Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Materials & Design 217 (2022) 110613

Contents lists available at ScienceDirect

Materials & Design


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

Design and fabrication of architected multi-material lattices with


tunable stiffness, strength, and energy absorption
Denizhan Yavas, Qingyang Liu, Ziyang Zhang, Dazhong Wu ⇑
Department of Mechanical and Aerospace Engineering, University of Central Florida, Orlando, FL 32816, 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

 The flexural modulus and strength of


the multi-material struts decrease
with the increasing core-to-strut
thickness ratio.
 The energy absorption capacity of the
multi-material struts is greater than
that of the individual constituents.
 Multi-material lattices with tunable
in-plane compression modulus and
strength are achieved.
 The multi-material lattices exhibit
greater energy absorption capacity
compared to those of the individual
constituents.

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

Article history: Modern engineering applications require advanced materials that are stronger, tougher, and lighter.
Received 10 November 2021 Nature offers various structural architectures to achieve superior mechanical properties. Inspired by bone
Revised 21 March 2022 with a spongy soft-core and a compact hard-shell, this study introduces architected multi-material lat-
Accepted 29 March 2022
tices with tunable mechanical properties such as stiffness, strength, and energy absorption. The archi-
Available online 31 March 2022
tected lattices are fabricated via multi-material fused filament fabrication. The hard polylactic acid
(PLA) shell of the lattice struts maintains the stiffness, while the soft thermoplastic polyurethane (TPU)
Keywords:
core enhances toughness and energy absorption capacity. The microstructural and mixed-mode fracture
Architected materials
Lattice structure
characteristics of the PLA-TPU interfaces were examined since the interfacial microstructure and adhe-
Multi-material 3D printing sion are crucial to stress transfer in multi-phase composite materials. This study explores the flexural
Finite element method behavior of the multi-material struts and reveals two distinct failure modes depending on the volume
Soft-hard interfaces fraction of the soft-core through 3-point bend tests. This study demonstrates the bio-inspired design
on a hexagonal planar lattice with a strut thickness-to-length ratio of 0.3. Tunable stiffness, strength,
and energy absorption are achieved by varying the core-to-strut thickness ratio. Compression test results
show that the architected lattices exhibit an energy absorption capacity that is about 2–3 times greater
than that of the constituent single-phase lattices.
Ó 2022 The Authors. 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/).

1. Introduction ited to expand the boundaries of the current material properties.


There is a wide range of strategies to achieve superior and tunable
Advanced engineering applications require stronger, tougher, mechanical properties such as mimicking nature’s materials with
and lighter materials [1]. However, traditional materials are lim- highly sophisticated architectures [2–6]. One approach is to create
micro-architected multi-material structures with a combination of
⇑ Corresponding author. two or more material phases in a unique microstructure, as
E-mail address: Dazhong.Wu@ucf.edu (D. Wu). inspired by biological materials such as bone [7] and nacre [8].

https://doi.org/10.1016/j.matdes.2022.110613
0264-1275/Ó 2022 The Authors. 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. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

For example, combining soft and hard constituents in a ‘‘brick–m core and hard (stiff) shell thermoplastic polymers, as inspired by
ortar” geometry can offer an excellent combination of strength the unique bio-composite architecture of bone uniting soft spongy
and toughness [9]. Alternatively, the concept of lattice materials core (trabecular bone) and hard compact shell (cortical bone) [2].
is another way of mimicking biological materials, in which the dis- 3D printing employing multi-material fused filament fabrication
tribution of porosity is engineered to achieve superior mechanical (FFF) is used to fabricate the proposed architected multi-material
properties [3,4]. Although the lattice materials were widely studied lattice structures as shown in Fig. 1(a). The soft core of the lattice
over the past decade, they have recently been gaining popularity is made of highly stretchable thermoplastic polyurethane (TPU)
with the latest advancements in additive manufacturing (hyper-elastic up to 400–500% strain), while the hard shell is made
techniques. of polylactic acid (PLA) thermoplastic with relatively greater stiff-
To date, there have been several attempts to design and manu- ness and strength (Young’s modulus of E = 2.8GPa and failure
facture lattice materials with high specific strength and toughness strength of rUTS ¼ 52:5MPa). In the proposed sandwich construc-
[10–19]. These lattice materials were mostly fabricated with either tion of the lattice struts, the hard shell sustains the in-plane stiff-
solid or hollow struts, which were made of a single-phase material ness of the lattice, while the soft core provides enhanced
(e.g., polymers, metals, or ceramics). These lattice materials exhib- toughness and energy absorption. In addition, the proposed design
ited either high strength (or stiffness) or high ductility (or energy allows tunable mechanical properties by varying the core-to-strut
absorption) depending on the mechanical behavior of the building thickness ratio c/t. The present study uses a planar honeycomb lat-
material. For instance, ceramic lattices exhibited ultra-high stiff- tice with a hexagonal unit cell and a strut length-to-thickness ratio
ness and strength with a low ductility due to their brittle nature of t=l ¼ 0:3 (t ¼ 3 and l ¼ 10mm) for the proof of concept. As an
[10], while polymeric lattices showed significantly high ductility example, Fig. 1(b) shows an FFF printed 3  3 honeycomb lattice
with limited strength [19]. Therefore, the challenge remains to with c=t ¼ 1=3.
achieve a well-balanced combination of different mechanical prop-
erties in single phase lattice materials. Although there has been an
effort on the development of multi-material lattices combining 2. Method
strength and ductility in the same material by means of hierarchi-
cal topologies [20,21], multi-material designs [22–28], and novel 2.1. 3D printing and sample preparation
cell geometries such as gyroid-type with triply periodic minimal
surfaces [29–31], there is still a huge need for the development An Ultimaker 3 Extended, FFF-based 3D printer with two
of stiffer and tougher lattice materials with high energy absorption. extruders, was used to fabricate specimens. One extruder was used
Especially, multi-material lattice architectures have not been fully to print PLA (MatterHackers, CA) filament with a nominal diameter
explored due to the material- and manufacturing-related limita- of 2.85 mm, where printing temperature was 200 °C and printing
tions. For example, one of the most critical issues related to mate- speed was 60 mm/s. The other extruder was used to print TPU
rial selection is the interfacial compatibility of the constituent 95A (Ultimaker, Netherlands) filament with a nominal diameter
materials. The mechanical properties of multi-material lattices of 2.85 mm, where printing temperature was 223 °C and printing
are significantly associated with the interfacial microstructure speed was 25 mm/s. The build plate temperature was set to
and adhesion between the constituents since stress is transferred 60 °C during the printing process. The layer thickness for both
through the interfaces. Moreover, as an example of the PLA and TPU 95A was 100 lm. XYZ build orientations were
manufacturing-related limitations, a multi-material lattice archi- adapted for all the samples according to the standard terminology
tecture with a polymer core and a metal shell did not allow the for additive manufacturing ISO/ASTM 52921–13 [32]. The chosen
researchers to investigate the effect of a wide range of shell thick- printing orientation allows the lattice struts to have the highest
ness on the mechanical properties due to the limited capability of bending stiffness as the printed filaments are aligned with their
the metal deposition on the core polymer lattice with uniform longitudinal axes.
thickness [23]. The tensile sample geometry was implemented from the stan-
To address these challenges, this study reports an extrusion- dard test method for tensile properties of plastics ASTM D638-14
based 3D printing approach for fabricating architected multi- [33]. To examine the mode I and mode II interlaminar fracture
material lattice materials with tunable mechanical properties such toughness, double cantilever beam (DCB) and end-notched flexural
as modulus, strength, and energy absorption. The struts of the pro- (ENF) test specimens were fabricated in accordance with ASTM
posed lattice material are designed by combining soft (compliant) D7905 [34] and ASTM D5528 [35], respectively. For both DCB

Fig. 1. Overview of the proposed work. (a) 3D schematic of multi-material FFF printing of multi-material honeycomb lattices, (b) An FFF printed 3  3 multi-material
honeycomb lattice with c=t ¼ 1=3 highlighting the details of the cell wall geometry.

2
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

images of the sample surface were recorded simultaneously. The


in-situ optical images of the sample were recorded using a high-
resolution digital monochrome camera of 2448  2048 pixels (FILR
Systems, Chameleon CM3-U3-50S5M-CS USB3) with a 35 mm lens
(Edmund Optics). An image acquisition rate of 1 Hz was employed
in all tests. The collected images were analyzed using 2D DIC soft-
ware Ncorr [38] to measure the full field strain distribution on the
side surfaces of the samples.

2.3.1. Tensile test


The tensile tests were conducted in accordance with ASTM
D638-14 [33] to determine the bulk mechanical properties of the
3D printed PLA and TPU materials. The tensile test samples were
Fig. 2. Schematics of the 3D printed samples, showing the printing orientation with sized as Type-V described in the standard testing method ASTM
respect to the printing bed. D638-14 [33]. Mechanical tensile wedge grips were used in the
tensile tests. The gage length between the grip surfaces was
adjusted to be 30 mm. A relatively low strain rate of 0.001 1/s
and ENF samples, a 1 mm thick TPU layer was printed between two was adopted to eliminate the rate dependent response of PLA
2.5 mm thick PLA adherents with an initial crack. To determine the and TPU. The average axial strain measured by DIC was used to
flexural properties of the multi-material struts, the multi-material confirm the strain evaluated by the crosshead displacement. The
struts with a length of 60 mm and thickness of 3 mm were fabri- measured force and displacement data were used to determine
cated for c=t ratios of 1/6, 1/3, 1/2, 2/3, and 3/4. The multi- the engineering stress–strain curves. The experimentally obtained
material struts were fabricated in compliance with the standard stress-stain curves were fed into the FE model to determine the
test geometry in ASTM D790 [36]. The multi-material 3  3 honey- constitutive models for both PLA and TPU materials.
comb lattices with only three different c/t ratios of 1/6, 1/3, and 2/3
were printed following the standard specimen geometry provided 2.3.2. Interfacial fracture tests
in ASTM D1621-16 [37], because the implemented c/t ratio resolu- All DCB and ENF tests were performed in accordance with the
tion was enough to capture the variation in the compressive prop- standard testing procedures ASTM D5528-13 [35] and ASTM
erties of the multi-material lattices with varying c/t ratio. The D7905-19 [34], respectively. Fig. 3 shows the DCB and ENF test set-
lattice density of about 32% was implemented with the strut ups and sample geometries. The DCB samples were attached to the
thickness-to-length ratio of t=l ¼ 0:3. A schematic of the sample testing frame through the loading blocks located at the end of the
orientations relative to the printing bed is provided in Fig. 2. Each sample arms. The initial crack was 42 mm, which yields a reason-
specimen was duplicated at least 3 times with 100% infill density able initial crack length-to-sample thickness ratio of about 6. The
and only one specimen was printed at a time. The specimens were upper arm of the DCB sample was loaded vertically as shown in
placed in a dryer box to eliminate the impact of humidity before Fig. 3(c). The ENF samples were tested in three-point bending con-
testing. figuration as shown in Fig. 3(d). The distance between the support
pins was adjusted to be 80 mm. A rigid wire with a diamater of
2.2. Microstructural characterization 0.5 mm was inserted between the sample arms to reduce the fric-
tion on the pre-crack surface. The initial crack length was 28 mm,
The microstructure of the PLA-TPU interface was examined which was greater than 70% of the half support length to maintain
using an optical microscope (Leica DM4 B). The 3D printed PLA- a stable crack growth during the tests.
TPU multi-material samples were first mounted in a low viscosity In both DCB and ENF tests, the sample side surfaces were
resin and hardener epoxy system (Buehler Epoxicure). Then, the painted by a white spray painter, which facilitates tracing the
cross-section of the samples was polished with 400, 600, 800, interfacial crack tip. Optical digital images of the interfacial crack
and 1200-grit grinding papers using a polisher (Buehler VibroMet zone were captured with a frame rate of 1 Hz. Approximately
2 Vibratory Polisher). The optical microscope images were exam- 400 images with a resolution of 15 lm/pixel were collected during
ined to evaluate microstructural features such as manufacturing- each test. The crack growth length was measured using a digital
induced voids at the PLA-TPU interface. image processing tool. The crack growth length was extracted with
an error of about 30 lm (i.e. an absolute error of 0.05% in the mea-
2.3. Mechanical and fracture tests sured crack length). The inset images in Fig. 3(c) and 3(d) show
typical crack growth observed under mode I and mode II fracture,
The mechanical tests include (1) tensile test for characterization respectively. All fracture tests were repeated on at least three sam-
of bulk mechanical properties of PLA and TPU materials, (2) inter- ples to determine the statistical variability. For the measured force
facial fracture tests of double cantilever beam (DCB), and end- P, displacement d, and instantaneous crack length a, the mode I
notched flexural (ENF) for characterization of the interfacial mode energy release rate was calculated following the standard data
I and mode II fracture strength and toughness, (3) three-point reduction method provided in ASTM D5528-13 [35] as,
bending test for characterization of the flexural behavior of the
multi-material struts, and (4) compression test for characterization 3 Pd
GI ¼ ð1Þ
of the in-plane compression behavior of the multi-material lattice 2 ba
materials. where b is the sample width. On the other hand, the mode II
All the mechanical tests were carried out under displacement energy release rate was calculated using the corrected beam the-
control using the Shimadzu Autograph AGS-X 10kN universal test- ory[34,39,40] as,
ing frame at room temperature. A ramp displacement profile with
a constant crosshead movement rate of 2 mm/min was applied. 9 P2 ða þ 0:42DÞ2
GII ¼ 2 3
ð2Þ
The applied cross-head displacement, force, and in-situ digital 16 Eb h

3
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

Fig. 3. The interfacial fracture test samples and testing configurations. (a) DCB and (b) ENF sample geometries, (c) DCB and (d) ENF test configurations. The inset images in (c,
d) show the interfacial crack growth profile.

R 0:15
where h is the half sample thickness, E is the effective flexural W ¼ 0 rðeÞde. In addition, the experimentally measured effec-
modulus of the sample, and D is the correction factor for the mea- tive flexural modulus of the 3D printed multi-material struts were
sured crack length, which takes shear deformation into account compared with the analytical estimations. The effective flexural
and was approximated as the half sample thickness, (i.e., D  h) modulus of the multi-material struts was analytically estimated
[41]. The results of the interfacial fracture experiments were also by the classical (i.e., Euler-Bernoulli) beam theory as,
used by the numerical simulations to determine the cohesive trac-
tion separation (TS) curves for mode I and mode II fracture. Icore Ishell
Eeff ¼ Ecore þ Eshell ð5Þ
I I
2.3.3. Three-Point bending tests of Multi-material struts where Ecore and Eshell are the core and shell’s Young’s moduli and
The initial elastic response of lattice materials under in–plane are substituted by the elastic moduli of TPU and PLA materials,
compression or tension is mainly governed by bending behavior respectively (i.e., Ecore ¼ ETPU and Eshell ¼ EPLA ). In addition, I is the
of the cell wall struts. Therefore, a proper understanding of the second moment of area of the entire cross section such that
flexural behavior of the multi-material struts is essential for I ¼ 1=12bt3 , Icore is the second moment area of the core
assessing the in-plane compression behavior of 3D printed multi- I ¼ 1=12bc3 , and Ishell is the second moment area of the shell, which
material lattice structures. For this purpose, we performed 3- can be estimated using the parallel axis theorem. Accordingly, Eq.
point bending tests on the 3D printed lattice struts with varying (5) becomes
core-to-strut thickness ratios. All 3-point bending tests were car- c3  3
ried out in accordance with ASTM D7264-07 [42] to determine 1 tc 3 ðt  cÞðt þ cÞ2
Eeff ¼ ETPU þ EPLA þ EPLA ð6Þ
the flexural modulus, strength, and energy absorption of the 3D t 4 t 4 t3
printed multi-material struts. The measured crosshead displace- While one can utilize a more complicated beam theory such as
ment and force were used to derive the maximum flexural stress Timoshenko beam theory, the classical beam theory is considered
rf and strain ef as, adequate for a first order approximation of the effective flexural
modulus of the multi-material struts in this study.
3PL
rf ¼ ð3Þ
2bt2
2.3.4. In-plane compression test of Multi-material lattice
and The in-plane compression behavior of the 3D printed multi-
material honeycomb lattices was characterized by the standard
6dt
ef ¼ ð4Þ in-plane compression test procedure ASTM C365-16 [43]. The com-
L2 pression loading was introduced through two 8 in. compression
where P is the measured force, d is the mid-span deflection, t is platens. The measured force and displacement were normalized
the sample thickness, b is the sample width, L is the support span by the nominal planar area (10  44 mm2) and height of the lattice
length. The initial slope of the flexural stress–strain curves was samples (46 mm) to calculate the nominal stress and strain,
taken as the effective flexural modulus and the peak stress value respectively. The initial slope of the stress–strain curves was eval-
was adapted as the flexural strength of the 3D printed multi- uated as the in-plane compression modulus and the peak stress
material struts. The energy absorption (i.e., strain energy density, was taken as the strength of the 3D printed multi-material lattices.
W) was calculated by calculation of the area under the flexural The energy absorption (i.e., strain energy density, W) was calcu-
stress–strain curves from ef ¼ 0 to ef ¼ 0:15 such that lated by evaluating the area under the stress–strain curves from
4
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

zero strain e ¼ 0 to the densification strain of about ed ¼ 0:60 such and C 2 ). In addition, it should be noted that both PLA and TPU
Re
that W ¼ 0 d rðeÞde. are assumed isotopic in the numerical simulations. It is widely
accepted that 3D-printed materials are highly orthotopic, signifi-
cantly compliant, and weaker in the printing direction [46,47].
2.4. Numerical modeling
However, in this study, the material orthotropy in the printing
direction was not considered since the bending-dominated defor-
Finite element (FE) simulations of tensile, DCB, ENF, and 3-point
mation of the examined lattices is mainly controlled by the in-
bend tests were performed to numerically examine the experimen-
plane mechanical properties. Therefore, it is reasonable to employ
tal results. The FE models were developed in a commercial FE soft-
isotropic material models in the numerical simulations. The mate-
ware ABAQUS 6.13 [44]. The multi-material strut and lattice
rial parameters of the utilized constitutive models were calibrated
sample geometries were modeled with 4-node continuum plane
using the experimentally obtained stress–strain curves. The cali-
stress quadrilateral elements with hourglass control (CPS4), while
brated material models were used in the numerical simulations
the fracture sample geometries were modeled with 4-node contin-
of DCB, ENF, and 3-point bend tests.
uum plane strain quadrilateral elements with hourglass control
The DCB and ENF test models employing the cohesive zone
(CPE4). The geometry and boundary conditions of each test were
model (CZM) were used to extract the interfacial mode I and mode
exactly adapted in the corresponding FE simulations. A conver-
II fracture strength ( r b, b
s ) and toughness (CI ; CII ) of the PLA-TPU
gence study was also performed for each model to verify the
interface, respectively. The model geometries and applied bound-
mesh-independency of the results. The tensile test geometry was
ary conditions of DCB and ENF models are presented in Fig. 4.
modeled to extract the constitutive material models of PLA and
One of the interfaces between TPU layer and PLA substrates was
TPU materials. An elasto-plastic constitutive material [45] model
modeled using cohesive interface to simulate the interfacial crack
with isotropic hardening was used to model the mechanical
growth, while the other was assumed to be perfectly tied. Such a
response of PLA. The yield strength was governed by a Von-Mises
modeling strategy was implemented based on the experimental
type isotropic plasticity model. On the other hand, a continuum
observations. A refined mesh size of 100 lm was used in the cohe-
mechanics-based hyper-elasticity material model such as
sive zone. The refined mesh size was chosen to be sufficiently
Mooney-Rivlin [41] was used for TPU. For rubber-like TPU, under
smaller than the cohesive zone length [48–50].
uniaxial tension, the true stress – nominal strain behavior can be
The cohesive interface was governed by the bilinear cohesive
expressed as
traction-separation curves [50]. The mode I and mode II cohesive
   
r ¼ 2 C 1 þ C 2 ð1 þ eÞ1  ð1 þ eÞ2  ð1 þ eÞ1 ð7Þ traction-separation (TS) curves contain three independent fracture
parameters: the initial slope of the curves ðK I ; K II ), the critical cohe-
While there are several alternative material models for model- b, b
sive stresses ( r s ), and cohesive fracture energy (CI ; CII ). For the
ing hyper-elastic response of the soft TPU such as Neo-Hookean or sake of simplicity, there is no coupling between mode I and mode
Arruda-Boyce models, the Mooney-Rivlin model was selected II fracture within initial linear deformation of the cohesive surface
because of its better prediction efficiency within the moderate and the initial slopes of mode I and mode II TS curves were taken as
strains (100–200%) with only two material parameters (i.e., C 1 equal to each other (i.e., K I ¼ K II ). The initial slopes of TS curves

Fig. 4. The finite element models employed for the simulations of (a) DCB, (b) ENF, and (c) flexural bending tests showing the details of the geometry and boundary
conditions in the numerical simulations, (d) a close-up view of the PLA-TPU interfaces highlighted as d in (a-c) showing the refined mesh along the cohesive interface between
the PLA and TPU materials, (e) the cohesive traction-separation curves of the PLA-TPU interface for mode I and mode II fracture.

5
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

were taken to be proportional to the adhesive modulus-to- ing prior to the fracture. In addition, the fracture planes of the PLA
thickness ratio such that K I ¼ K II ¼ bE=h. Here, b is a scaling con- samples were found to be normal to the axial stress direction. On
stant much greater than unity and is taken as b = 100 [50], which the other hand, the stress–strain curve of the TPU sample shows
yields K I ¼ K II  105 MPa=mm: A quadratic nominal traction crite- a strain-dependent stiffness within the initial portion (e < 20%),
rion was employed to evaluate the mixed-mode critical cohesive then reveals almost a linear monotonic deformation above 25%
strength for damage nucleation, while the critical mixed-mode strain. While the entire stress-stain curve of TPU was not shown
fracture toughness for the damage evolution was governed by in Fig. 5, the failure stress was recorded as 14.8 MPa at 440% strain.
the linear power-law criteria in the cohesive surface. The fracture The FE simulations with the calibrated material constants were
strength and toughness of the mode I and mode II TS curves were able to reproduce the experimental stress–strain curves with high
independently determined by matching the experimentally mea- accuracy. A summary of the material parameters of the employed
sured force–displacement and crack resistance curves from the constitutive laws is given in Table 1.
DCB and ENF experiments, respectively. A parametric study was
carried out to fine tune the cohesive fracture parameters. 3.2. Microstructural characterization of PLA-TPU interface

3. Results and discussion The morphology of the interface between 3D printed PLA and
TPU was examined. Fig. 6(a) shows a 3D schematic of the 3D
3.1. Mechanical behavior of PLA and TPU printed PLA-TPU sample used in the microstructural characteriza-
tion. The PLA-TPU interface was inspected on two different planes:
Tensile tests were performed to characterize the mechanical (1) a side surface that is parallel to the build direction and (2) a top
properties of the 3D printed PLA and TPU. The experimental surface that is orthogonal to the build direction. The PLA-TPU inter-
stress–strain curves were used to calibrate the parameters of elas- face morphology on the side surface shows a saw-tooth shaped
tic–plastic and Money-Rivlin hyper-elastic constitutive models for waviness with presence of interfacial voids as shown in Fig. 6(b).
PLA and TPU, respectively. Fig. 5 shows the experimentally The saw-tooth shaped interfacial morphology can be attributed
obtained engineering stress–strain curves of PLA and TPU, along to the limited x-y stage precision of the 3D printer. The limited pre-
with the numerically derived counterparts. The experimental cision results in small swings in the location of the interface along
stress–strain curves are selected from three different tensile tests the build direction, which causes the formation of a zig-zag pat-
and represent the mean mechanical responses. The stress–strain tern. In fact, such a rough interfacial profile can be beneficial for
curve of PLA shows a linear-elastic response up to about 2.5–3% manufacturing multi-material materials with enhanced interfacial
strain and yielding takes place at about 3% strain and 50 MPa adhesion by means of mechanical interlocking. The size of the
stress. Then, it shows a strain-softening behavior after the yield interfacial voids can be as large as 125 lm  250 lm as shown
point and the ultimate failure happens at about 6% strain and in Fig. 6(c). In addition, Fig. 6(d) shows the PLA-TPU interfacial
44 MPa stress. The PLA tensile samples showed a very minor neck- morphology on the top surface. Unlike the interfacial morphology
on the side surface, the PLA-TPU interface on the top surface
reveals a straight interfacial morphology with presence of bulk
voids in the TPU layer near the interface. The close-up views of
the interface presented in Fig. 6(e) and 6(f) show an intact inter-
face, with a series of bulk voids about 40–50 lm away from the
interface. The size of the bulk voids was assessed to be about
250 lm  50 lm.

3.3. Fracture behavior of PLA-TPU interface

The interfacial mode I and mode II fracture behavior of the PLA-


TPU interface was examined via DCB and ENF tests to quantita-
tively assess interfacial adhesion between the constituent materi-
als PLA and TPU. Fig. 7(a) shows the experimentally measured
force–displacement curves obtained from two DCB samples along
with the numerically obtained counterparts. The force–displace-
ment curves typically show a linear trend until the interfacial
opening stress reaches the critical level. Thereafter, the force
reduces monotonically with the commencement of the interfacial
Fig. 5. The experimentally obtained engineering stress–strain curves of PLA and crack growth. One of the force–displacement curves exhibits a sec-
TPU materials, along with the numerically derived counterparts. The inset images
ondary peak force. This can be attributed to arrest of interfacial
show the numerically derived principal strain distributions in PLA and TPU tensile
samples at 6% and 100% axial strain, respectively.
crack because of locally enhanced adhesion at the PLA-TPU
interface.
The corresponding mode I fracture energy release rate curves
derived by the beam theory (Eq. (1)) are shown in Fig. 7(c) as a
Table 1 function of the crack length (i.e., mode I interfacial crack resistance
Material constants of elastic–plastic and Mooney-Rivlin hyperplastic constitutive curves). The energy required for the crack nucleation was found to
models of PLA and TPU, respectively.
be about GI ¼ 40J=m2  60J=m2 . One of the DCB samples showed a
Material Elastic Plastic decreasing crack resistance curve, while the other exhibited a
PLA E = 2800 MPa, ry = 50.1 MPa, ep =0rUTS = 52.5 MPa, modest increasing crack-resistance curve. The steady-state level
m=0.45 ep =0.06 of the energy release rate was measured to be about
Mooney-Rivlin Hyper-elasticity
GIc ¼ 45J=m2 , which was taken as the interfacial mode I fracture
TPU C 2 = 4.39 MPa C 1 = 1.15 MPa
toughness (CI ¼ GIc ). In addition, the FE simulations of the DCB test
6
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

Fig. 6. (a) A schematic of the 3D printed PLA-TPU material showing the build direction and planes of the microscopic inspection, (b) optical microscope image of the PLA-TPU
interface on the side surface highlighting the saw-tooth shaped interfacial morphology and a large interfacial void, (c) a close-up view of area c shown in (b), (d) optical
microscope image of the PLA-TPU interface on the top surface highlighting the straight interfacial morphology with presence of bulk voids in the TPU layer near the interface,
(e) a close-up view of area e in (d), and (f) a close-up view of area f in (e) showing the intact interface and bulk void next to the interface.

Fig. 7. The experimentally measured force–displacement curves obtained from (a) DCB and (b) ENF tests and the corresponding interfacial crack resistance curves for (c)
mode I and (d) mode II interfacial fracture. Blue/red colors denote two different experimental data. The dashed lines in all graphs show the numerically derived counterparts
with the upper and lower bounds of cohesive fracture strength and toughness. (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this article.)

7
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

Table 2 (i.e., mode II interfacial crack resistance curves) derived from the
Mode I and mode II fracture strength and toughness of the examined PLA-TPU corrected beam theory (Eq. (2)). Both ENF samples exhibited
interface.
increasing crack resistance rate curves with a crack nucleation
Parameter Mode I Mode II energy about GII ¼ 150J=m2  300J=m2 . The increasing trend can
Initial slope of traction- K I ¼ 105 MPa=mm K II ¼ 105 MPa=mm be related to the frictional sliding of the interfacial crack surfaces,
separation curve which contributes to the system energy. The upper and lower
Cohesive strength b ¼ 1.0 0:2MPa
r b
s ¼ 2:7  0:5MPa bound of the mode II fracture parameters were derived by FE sim-
Cohesive toughness CI = 48  10 J/m2 CII = 220 7 0 J/m2
ulations of the ENF test. A good agreement between the experi-
mentally and numerically obtained results was observed as
shown in Fig. 7(b) and 7(d). The average of the numerically
extracted mode II fracture parameters and the standard deviation
were performed to numerically extract the upper and lower bound are also provided in Table 2. Noteworthy is that the mode II frac-
of the mode I fracture parameters by matching the experimentally ture strength and toughness of the examined PLA-TPU interface
measured force–displacement and crack-resistance curves[59,60]. were found to be 3–4 folds of the mode I counterparts. This dis-
The numerically derived force–displacement and crack-resistance crepancy between mode I and mode II fracture toughness is attrib-
curves showed a good agreement with the experimental counter- uted to the fracture mode-induced toughening in soft adhesive-
parts. The average of the numerically extracted mode I fracture hard substrate systems, which also is common for a wide range
parameters and the standard deviation are tabulated in Table 2. of adhesives [51,52].
Fig. 7(b) shows the experimentally measured force–displace-
ment curves obtained from two ENF samples along with the
numerically derived counterparts. Initially, the force is linearly 3.4. Flexural behavior of Multi-material struts
increasing with the increasing center-line displacement. Upon
reaching the critical interfacial shear stress level at the PLA-TPU We examined the flexural behavior of the multi-material struts
interface, the interfacial crack starts propagating, which results in with varying core-to-strut thickness ratio c=t by 3-point bend tests.
a slight force drop. The whole crack growth process is completed Fig. 8(a) shows a representative set of the experimentally mea-
within this force plateau region. Then, the force begins increasing sured flexural stress–strain curves with varying c=t ratios. The
since the interfacial crack arrest at the interface underneath the ratios of 0 and 1 stand for the single-phase PLA and TPU samples,
loading pin due to a large crack-closure stress. Fig. 7(d) shows respectively. The flexural stress–strain curve when the ratio
the corresponding mode II fracture energy release rate curves c=t ¼ 0 shows a linear deformation with a modest nonlinearity

Fig. 8. (a) The experimentally measured flexural stress–strain curves for varying core-to-strut thickness ratios. The variations of (b) flexural modulus, (c) strength, and (d)
strain energy density as a function of the core-to-strut thickness ratio. The inset images in (a) illustrate the 3-point bend configuration and geometry of a multi-material strut.
The symbols in (b-d) show the experimental measurements (open and closed diamonds stand for the interfacial fracture and shell failure, respectively), while the dashed lines
in (b-d) show the numerical and analytical predictions.

8
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

up to a critical stress level and then a sudden force-drop because of PLA-TPU interfaces due to a better printing quality. The printing
semi-brittle fracture. On the other hand, the flexural stress–strain quality of the TPU and PLA interface was observed to be relatively
curve when the ratio c=t ¼ 1 shows a rising trend up to almost random due to the restricted capabilities of the utilized 3D printer.
0.15 strain without any trace of permanent deformation. The flex- In addition, no meaningful correlation was noticed between the
ural stress–strain curves of the multi-material struts reveal a linear printing quality of the PLA-TPU interface and the geometric param-
deformation within the initial portion of the loading. Upon exceed- eters such as c=t. This observation suggests that the flexural prop-
ing the critical stress level (i.e., either the critical interfacial erties of the multi-material struts can be significantly enhanced by
strength of the PLA-TPU interface or the yield strength of the PLA the precise control of the printing parameters.
shells), the flexural stress reduces to a lower stress plateau and To understand two distinct failure modes (i.e., shell cracking
maintains the stress level up to about 0.15 strain. It should be men- and interfacial cracking), Fig. 9(a) shows two flexural stress–strain
tioned that none of the multi-material struts revealed catastrophic curves obtained from 3-point bend tests of two different multi-
failure. The tail of the flexural stress–strain curves with a plateau material struts with c/t ¼ 1=2. These two samples showed either
stress contributed to the energy abortion capacity of the multi- shell or interfacial failure. The corresponding in-situ shear strain
material struts under bending deformation. The flexural tests of distributions captured at three different stages of the bending
the multi-material struts were repeated on 21 samples with 7 dis- deformation are shown in Fig. 9(b) and 9(c) for the shell cracking
tinct c=t ratios including c=t ¼ 0; 1=6; 2=6; 3=6; 4=6; 3=4 and 1. and interfacial cracking failure, respectively. The stress–strain
The variations of the experimentally measured flexural modu- curve of the multi-material strut failed due to shell cracking shows
lus, strength, and strain energy density as a function of c=t were a linear increase until the yield strength of the PLA shell is attained.
summarized in Fig. 8(b), 8(c), and 8(d), respectively. As anticipated, Then, the lower PLA shell undergoes yielding due to the tensile
the flexural modulus shows a gradual decrease as c=t increases. stress, which results in a plateau in the stress–strain curve. Upon
However, the measured decreasing trend was found to be much the tensile stress reaching a critical level, shell cracking occurs
more severe than the predicted variation by the standard beam and results in a large stress drop in the stress–strain curve. Despite
theory (Eq (5)). In other words, the analytical estimation overesti- the unstable failure, the strut remains its load carrying capacity
mated the flexural modulus of the multi-material struts. There are with a steady stress at about 8 MPa.
two plausible explanations for the observed discrepancy. First, the Inspecting the corresponding shear strain maps by DIC shown
standard beam theory is valid for perfectly elastic core and shell in Fig. 9(b), one can note that the TPU core reveals a homogenous
materials, which is not completely accurate in this case since shear deformation until the yielding point (point B), and the shell
TPU is highly hyper-elastic. In addition, the standard beam theory cracking can be easily seen at the bottom shell at the centerline
does not include the contribution of shear deformation on the of the sample (point C). The stress–strain curve of the multi-
measured stiffness, which can cause the observed overestimation material strut that failed due to interfacial cracking shows a linear
considering the short support span length in 3-point bend tests increase until the critical interfacial stress level at point E, at which
[53]. Second, such a discrepancy can be associated with the large interfacial crack nucleates in the upper PLA-TPU interface as shown
mismatch between the core and shell moduli (EPLA =ETPU  50). in Fig. 9(c). Then, the flexural stress gradually decreases as the
On the other hand, the FE simulations of the 3-point bend tests interfacial crack propagates along the upper PLA-TPU interface
of the struts provide a remarkably decent match with the experi- and reaches a steady stress level at about 13 MPa. The shear strain
mental variation of the flexural modulus. FE simulations employ- map captured at point F as shown in Fig. 9(c) clearly displays the
ing the CZM were performed using the upper and lower bounds propagation of the interfacial crack toward the centerline of the
of the cohesive fracture parameters. In addition, the FE simulations sample. The pre-mature failure of the multi-material struts due
with perfectly tied PLA-TPU interfaces were conducted to simulate to the weak interface limits the flexural strength. Such a weak
the perfect interfacial bonding and explore the limits of the multi- PLA-TPU interface can also be linked to high interfacial void con-
material struts. All the FE predictions shown in Fig. 8(b) match centration. In addition, the soft core does not undergo large defor-
each other, as the cohesive interface surface does not affect the ini- mation due to the limited strength, which results in an insufficient
tial bending stiffness of the samples [45]. Similar to the flexural energy absorption capacity. Therefore, the combined experimental
modulus, the flexural strength exhibited a gradual reduction with and numerical results suggest that it is important to tailor the
the increasing c=t as shown in Fig. 8(c). interfaces between the phases in the design and manufacturing
Lastly, the energy absorption capacity of the multi-material of multi-material struts to achieve their full potential.
struts was evaluated through the flexural strain energy density,
which is simply the area under the flexural stress–strain curve. 3.5. In-plane compression behavior of Multi-material lattices
Unlike the flexural modulus and strength, the strain energy density
exhibited a stimulating correlation with the c=t ratio. For c=t < 0:2, The in-plane compression behavior of the 3  3 honeycomb lat-
the multi-material struts provided equal or larger energy absorp- tices made of multi-material struts with varying c=t ratios was
tion than those of the constituent materials. For c=t > 0:2, the characterized by compression tests. Fig. 10(a) shows a set of nom-
strain energy density gradually decreases to the lower limit evalu- inal stress–strain curves for c=t ¼ 0; 1=6; 2=6; 4=6; and 1. The fully
ated for the bare TPU sample. Most of the experimental measure- PLA honeycomb lattice (c=t ¼ 0) exhibits a linear deformation up
ments fall into the region between the upper and lower limits to a critical stress level of about 9 MPa. Then a series of cracks
evaluated by the FE simulations employing CZM. It is noteworthy nucleate perpendicular to the struts near the nodes. The shear
that the flexural strength and strain energy density of two multi- strain map in Fig. 10(b) shows the strain concentrations around
material struts with c=t ¼ 3=6 and 3=4 (represented by closed dia- the cracks, which are marked by the arrows. Then, these cracks
monds in Fig. 8(c) and 8(d)) were measured to be very close to the unstably propagate and result in catastrophic failure, which is
strength limit estimated by the FE simulations employing observed as a large stress drop in the stress–strain curve. Following
perfectly-bonded PLA-TPU interfaces. These two samples revealed the failure, the fully PLA honeycomb lattice completely loses its
a failure due to shell cracking instead of interfacial failure of the loading carrying capacity. On the other hand, the nominal stress–
PLA-TPU interface. The larger flexural strength and strain energy strain curve of the fully TPU lattice (c=t ¼ 1) shows a rising trend
density measured on these two samples were attributed to the up to almost 0.35 strain and reaches a plateau level at about
observed failure transition from interfacial fracture to shell crack- 0.6 MPa. The fully TPU lattice recovered to its original shape upon
ing. This is a result of improved interfacial fracture properties of the removal of loading.
9
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

Fig. 9. (a) The experimentally measured flexural stress–strain curves of two different multi-material struts with c=t ¼ 1=2 showing shell and interfacial failure. The
corresponding shear strain distributions measured by DIC at A, B, C and D, E, F stages during the deformation of samples that exhibited (b) shell cracking and (c) interfacial
failure, respectively.

Fig. 10. (a) A representative set of nominal stress–strain curves obtained from in-plane compression tests of the multi-material honeycomb lattices with different c=t ratios.
The vertical strain eyy and shear strain exy maps measured by DIC captured at the peak stresses at points P and A for (b) c=t ¼ 0 and (c) c=t ¼ 1=6, respectively. (d) The optical
images showing the progressive damage of the multi-material honeycomb lattice with c=t ¼ 1=6 captured at points A, B, C, D, and E as marked in (a).

The stress–strain curves of the multi-material honeycomb lat- lattice, the multi-material lattices exhibit a progressive failure with
tices show a linear deformation up to a critical stress level. Then, multiple peak stresses. The existence of the soft core prevents the
with failure of the PLA-TPU interfaces, the compressive stress lattice material from a catastrophic failure and maintains the over-
reduces to a lower bound of about 0.6 MPa. Fig. 10(c) shows the all load carrying capacity of the lattice. Accordingly, the observed
vertical and shear strain maps measured by DIC at the first peak progressive failure of the multi-material lattices leads to a substan-
stress (point A in Fig. 10(a)) for the multi-material honeycomb lat- tial increase in energy absorption compared to the fully PLA lattice,
tice with c=t ¼ 1=6 . Based on the vertical strain field, one can find even though the strength is compromised by the multi-material
that the horizontal struts denoted by the arrows reveal a high pos- lattice design. In addition, each of the stress peaks in the stress–
itive vertical strain, which is the evidence of the interfacial failure strain curves of the multi-material lattices is associated with the
at these PLA-TPU interfaces. This suggests that the compressive collapse of a row of unit cells. Fig. 10(d) depicts the optical images
strength of the multi-material lattice correlates with the interfacial of the progressive damage for the multi-material lattice with
strength of the PLA-TPU interface. However, unlike the fully PLA c=t ¼ 1=6 captured at points A, B, C, D, and E as marked in Fig. 10

10
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

Fig. 11. The variation of in-plane compression (a) modulus, (b) strength, and (c) strain energy density as a function of the core-to-strut thickness ratio, c=t, (error bars denote
the standard deviation) (d) An Ashby plot depicting specific energy vs. density for the comparison of the energy absorption capacity of the proposed multi-material lattices in
the current study with single-phase polymeric [54–56] and metallic [57,58] lattices with various cell geometries and architectures from the relevant literature.

(a). During each collapse, the multi-material lattices revealed 45° multiple force fluctuations, which in turn leads to the substantial
fracture bands, which results in a slightly unsymmetric collapse energy absorption. The specific energy absorption capacity of the
during the densification. The formation of shear bands can be proposed multi-material lattices can be further improved by the
related to the multi-material strut architecture with the brittle fabrication of stronger PLA-TPU interfaces. Fig. 11(a-c) also suggest
PLA shells as well as bending-dominated deformation of honey- that compression modulus, compression strength, and energy
comb lattices [29,31]. However, the fully TPU lattice showed a absorption capacity of the multi-material lattices can be tuned
symmetric collapse due to the ductile nature of the material. It by varying the core-to-strut thickness ratio.
should also be noted that the multi-material lattice experiences Fig. 11(d) shows a typical Ashby graph of specific energy
densification upon exceeding about 60% strain, which leads to a absorption vs. density to compare energy absorption capacity of
rise in the nominal stress. the proposed multi-material lattices in the current study with
The in-plane compression tests of the multi-material lattices single-phase polymeric [54–56] and metallic [57,58] lattices with
were repeated on 15 samples with 5 distinct c=t ratios including various cell geometries and architectures from literature. The
c=t ¼ 0; 1=6; 2=6; 4=6; 1. The variations of the experimentally mea- single-phase polymeric lattices characteristically achieve specific
sured in-plane compression modulus (MPa), strength (MPa), and energy absorption of 0.05–0.5 MJ/m3 within the density range of
specific energy absorption (MJ/m3) with the varying c=t are shown 5–400 kg/m3. Such a wide variability is associated with the cell
in Fig. 11 (a-c). One can observe a drastic change in both in-plane geometries. Whereas, within the range of density, the single-
compression modulus and strength as the c=t ratio increases from phase stainless steel (SS) lattices maintain an order of magnitude
0 to 1=6. This suggests that even a thin layer of soft-core inclusion greater energy absorption. Noteworthy is that the proposed
in the lattice walls causes more than 2-fold reduction in both in- multi-material design raises the energy abortion capacity of the
plane compression modulus and strength. For c=t > 1=6, the in- polymeric lattices to the level of that of the metallic lattices with
plane compression modulus and strength exhibit a gradual standard cell geometries. However, it should be mentioned that
decrease with the increasing c=t. It should be noted that the mea- newly emerging metallic lattices with novel cell geometries such
sured in-plane modulus and strength are still limited since all the as a gyroid-type triply periodic minimal surfaces can greatly out-
lattice samples exhibit interfacial failure due to improper interfa- perform the regular metallic lattices and maintain energy absorp-
cial adhesion between soft and hard phases. Therefore, based on tion capacity of 50–60 MJ/m3 [31]. With the further improvement
the insight gained from the flexural test of the multi-material of the soft/hard-phase interfaces, the proposed multi-material lat-
struts, the in-plane compression modulus and strength of the tices can possibly make the polymeric lattice materials more
multi-material lattice can be further improved by stronger PLA- attractive for the practical applications.
TPU interfaces.
On the other hand, the multi-material lattice exhibits much
greater strain energy density than those of the individual con- 4. Conclusions
stituent materials for all the examined c=t ratios. The substantially
increased specific energy absorption capacity arises from the In summary, architected multi-material lattices with tunable
observed progressive failure of the multi-material lattices. As mechanical properties were designed and fabricated via FFF-
shown in the force–displacement curves in Fig. 10(a), the catas- based 3D printing. First, the microstructural characteristics of the
trophic failure of the hard lattice limits the energy absorption, PLA-TPU interfaces were studied via optical microscopy. Then,
while the multi-material lattice reveals a progressive damage with interfacial mode I and mode II fracture strength and toughness of
11
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

the PLA-TPU interface were quantitatively characterized by a com- Data Availability


bined experimental and numerical study. The flexural behavior of
the multi-material struts with varying core-to-strut thickness c=t The raw/processed data required to reproduce these findings
ratios was explored by 3-point bend tests. Lastly, in-plane com- cannot be shared at this time as the data also forms part of an
pression tests were performed to quantitatively determine the ongoing study.
in-plane compression modulus, strength, and energy absorption
capacity of the 3D printed architected multi-material lattices with
varying c=t ratios. The following conclusions were drawn from this Declaration of Competing Interest
study:
The authors declare that they have no known competing finan-
1. A typical 3D printed PLA-TPU interface exhibited distinct cial interests or personal relationships that could have appeared
microstructural features depending on the build orientation. to influence the work reported in this paper.
The PLA-TPU interface on a plane parallel to the build direction
revealed a saw-tooth shaped waviness with presence of random
interfacial voids. Such a rough interfacial morphology can play a References
significant role in improving interfacial adhesion by means of
[1] F. Barthelat, Architectured materials in engineering and biology: fabrication,
mechanical interlocking. However, the PLA-TPU interface on a structure, mechanics and performance, Int. Mater. Rev. 60 (8) (2015) 413–430.
plane orthogonal to the build direction showed a flat morphol- [2] L.J. Gibson, M.F. Ashby, Cellular Solids: Structure and Properties, Cambridge
ogy with a series of voids near the interface. University Press, Cambridge UK, 1999.
[3] M.F. Ashby, Materials Selection in Mechanical Design, Elsevier Science,
2. The experimentally measured and numerically calibrated mode Amsterdam, Netherlands, 2010.
I and mode II fracture strength and toughness of the PLA-TPU [4] N.A. Fleck, V.S. Deshpande, M.F. Ashby, Micro-architectured materials: past,
interface were r b ¼ 1.0 0:2MPa , b s ¼ 2:7  0:5MPa and present and future, Proceedings of the Royal Society A: Mathematical, Physical
and Engineering Sciences 466 (2121) (2010) 2495–2516.
CI ¼ 48  10J=m2 , CII ¼ 220  70J=m2 , respectively. The large [5] E.B. Joyee, A. Szmelter, D. Eddington, Y. Pan, Magnetic field-assisted
discrepancy between the mode I and mode II fracture parame- stereolithography for productions of multimaterial hierarchical surface
structures, ACS Appl. Mater. Interfaces 12 (37) (2020) 42357–42368.
ters is attributed to the fracture mode-induced toughening.
[6] M. Lei, W. Hong, Z. Zhao, C. Hamel, M. Chen, H. Lu, H.J. Qi, 3D printing of auxetic
3. The flexural modulus and strength of the multi-material struts metamaterials with digitally reprogrammable shape, ACS Appl. Mater.
of TPU core and PLA shell gradually decrease with the increas- Interfaces 11 (25) (2019) 22768–22776.
ing core-to-strut thickness ratio. The analytical estimation [7] K.A. Hing, Bone repair in the twenty–first century: biology, chemistry or
engineering?, Philosophical Transactions of the Royal Society of London, Series
based on the beam bending theory overestimated the flexural A: Mathematical, Physical and Engineering Sciences 362 (1825) (2004) 2821–
modulus with the varying core-to-strut thickness ratios. 2850.
4. The energy absorption capacity of the multi-material struts [8] M.A. Meyers, P.-Y. Chen, A.-Y.-M. Lin, Y. Seki, Biological materials: structure
and mechanical properties, Prog. Mater Sci. 53 (1) (2008) 1–206.
under bending was found to be greater than that of the individ- [9] A. Jackson, J.F. Vincent, R. Turner, The mechanical design of nacre, Proceedings
ual constituents up to a threshold c=t ratio of 0.2. The numerical of the Royal society of London. Series B. Biological sciences 234(1277) (1988)
results suggested that both the flexural strength and energy 415-440.
[10] L.R. Meza, S. Das, J.R. Greer, Strong, lightweight, and recoverable three-
absorption capacity of the multi-material struts can be further dimensional ceramic nanolattices, Science 345 (6202) (2014) 1322–1326.
improved by enhancing the adhesion at the PLA-TPU interfaces. [11] T.A. Schaedler, A.J. Jacobsen, A. Torrents, A.E. Sorensen, J. Lian, J.R. Greer, L.
5. The flexural strength and energy absorption capacity of the Valdevit, W.B. Carter, Ultralight metallic microlattices, Science 334 (6058)
(2011) 962–965.
multi-material struts under bending are significantly depen-
[12] M.F. Ashby, T. Evans, N.A. Fleck, J. Hutchinson, H. Wadley, L. Gibson, Metal
dent on the interfacial adhesion strength of the PLA-TPU inter- foams: a design guide, Elsevier Science, Amsterdam, Netherlands, 2000.
faces. The multi-material strut with strong PLA-TPU interfaces [13] L. Montemayor, J. Greer, Mechanical response of hollow metallic nanolattices:
combining structural and material size effects, J. Appl. Mech. 82 (7) (2015).
resulted in shell cracking failure, which leads to a notably
[14] X. Zheng, W. Smith, J. Jackson, B. Moran, H. Cui, D.a. Chen, J. Ye, N. Fang, N.
higher flexural strength and strain energy density. On the other Rodriguez, T. Weisgraber, C.M. Spadaccini, Multiscale metallic metamaterials,
hand, the multi-material strut with weak PLA-TPU interfaces Nat. Mater. 15 (10) (2016) 1100–1106.
revealed relatively less flexural strength and strain energy den- [15] X. Zheng, H. Lee, T.H. Weisgraber, M. Shusteff, J. DeOtte, E.B. Duoss, J.D. Kuntz,
M.M. Biener, Q.i. Ge, J.A. Jackson, S.O. Kucheyev, N.X. Fang, C.M. Spadaccini,
sity due to pre-mature interfacial failure. Ultralight, ultrastiff mechanical metamaterials, Science 344 (6190) (2014)
6. Tunable in-plane compression modulus and strength of the 1373–1377.
multi-material lattices were achieved by varying the core-to- [16] M. Jamshidian, N. Boddeti, D. Rosen, O. Weeger, Multiscale modelling of soft
lattice metamaterials: Micromechanical nonlinear buckling analysis,
strut thickness ratio. The compression modulus and strength experimental verification, and macroscale constitutive behaviour, Int. J.
of the multi-material lattices reveal a similar trend exhibited Mech. Sci. 188 (2020) 105956.
in the flexural modulus and strength of the multi-material [17] D. Wang, Y.i. Xiong, B. Zhang, Y.-F. Zhang, D. Rosen, Q.i. Ge, Design framework
for mechanically tunable soft biomaterial composites enhanced by modified
struts. The failure strength of the multi-material lattice materi- horseshoe lattice structures, Soft Matter 16 (6) (2020) 1473–1484.
als is decided by the interfacial fracture strength of the PLA-TPU [18] K. Yu, H. Du, A. Xin, K.H. Lee, Z. Feng, S.F. Masri, Y. Chen, G. Huang, Q. Wang,
interfaces. Therefore, by improving adhesion between the con- Healable, memorizable, and transformable lattice structures made of stiff
polymers, NPG Asia Mater. 12 (1) (2020) 1–16.
stituent phases, the in-plane compression modulus and
[19] Y. Jiang, Q. Wang, Highly-stretchable 3D-architected mechanical
strength of the multi-material lattices can be significantly metamaterials, Sci. Rep. 6 (1) (2016) 1–11.
enhanced. [20] T.A. Schaedler, W.B. Carter, Architected cellular materials, Annu. Rev. Mater.
Res. 46 (1) (2016) 187–210.
7. Tunable energy absorption capacity of the multi-material lat-
[21] T.S. Lumpe, J. Mueller, K. Shea, Tensile properties of multi-material interfaces
tices was achieved by varying the core-to-strut thickness ratio. in 3D printed parts, Mater. Des. 162 (2019) 1–9.
The multi-material lattices showed significantly greater (about [22] J. Song, L. Gao, K.e. Cao, H. Zhang, S. Xu, C. Jiang, J.U. Surjadi, Y. Xu, Y. Lu, Metal-
2–3 folds) energy absorption capacity compared to those of the coated hybrid meso-lattice composites and their mechanical
characterizations, Compos. Struct. 203 (2018) 750–763.
individual constituents for all the examined core-to-strut thick- [23] T. Juarez, A. Schroer, R. Schwaiger, A.M. Hodge, Evaluating sputter deposited
ness ratios. The fully PLA lattice exhibited semi-brittle failure, metal coatings on 3D printed polymer micro-truss structures, Mater. Des. 140
while all the multi-material lattices showed a progressive fail- (2018) 442–450.
[24] X. Zhang, J. Yao, B. Liu, J. Yan, L. Lu, Y.i. Li, H. Gao, X. Li, Three-dimensional high-
ure with multiple stress peaks during the deformation, which entropy alloy–polymer composite nanolattices that overcome the strength–
increases the energy absorption capacity. recoverability trade-off, Nano Lett. 18 (7) (2018) 4247–4256.

