Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Materials and Design 169 (2019) 107655

Contents lists available at ScienceDirect

Materials and Design

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

Failure and energy absorption characteristics of four lattice structures


under dynamic loading
Nan Jin a, Fuchi Wang a,b, Yangwei Wang a,b,⁎, Bowen Zhang a, Huanwu Cheng a,b, Hongmei Zhang a,b
a
School of Materials Science and Engineering, Beijing Institute of Technology, Beijing 100081, China
b
National Key Laboratory of Science and Technology on Materials Under Shock and Impact, Beijing 100081, China

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

• Two novel diamond lattice structures


(Dfcc and Dhex) and the typical FCC
and BCC lattice structures were de-
signed.
• A numerical simulation model has been
developed, and was consistent well
with the experimental results.
• Using a power law function to relate dif-
ferent mechanical properties to rod di-
ameter or relative density for each
lattice type.
• Stretching-dominated structures often
exhibit better dynamic mechanical
properties than bending-dominated
structures.

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

Article history: Two novel diamond lattice structures (Dfcc and Dhex) and the typical face-centered cubic (FCC) and body-
Received 13 December 2018 centered cubic (BCC) lattice structures are manufactured by selective laser melting (SLM), and the effects of
Received in revised form 9 February 2019 cell topology and relative density on the dynamic behavior of these structures are studied through a combination
Accepted 10 February 2019
of experimental tests and numerical simulation. It is observed that in Dfcc and Dhex, failure initiates at the con-
Available online 14 February 2019
nection points between rods in the middle of the structure, causing a sudden drop of the measured stress value.
Keywords:
However in FCC and BCC, failure initiates in the face-truss junction. Generally, the FCC and BCC are dominated by
Lattice structures stretching and bending of the rods respectively, whereas the Dfcc and Dhex are a mixture of the two deformation
Mechanical properties modes. The results show that the mechanical properties of the lattice structures with different relative density
Failure analysis can be described by a power law function. Moreover, for the lattice structures with the same rod diameter of
Energy absorption 0.8 mm, FCC out-performs other structures in terms of specific strength, specific modulus and energy absorption.
This gives evidence that lattice structures with the stretching-dominated deformation mode are more likely to
exhibit better mechanical properties under dynamic loading.
© 2019 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

Lightweight lattice material with high strength, high modulus and


⁎ Corresponding author at: School of Materials Science and Engineering, Beijing
good energy absorption capacity has a great potential of application in
Institute of Technology, Beijing 100081, China. many fields such as automobile, aerospace and military industry [1,2].
E-mail address: wangyangwei@bit.edu.cn (Y. Wang). Recently, additive manufacturing (AM) such as electron beam melting

https://doi.org/10.1016/j.matdes.2019.107655
0264-1275/© 2019 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/).
2 N. Jin et al. / Materials and Design 169 (2019) 107655

Fig. 1. Configurations of (a) Dfcc, (b) Dhex, (c) BCC, (d) FCC lattice structures.

(EBM) [3–5] and selective laser melting (SLM) [6–8] are increasingly bearing capacity, while the tetrahedron lattice material performed the
used to fabricate lattice structures. Compared with traditional methods, worst. Harris et al. [10] optimized the honeycomb structure by replacing
it has the advantages of high precision and the ability to fabricate com- the wall with lattice truss, which improved its mechanical properties.
plex structures. Ullah et al. [11] studied the failure and energy absorption characteristics
Previous studies have showed that the mechanical properties of lat- of Kagome, bcc, f2cc and f2bcc lattice structures with one unit cell, and
tice material can be improved by optimize the core structure. Guo et al. found that the Kagome structure performed best in terms of specific
[9] studied the static mechanical properties of square, tetrahedron, strength and energy absorption capacity. Hasan et al. [12] found that
hexagon and quadrate lattice material with the same cross-section the drop weight impact performance of the sandwich panels with tita-
size, and found that the hexagon material showed the optimal overall nium Kagome lattice structure was better than the aluminum

Fig. 2. Structural unit of (a) Dfcc and Dhex, (b) BCC, (c) FCC.

Fig. 3. Lattice specimens studied in the experiment: (a) Dfcc-0.8, (b) Dhex-0.8, (c) BCC-0.8 (d) FCC-0.8.
N. Jin et al. / Materials and Design 169 (2019) 107655 3

Table 1 of these structures under dynamic loading were studied through a com-
Details of the lattice structures studied in the experiment. bination of experimental tests and numerical simulation.
Structure Rod diameter Rod diameter Apparent density Relative
d [mm] (design) d [mm] (actual) ρ [kg/m3] density ρ
2. Experimental study
Dfcc-0.8 0.8 0.8 211.7 0.049
Dhex-0.8 0.8 0.8 115.6 0.027 2.1. Design and manufacture of four lattice structures
FCC-0.8 0.8 0.8 1129.9 0.262
BCC-0.8 0.8 0.8 513.2 0.119
The design of the four lattice structures is based on the crystal struc-
ture [23]. We try to verify whether lattice materials with the same crys-
tal structure of hard materials have better mechanical properties. The
honeycomb structure. Meza et al. [13] created a hollow-tube nanolattice four lattice structures investigated in this work are shown in Fig. 1.
with octet-truss geometry, and reported that optimizing the wall Dfcc has the same structure as diamond, which is one of the hardest ma-
thickness-to-radius of the tubes can result in ductile-like deformation terials in nature [24]. Dhex has the same structure as lonsdaleite, also
and recoverability. Ahmadi et al. [14] found that the compressive prop- called hexagonal diamond in reference to the crystal structures.
erties of lattice material increased with the increase of strut size and the Lonsdaleite is simulated to be 58% harder than diamond on the 〈100〉
relationship was a power law type. Wang et al. [15] and Thomas et al. face and to resist indentation pressures of 152 GPa, whereas diamond
[16] reported the similar results in their study. Zargarian et al. [5] stud- would break at 97 GPa [25]. Therefore, we chose the Dfcc and Dhex lat-
ied the fatigue behavior of truncated cuboctahedron, rhombic dodeca- tice structures, and compare their mechanical properties with the body-
hedron and diamond lattice structures and found that their fatigue centered cubic (BCC) and face-centered cubic (FCC) lattice structures.
behavior could be described by a power law function. The coefficient The four lattice structures with limited number of cells were adopted
of the power law function was depend on cell geometry, relative density for both considering reducing the boundary effect and computing time
and the fatigue behavior of the bulk material, while the exponent was in numerical simulation. The diamond lattice structure (Dfcc) and the
only depend on the fatigue properties of the bulk material. hexagonal diamond lattice structure (Dhex) have the same structural
The deformable core plays such an important role in the load bearing unit as depicted in Fig. 2(a). However in Dfcc, all the rods are in the stag-
process, thus superior structure is significant in various applications. gered conformation, thus causing all four cubic-diagonal directions to
Load bearing strength and modulus in quasi-static compression have be equivalent; while in Dhex, the rods between layers are in the eclipsed
been discussed by many researchers [9–11,14–17], while fewer re- conformation, which defines the axis of hexagonal symmetry. The third
search studies include dynamic mechanical properties [18–22]. In this and fourth geometry depict the body-centered cubic (BCC) and face-
study, two novel diamond lattice structures (Dfcc and Dhex) and the centered cubic (FCC) lattice structures and their structural units are
typical face-centered cubic (FCC) and body-centered cubic (BCC) lattice shown in Fig. 2(b) and (c) respectively. The size of the structural units
structures were manufactured by selective laser melting (SLM), and the is controlled by the length of the cube they occupied and the length is
effect of cell topology and relative density on the mechanical properties kept as a constant value of 5 mm.

Fig. 4. SEM observations of the lattice specimens: (a) an individual broken rod, (b) base material.

Fig. 5. Schematic of the SHPB apparatus.


4 N. Jin et al. / Materials and Design 169 (2019) 107655

Fig. 6. FE mesh of lattice structures: (a) Dfcc-0.8, (b) Dhex-0.8, (c) BCC-0.8, (d) FCC-0.8.

Table 2
Details of the lattice structures studied in the numerical simulation.

Structure-rod diameter d Relative Structure-rod diameter d Relative Structure-rod diameter d Relative Structure-rod diameter d Relative
[mm] density ρ [mm] density ρ [mm] density ρ [mm] density ρ

Dfcc-0.8 0.049 Dhex-0.8 0.027 FCC-0.3 0.044 BCC-0.5 0.049


Dfcc-1.0 0.076 Dhex-1.0 0.041 FCC-0.4 0.075 BCC-0.8 0.119
Dfcc-1.2 0.109 Dhex-1.2 0.061 FCC-0.5 0.118 BCC-1.0 0.178
Dfcc-1.4 0.147 Dhex-1.4 0.080 FCC-0.6 0.159 BCC-1.2 0.245
Dfcc-1.6 0.190 Dhex-1.6 0.103 FCC-0.7 0.208 BCC-1.4 0.319
Dfcc-1.8 0.235 Dhex-2.0 0.159 FCC-0.8 0.262 BCC-1.6 0.396
Dfcc-2.0 0.285 Dhex-2.2 0.185 FCC-0.9 0.319 BCC-1.8 0.476
Dfcc-2.2 0.339 FCC-1.0 0.379
FCC-1.1 0.439

The specimens were fabricated by SLM and the constituents of the mainly composed of continuous α phase and discretely distributed flat
base material were Ti-6Al-4V. The prepared specimens were kept at rod β phase (Fig. 4(b)).
800 °C in furnace for 2 h and then cooled down. Fig. 3 shows the speci-
mens which are composed of two face sheets with a thickness of 2.2. SHPB experimental apparatus
0.50 mm and the cores between them. The structural parameters of all
the lattice structures used in the experiment are listed in Table 1. Divid- The Split Hopkinson Pressure Bar (SHPB) is a typical experimental
ing the mass of the lattice structures by the volume they occupy gives apparatus for testing dynamic behavior of materials at strain rates rang-
their apparent density ρ, and the ratio of the apparent density ρ to the ing from 102 s−1 to 104 s−1, which has also been used for testing lattice
base material density ρs is the relative density ρ. In the experiment, materials [19,26,27]. The experimental tests were executed on the alu-
we compare the mechanical behavior of the four lattice structures minum bar with a diameter of 37 mm. The length of the bullet was
with the same rod diameter of 0.8 mm (Dfcc-0.8, Dhex-0.8, FCC-0.8 400 mm, while the incident and transmission bar were 2000 mm. The
and BCC-0.8). