. It is must be noted that an equal t value does not correspond to the same relative density between the four
TPMS lattices. Therefore, unit cells with t values ranging from 0 to -14 were used to vary the relative density of the
unit cells and the two most similar relative densities across all four TPMS unit cells were selected to be additively
manufactured for testing. The reason for the selection of this range of t values lies in the fact that when a value of t
lower than – 14, the unit cells lose their structural integrity. An example of the varying relative densities of the Diamond
unit cell can be seen in figure 1
(a) (b) (c)

FIGURE 1. Diamond lattice at (a) t=-4, (b) t=-6, (c) t=-8

As can be seen, as the t value decreases, the thickness of the unit cell struts decreases accordingly, resulting in
lower relative density. For creation of the cubic lattice, Solid Works was used, and no equations were used to create
the lattice therefore, the relative density of the Cubic lattice is not characterized by the t parameter as with the TPMS
lattices.
FDM was used to manufacture the lattices. FDM involves heating a thermoplastic filament (Biocompatible PLA
in the case of this study) until it reaches its melting point. The PLA is dispensed from a nozzle at 215 ° C with a speed
that ranges between 40 to 90 mm/s. The nozzle moves in the x, y, and z axes to create the desired geometries. Layers
of the PLA which are 0.2 mm in height are extruded and fused together when they cool down.
For analysis of the accuracy of FDM to create the lattices, an error analysis between the theoretical relative density
(ρ*) and experimental relative density of the lattices was performed. First, the theoretical volume of the lattices was
obtained using the additive manufacturing software Netfabb. The following equation (5) was then used to determine
the theoretical relative density of the lattices. In this study the total volume of the lattices was 25 ൈ 25 ൈ 25 mm3.
௏௢௟௨௠௘௢௙ௌ௢௟௜ௗ
ߩ ‫כ‬ൌ  (5)
்௢௧௔௟௏௢௟௨௠௘௢௙௅௔௧௧௜௖௘
The Archimedes principle was used to measure the experimental relative density of the lattices. A 100 ml measuring
cylinder was filled with 50 ml of distilled water. The lattices were then submerged in the water (one at a time) and the
displacement of the water volume was measured. The difference between the displacement and the original volume
was then obtained as the actual volume of the lattices in millilitres. The volume was then converted to the appropriate
units and equation (5) was used to obtain the experimental relative density.
To analyse the mechanical properties of the lattices, compression testing was performed on the samples. The
compression test is used to determine the lattices’ properties because this study aims to simulate the conditions that
tissue scaffolds would be subjected to in the human body. The compression testing was performed under the ASTM
(D1621). Lattices with a 5x5x5 unit cells with a total volume of 25 × 25 × 25 mm3. were produced. Additionally, at
least five samples of each specimen were tested as dictated by the standard. The specimens were conditioned by being
placed at least 24 hours prior to testing in the lab where the testing was performed. The testing was performed using
an Instron 5965 universal material testing frame equipped with a 25 kN load cell with a crosshead speed of 2.5 mm/min.
The test was performed until the specimens reached a strain percentage of 13% where key properties such as the
compressive stress were determined.

RESULTS AND DISCUSSION


The samples were successfully printed with no need for additional support during printing. The lattices with both
relative densities can be seen in Figure 2. Table 2 demonstrates a comparison between the expected relative density of
the lattices against the experimental relative density with the error analysis to determine the accuracy of the FDM
method. The naming convention in the table uses the initial letter of the lattice name followed by its corresponding t
value. For example, a Primitive lattice with t value = -3 would be referred to as P3. The results showed varying
accuracies across all the lattice and no specific trend was observed. The highest error was recorded by the Primitive
lattice (ɏȗൌ 30%) at 15% error followed by the Diamond lattice (ρ*=29%) and the I-WP (ɏȗൌ 50%) at 10.6% and
10.4% respectively. Whereas the Primitive (ɏȗൌ 50%) and Diamond (ɏȗൌ 48%) lattices recorded significantly low
error percentages with 1.4% and 0.5% respectively.
t=6 t = 10 t=0 t=5

(a) (b)

t=0 t=6 t=0 t=7

(c) (d)

ɏȗ = 53% ɏȗ = 29%

(e)

FIGURE 2. The (a) Primitive, (b) Diamond, (c) Gyroid, (d) I-WP, and (d) Cubic lattices with both relative densities

TABLE 2. Error analysis of experimental theoretical density

Experimental ɏȗ
Lattice Theoretical ɏȗ 1 2 3 Error %
P3 50.4 51.3 51.24 50.8 1.4
P9 30.8 35 35.1 36.4 15.3
D0 48.1 47.7 48.1 47.8 0.5
D5 28.8 32.3 31.9 31.4 10.6
G0 49.3 46.7 46.9 46.7 5.1
G6 30 31.6 31 31.2 4.2
I-WP1 50 55.3 55.2 55.1 10.4
I-WP8 29.5 31.7 31.8 31.3 7.1
C53 53 55.9 56.8 56.9 6.7
C29 29 31.7 31.9 31.9 9.8

The extracted mechanical properties were the compressive modulus, compressive stress, and yield strain of the
lattices. The obtained results were tabulated, and the data was grouped according to the relative densities of the lattices
in Table 3.
TABLE 3. Mechanical Properties of TPMS lattices with ɏȗ= 48-53% and ɏȗ = 29-30 %

Elastic Modulus (MPa) Compressive Strength (MPa) Yield Strain (%)


Lattice Type ɏȗ = 48-53 % ɏȗ = 29-30 % ɏȗ = 48-53 % ɏȗ = 29-30 % ɏȗ = 48-53 % ɏȗ = 29-30 %
Primitive 1285±74 597±32 15.15±1.78 5.76±0.18 1.63±0.08 1.52±0.12
Diamond 1308±68 475±8 11.77±0.28 5.22±0.04 1.59±0.14 1.66±0.08
Gyroid 1619±25 554±21 14.85±0.20 5.47±0.05 1.54±0.12 1.73±0.04
I-WP 1270±198 220±11 5.99±3.41 2.59±0.32 0.80±0.38 1.58±0.23
Cubic 1930±161 749±61 19.85±1.58 6.79±0.38 1.64±0.14 1.38±0.07

As illustrated in Table 3, the highest elastic modulus and compressive stress recorded were attained by the Cubic
lattice (ɏȗ = 53 %) at 1930 MPa and 19.85 MPa respectively. It is worth noting that all four TPMS lattice types
experienced failure around very similar strain percentages except for IW which recorded the least strain percentage.
Furthermore, the Gyroid lattice possessed the second highest stiffness, but the maximum compressive strength did not
correspond to the Compressive modulus and was lower than the compressive strength of the Primitive lattice. The I-
WP lattice illustrated the lowest performance amongst all as it possessed the lowest elastic modulus, compressive
strength, and failed after being subjected to the least strain amongst all four lattice types. The Primitive, Diamond, and
Gyroid lattices failed at similar yield strain percentages; the cubic lattice failed after the greatest yield strain while the
I-WP lattice failed after the least strain percentage. Figure 3 presents the stress-strain graphs of the lattice structures
and gives a further understanding of the deformation behaviour of the lattices.
Reducing the relative density (ɏȗ) of the lattices incurs significant changes in the mechanical performance of the
lattice structures as can be seen in Table 3. All TPMS lattices illustrated a great reduction in the compressive modulus,
as well as the compressive strength. Furthermore, the cubic lattice was also observed to possess the highest mechanical
performance amongst the lattices. Indicating that at lower relative densities, a strut-based topology is more desirable.
The Gyroid lattice also performed well in a lower density, while the I-WP lattice also demonstrated the least mechanical
performance. Figure 3 illustrates the stress-strain graphs of the TPMS lattices with ɏȗ = 48-53%

45
40
A B
35
30
Stress(MPa)

P3
25
D0
20
G0
15
IW1
10
C (ρ*=53%)
5
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Strain(mm/mm)

FIGURE 3. Stress-Strain graphs comparing the five TPMS lattice types (ɏȗ = 48-53%)

The Diamond, Gyroid, Cubic and I-WP lattices exhibited very similar behaviour as can be seen from Figure 3.
Their graphs start with the elastic region after which plastic deformation begins. The plastic deformation region
represents a layer-by-layer collapse in which the initial response in the stress-strain graphs indicates a slight drop in
compressive stress required to deform the lattices. End of the elastic region for the Diamond, Gyroid, and I-WP lattices
were all observed at around the same strain percentage. The dotted line A on figure 3 illustrates the end of elastic
region for all the lattices occurring at very similar strain percentages. The drop in stress can be explained by the failure
of the mid-section layer in the lattices. The stress drops slowly but consistently until all four graphs experience a very
sharp increase in the compressive stress at higher strain or core deformation percentages. This increase in stress can
be attributed to the densification stage that the lattices undergo as the collapsed layers become very closely packed
together and illustrates higher resistance to the compressive stress. Dotted line B (Figure 3) illustrates densification
starts for the Diamond, Gyroid, and I-WP lattices at approximately 5% strain. While for the Cubic lattice, where the
layers deform almost entirely in unison, densification begins slightly sooner.
The Primitive lattice illustrated a slightly different deformation than the other four lattices. Furthermore, the
deformation behaviour did include the elastic region in the same manner as the other lattices but there was no observed
drop in compressive stress at the onset of plastic deformation. Instead, the compressive stress seemed to increase after
the end of the elastic region. This can be explained by the fact that contrary to the four other lattices, the Primitive
lattice had all the layers deforming in unison and no layer deformed more than the other at the end of the elastic region.
Therefore, densification starts earlier in the Primitive lattice than with the other lattices. Primitive lattice deformation
was governed by a stretching dominated behaviour where the struts can be seen to buckle at higher strains whereas at
the yield strain, it seems that there is little observable deformation.
8

6
Stress(MPa)

5 P9
4 D5

3 G6

2 IW8
C (ρ*=29%)
1

0
0 0.02 0.04 0.06 0.08 0.1 0.12
Strain(mm/mm)

FIGURE 4. Stress-Strain graphs comparing the four TPMS lattice types (ɏȗ = 29-30%)

Figure 4 illustrates the stress-strain of the lower relative density lattices. All TPMS lattices can be observed to
behave in a very similar manner to their higher relative density counter-parts. The same stages can be observed in the
stress-strain graph as in Figure 3. The only difference between the two graphs lies in the magnitude of drop in stress
at the onset of plastic deformation where the drop in compressive stress is much sharper than with the higher relative
density lattices. Additionally, the difference between the I-WP lattice in comparison to the other lattices was much
greater than under a high relative density. The stress-strain graph of the cubic lattice illustrates a continuous drop in
and elevation in compressive stress. The drop in stress signifies the collapse of one of the lattice layers, followed by
the densification as the layers become more packed into a smaller region. The effects of deformation are more
pronounced in the lattices at a lower relative density. All five lattices fail at approximately the same strain percentage
(1.3%-1.6%). The Primitive lattice maintained the early densification at the onset of plastic deformation. Whereas the
Diamond, Gyroid, and I-WP lattices began densification at around 6% strain. Figures 5, 6, and 7 demonstrate the
deformation behaviour of the Primitive, Diamond, and Gyroid lattices at the yield strain and at 5% strain. The Cubic
lattice on the other-hand demonstrated densification slightly later than the other lattices at 7% strain. The recurring
“bumps” in the stress-strain graph of the cubic lattice can be attributed to the subsequent failure in each of its layers.
Figure 5, 6, and 7 present the deformation behaviour of the Diamond, Gyroid, and Primitive lattices (both relative
densities) respectively. The Diamond and Gyroid lattices showed a bending dominated behaviour in which the
Diamond lattice showed diagonal shear bands that are more visible at 5% strain as can be seen in Figures 5 and 6
highlighted with aid of the blue dotted lines. The Gyroid lattice represented the stretching dominated behaviour through
strut bending. The dotted lines Figure 6 illustrate the stretching of the of the Gyroid struts. Whereas the Primitive
lattice illustrated a stretching dominated behaviour in which the struts buckle at higher strain percentages. This can be
clearly observed in Figures 7c. and 7d. At the Yield strain of all the presented lattices, little to no deformation can be
observed. The deformation becomes much more apparent at 5% deformation. Notably, the lower density lattices
presented very similar deformation to the higher density lattices. The difference being how much more visible the
deformation becomes at lower relative densities. The cubic lattices illustrated an identical behaviour to the Primitive
lattices due to their similar structures while the I-WP lattices illustrated a similar deformation behaviour to the Gyroid
and Diamond lattices.

(a) (b) (c)

(d) (e) (f)

FIGURE 5. Deformation stages of Diamond lattice (ɏȗ=48%). at (a) 0% Strain, (b) 1.59% Strain, and (c) 5% Strain. and
deformation of Diamond lattice (ɏȗ=29%) at (d) 0% Strain, (e) 1.67% Strain, and (f) 5% strain.

(a) (b) (c)

`
(d) (e) (f)

FIGURE 6. Deformation stages of Gyroid lattice (ɏȗ=49%). at (a) 0% Strain, (b) 1.54% Strain, and (c) 5% Strain. and
deformation of Diamond lattice (ɏȗ=30%) at (d) 0% Strain, (e) 1.74% Strain, and (f) 5% strain
(a) (b) (c)

(d) (e) (f)

FIGURE 7. Deformation stages of Primitive (ɏȗ=50%) lattice at (a) 0% Strain, (b) 1.31% Strain, and (c) 5% Strain. and
deformation of Primitive (ɏȗ=30%) lattice at (d) 0% Strain, (e) 1.14% Strain, and (f) 5% Strain

The error analysis of the lattices’ experimental relative density (Table 2) illustrates that even though the Primitive
and Diamond lattices (ɏȗ=50%, and ɏȗ= 48%) respectively showed remarkable accuracy, several other lattices
demonstrated high levels of inaccuracy that attained error percentages of up to 15%. There were no observed patterns
in the high error percentages. For instance, the high complexity of lattices such as the Diamond and Gyroid did not
show a direct co-relation with the elevated inaccuracy percentages as the Diamond lattice with lower relative density
recorded the least error amongst all the lattices. Furthermore, there was also no co-relation between the relative density
of the lattices and the error percentages. Therefore, the errors can be attributed to both the FDM method as well as the
3D printer used. The errors present themselves in the form of additional strings of the PLA filament that were not
detached after manufacture of the lattices was complete and may have affected the physical attributes of the lattices.
The pores were observed to not be smooth but rather bumpy and inconsistent. Another factor that may have contributed
to the inaccuracies is the rough surface finish of the printer. Therefore, it is recommended that the FDM is tested with
more accurate 3D printers that possess smaller nozzle diameters and under different settings such as the nozzle speed.
The data obtained from the compression test sheds light on whether TPMS lattices can be utilised in the human
body as tissue scaffolds. For instance, the high compressive strength recorded by lattices with 48%-50% relative
density illustrates their potential for use in high stiffness applications. The lower Compressive modulus of the I-WP
lattice does not render it ineffective as a potential tissue scaffold but rather serves as an indicator that it can be used in
applications where lower stiffness is required. Furthermore, analysis of the lower relative density lattices indicated that
Cubic lattice illustrated the highest compressive modulus amongst all lattices. In terms of the mechanical properties,
it was apparent that the cubic lattice was superior to the TPMS lattices under both the higher and lower relative
densities. This superiority can be attributed to their rectangular struts that are known for their mechanical performance
as opposed to circular struts as observed in the Primitive lattice. In addition, it can be said that the higher performance
of the Gyroid and Diamond lattices to the Primitive lattice is due to the high interconnectivity of the structures which
is much more pronounced in the higher relative density setting. With lower relative density, it seems that the strut
geometry is more suitable to produce lattices with higher stiffness as well as larger pore sizes. This was evident as the
top two performing lattices under compression were the Cubic and Primitive lattices. This can be attributed to the more
efficient load transfer mechanism of the Primitive lattice due the round structure at the joints. However, it is important
to highlight that even though the cubic lattice did record the highest mechanical performance, they may not be suitable
for use as scaffolds due to their simple topology that may not resemble that of surrounding human tissues in addition
to their characteristic sharp edges which can damage the surrounding tissues.
To further understand the similarity between TPMS lattice mechanical properties and human tissue, mechanical
properties of several human tissues in Table 4 were compared to those of the TPMS lattices. Table 4 states that the
Cancellous bone can possess a wide range of stiffness from 0.02-0.5 GPa. It was found that the Diamond and I-WP
(with relative densities 29%) lattice structures possess stiffness magnitudes that reside within that range. Additionally,
the tendon stiffness lies within the range of 0.065 – 0.541 GPa which coincides with the stiffness of the Diamond, I-
WP, and Cubic lattices also with a relative density of 29%-30%. The lattices with higher relative density are much
higher than the tissues referenced in the above figure, meaning that those lattices would be needed in much higher
stiffness applications such as in dental applications. Additionally, most of the tissues in the figure possess much lower
stiffness than the tested lattices. To attain similar mechanical properties, further studies are to be conducted to tested
lattices with lower relative densities or perhaps even other types of TPMS lattice structure.

TABLE 4. Mechanical properties of several human tissues

Tissue Type Young's Modulus References


Cancellous bone 0.05-0.5 GPa [12]
Cortical bone 7-30 GPa [12]
Tendon 0.56 GPa [13]
Subcutaneous adipose tissue 2.9 kPa [14]
Lung 1-5 kPa [15]
Liver 5 kPa [16]
Stomach 1.9 kPa [16]

The results obtained from literature summarised in Table 1 included metallic TPMS lattices which explains the
large differences in the compressive modulus of the lattices created in this study. Furthermore, when the results from
this study are compared with literature in which the lattices were created from Polymers, the lattices in this study
possessed a much higher compressive modulus. For instance, the Primitive lattice in this study demonstrated a
compressive modulus of 597 MPa while when created from photopolymer resin using Stereolithography [7], the
Primitive lattice demonstrated a compressive modulus of only 257 MPa. The same trend was observed for the Diamond
and Gyroid lattices. This difference can be attributed to the difference in properties of PLA and the photopolymer
resin. Similar results were observed when FDM was employed to create the lattices from ceramic paste [10], the
difference between the compressive modulus of the Primitive lattice was recorded at 355 MPa. The differences can be
attributed to the difference in material properties. Furthermore, the results from a study that compared the performance
of a Cubic lattice to the Diamond lattice [5] showed that the Cubic lattice outperformed the Diamond lattice in the
compression test which agrees with the results obtained in this study.
The results from this study agree with the results obtained from literature, bearing in mind the differences of
material properties used to manufacture the lattices. This indicates that FDM holds a true potential to be utilized as a
cheaper alternative to manufacture biocompatible tissue scaffolds. However, as observed in the error analysis (Table
2), there were errors attained in the experimental relative density in comparison to the expected relative density.

CONCLUSIONS
Compression tests showed that under both relative density ranges, the cubic lattice outperformed TPMS lattices.
This however does not indicate that TPMS lattices are not suitable as Table 4 illustrated that they have similar
mechanical properties to some human tissues. Upon comparison of the lattices’ mechanical properties with that of
human tissues, it was observed that lattices with higher relative densities proved to be much stiffer than the reviewed
human tissues. While with lower relative density the Diamond, Gyroid and I-WP proved to suitable for use in two
tissue types (Cancellous bone and Tendon) in terms of the similarity in mechanical performance. The cubic lattice
performed much better than expected as it recorded the highest mechanical performance under both relative densities.
Further testing of different TPMS lattices with a wider range of relative densities by utilizing all the t values in the
permissible spectrum (t = 0 to t = -14) using FDM must be performed in a comprehensive manner to further clarify the
understanding of which lattices can be utilized as tissue scaffolds. For better understanding of the lattices’ ability to
perform as functional tissue scaffolds, in vitro studies can be performed to determine their ability to support cell growth
and proliferation. The tests can be utilised to determine which of the structures lead to optimum cell growth.
Furthermore, the error analysis of the relative density of the scaffolds illustrated that there is still room for
improvement in the technology to attain optimum results for integration of the manufactured lattices as functional
tissue scaffolds.
REFERENCES
1. N. Hauptmann, Q. Lian, J. Ludolph, H. Rothe, G. Hildebrand, and K. Liefeith, “Biomimetic Designer
Scaffolds Made of D , L -Lactide- e - Caprolactone Polymers by 2-Photon Polymerization,” vol. 25, no. 3,
pp. 167–186, 2019.
2. F. J. O. Brien, “Biomaterials & scaffolds for tissue engineering,” Mater. Today, vol. 14, no. 3, pp. 88–95,
2011.
3. G. Criscenti, C. De Maria, G. Vozzi and L. Moroni, “Characterization of Additive Manufactured Scaffolds,”
3D Printing and Biofabrication, pp. 55-78, 2018.
4. J. Shi, L. Zhu, L. Li, Z. Li, J. Yang, and X. Wang, “A TPMS-based method for modeling porous scaffolds for
bionic bone tissue engineering,” Sci. Rep. 8, 7395, 2018.
5. J. Kadkhodapour, H. Montazerian, A. C. Darabi, and A. P. Anaraki, “Failure mechanisms of additively
manufactured porous biomaterials : Effects of porosity and type of unit cell,” J. Mech. Behav. Biomed. Mater.,
vol. 50, pp. 180–191, 2015.
6. L. Han and S. Che, “An Overview of Materials with Triply Periodic Minimal Surfaces and Related Geometry:
From Biological Structures to Self-Assembled Systems,” Adv. Mater., vol. 30, no. 17, pp. 0935–9648, 2018.
7. X. Zheng, Z. Fu, K. Du, C. Wang, and Y. Yi, “Minimal surface designs for porous materials : from
microstructures to mechanical properties,” J. Mater. Sci., vol. 53, no. 14, pp. 10194–10208, 2018.
8. O. Al-ketan, R. Rowshan, and R. K. A. Al-rub, “Topology-Mechanical Property Relationship of 3D Printed
Strut , Skeletal , and Sheet Based Periodic Metallic Cellular Materials Topology-mechanical property
relationship of 3D printed strut , skeletal , and sheet based periodic metallic cellular materials,” no. March
2019, 2018.
9. I. Maskery, L. Sturm, A.O. Aremu, A. Panesar, C.B. Williams, C.J. Tuck, R.D. Wildman, I.A. Ashcroft and
R.J.M. Hague, “Insights into the mechanical properties of several triply periodic minimal surface lattice
structures made by polymer additive manufacturing,” Polymer, vol. 152, pp. 62–71, 2018.
10. M. Werner, P. Pieranski, J. Yu, and E. Nie, “Mechanical properties of ceramic structures based on Triply
Periodic Minimal Surface ( TPMS ) processed by 3D printing Mechanical properties of ceramic structures
based on Triply Periodic Minimal Surface ( TPMS ) processed by 3D printing,” pp. 8–14, 2017.
11. D. Yoo, “Computer-aided Porous Scaffold Design for Tissue Engineering Using Triply Periodic Minimal
Surfaces,” vol. 12, no. 1, pp. 61–71, 2011.
12. B. Substitutes, N. A. Nawawi, A. S. F. Alqap, and I. Sopyan, “Recent Progress on Hydroxyapatite-Based
Dense Biomaterials for Load Bearing Recent Progress on Hydroxyapatite-Based Dense Biomaterials for Load
Bearing Bone Substitutes,” no. March, 2011.
13. T. Clayton, T. Mckee, J. A, P. Russel, and C. Murphy, “Indentation versus tensile measurements of Young’s
modulus for soft biological tissues,” Tissue Eng. Part B, vol. 17, no. 3, 2011.
14. N. Alkhouli, J. Mansfield, E. Green, J. Bell, B. Knight, N. Liversedge, J.C. Tham, R. Welbourn, A.C. Shore,
K. Kos and C.P. Winlove, “The mechanical properties of human adipose tissues and their relationships to the
structure and composition of the extracellular matrix,” Am J Physiol Endocrinol Metab, vol. 305(12), pp.
1427–1435, 2013.
15. B. Hinz, “Mechanical Aspects of Lung Fibrosis A Spotlight on the Myofibroblast,” vol. 9, pp. 137–147, 2012.
16. W.C. Yeh, Y.M. Jeng, H.C. Hsu, P.L. Kuo, M.L. Li, P.M. Yang, P.H. Lee and P.C. Li, "Young's modulus
measurements of human liver and correlation with pathological findings," 2001 IEEE Ultrasonics
Symposium. Proceedings. An International Symposium, Atlanta, GA, USA, vol. 2, pp.1233-1236, 2001

View publication stats

You might also like