12
D. Yavas, Q. Liu, Z. Zhang et al. Materials & Design 217 (2022) 110613

[25] J. Mueller, J.R. Raney, K. Shea, J.A. Lewis, Architected lattices with high stiffness [42] ASTM D7264 / D7264M-07, Standard Test Method for Flexural Properties of
and toughness via multicore–shell 3d printing, Adv. Mater. 30 (12) (2018) Polymer Matrix Composite Materials, ASTM International, West
1705001, https://doi.org/10.1002/adma.v30.1210.1002/adma.201705001. Conshohocken, PA, 2007.
[26] T. Stanković, J. Mueller, P. Egan, K. Shea, A generalized optimality criteria [43] Astm, C365 / C365M–16, Standard Test Method for Flatwise Compressive
method for optimization of additively manufactured multimaterial lattice Properties of Sandwich Cores, ASTM International (2016).
structures, J. Mech. Des. 137 (11) (2015). [44] M. Smith, ABAQUS/Standard User’s Manual, Version 6.13, Dassault Systèmes
[27] Y.-C. Chan, K. Shintani, W. Chen, Robust topology optimization of multi- Simulia Corp, Providence, RI, 2013.
material lattice structures under material and load uncertainties, Frontiers of [45] M. Mooney, A theory of large elastic deformation, J. Appl. Phys. 11 (9) (1940)
Mechanical Engineering 14 (2) (2019) 141–152. 582–592.
[28] J.W. Boley, W.M. van Rees, C. Lissandrello, M.N. Horenstein, R.L. Truby, A. [46] D. Yavas, Z. Zhang, Q. Liu, D. Wu, Fracture behavior of 3D printed carbon fiber-
Kotikian, J.A. Lewis, L. Mahadevan, Shape-shifting structured lattices via reinforced polymer composites, Compos. Sci. Technol. 208 (2021) 108741.
multimaterial 4D printing, Proc. Natl. Acad. Sci. 116 (42) (2019) 20856–20862. [47] Y. Song, Y. Li, W. Song, K. Yee, K.-Y. Lee, V.L. Tagarielli, Measurements of the
[29] L. Yang, C. Yan, W. Cao, Z. Liu, B.o. Song, S. Wen, C. Zhang, Y. Shi, S. Yang, mechanical response of unidirectional 3D-printed PLA, Mater. Des. 123 (2017)
Compression–compression fatigue behaviour of gyroid-type triply periodic 154–164.
minimal surface porous structures fabricated by selective laser melting, Acta [48] A. Hillerborg, M. Modéer, P.-E. Petersson, Analysis of crack formation and crack
Mater. 181 (2019) 49–66. growth in concrete by means of fracture mechanics and finite elements, Cem.
[30] C. Yan, L. Hao, A. Hussein, P. Young, Ti–6Al–4V triply periodic minimal surface Concr. Res. 6 (6) (1976) 773–781.
structures for bone implants fabricated via selective laser melting, J. Mech. [49] T.L. Anderson, Fracture mechanics: fundamentals and applications, CRC
Behav. Biomed. Mater. 51 (2015) 61–73. press2017.
[31] C. Zhang, H. Zheng, L. Yang, Y. Li, J. Jin, W. Cao, C. Yan, Y. Shi, Mechanical [50] P.H. Geubelle, J.S. Baylor, Impact-induced delamination of composites: a 2D
responses of sheet-based gyroid-type triply periodic minimal surface lattice simulation, Compos. B Eng. 29 (5) (1998) 589–602.
structures fabricated using selective laser melting, Mater. Des. 214 (2022) [51] K.M. Liechti, Y.S. Chai, Asymmetric Shielding in Interfacial Fracture Under In-
110407, https://doi.org/10.1016/j.matdes.2022.110407. Plane Shear, J. Appl. Mech. 59 (2) (1992) 295–304.
[32] ISO / ASTM52921-13, Standard Terminology for Additive Manufacturing— [52] V. Tvergaard, J.W. Hutchinson, The influence of plasticity on mixed mode
Coordinate Systems and Test Methodologies, ASTM International, West interface toughness, J. Mech. Phys. Solids 41 (6) (1993) 1119–1135.
Conshohocken, PA, 2019. [53] J. He, M.Y. Chiang, D.L. Hunston, Assessment of sandwich beam in three-point
[33] ASTM D638-14, Standard Test Method for Tensile Properties of Plastics, ASTM bending for measuring adhesive shear modulus, J. Eng. Mater. Technol. 123 (3)
International, West Conshohocken, PA, 2014. (2001) 322–328.
[34] ASTM D7905 / D7905M-19e1, Standard Test Method for Determination of the [54] F.N. Habib, P. Iovenitti, S.H. Masood, M. Nikzad, Fabrication of polymeric lattice
Mode II Interlaminar Fracture Toughness of Unidirectional Fiber-Reinforced structures for optimum energy absorption using Multi Jet Fusion technology,
Polymer Matrix Composites, ASTM International, West Conshohocken, PA, Mater. Des. 155 (2018) 86–98.
2019. [55] M. Mohsenizadeh, F. Gasbarri, M. Munther, A. Beheshti, K. Davami, Additively-
[35] ASTM D5528-13, Standard Test Method for Mode I Interlaminar Fracture manufactured lightweight Metamaterials for energy absorption, Mater. Des.
Toughness of Unidirectional Fiber-Reinforced Polymer Matrix Composites, 139 (2018) 521–530.
ASTM International, West Conshohocken, PA, 2013. [56] X. Chen, Q. Ji, J. Wei, H. Tan, J. Yu, P. Zhang, V. Laude, M. Kadic, Light-weight
[36] ASTM D790-17, Standard Test Methods for Flexural Properties of Unreinforced shell-lattice metamaterials for mechanical shock absorption, Int. J. Mech. Sci.
and Reinforced Plastics and Electrical Insulating Materials, ASTM International, 169 (2020) 105288.
West Conshohocken, PA, 2017. [57] Y. Wang, X. Ren, Z. Chen, Y. Jiang, X. Cao, S. Fang, T. Zhao, Y. Li, D. Fang,
[37] ASTM D1621-16, Standard Test Method for Compressive Properties of Rigid Numerical and experimental studies on compressive behavior of Gyroid lattice
Cellular Plastics, ASTM International, West Conshohocken, PA, 2016. cylindrical shells, Mater. Des. 186 (2020) 108340.
[38] J. Blaber, B. Adair, A. Antoniou, Ncorr: open-source 2D digital image correlation [58] X. Cao, S. Duan, J. Liang, W. Wen, D. Fang, Mechanical properties of an
matlab software, Exp. Mech. 55 (6) (2015) 1105–1122. improved 3D-printed rhombic dodecahedron stainless steel lattice structure of
[39] Y. Wang, J. Williams, Corrections for mode II fracture toughness specimens of variable cross section, Int. J. Mech. Sci. 145 (2018) 53–63.
composites materials, Compos. Sci. Technol. 43 (3) (1992) 251–256. [59] Denizhan Yavas et al., Mode-I fracture toughness and surface morphology
[40] M.F.S.F. de Moura, M.A.L. Silva, A.B. de Morais, J.J.L. Morais, Equivalent crack evolution for contaminated adhesively bonded composite structures.,
based mode II fracture characterization of wood, Eng. Fract. Mech. 73 (8) Composite Structures (2018), https://doi.org/10.1016/
(2006) 978–993. j.compstruct.2018.07.014.
[41] G. Marckmann, E. Verron, Comparison of hyperelastic models for rubber-like [60] Denizhan Yavas et al., Prediction of interfacial surface energy and effective
materials, Rubber Chem. Technol. 79 (5) (2006) 835–858. fracture energy from contaminant concentration in polymer-based interfaces,
Journal of Applied Mechanics (2017), https://doi.org/10.1115/1.4035931.

13

You might also like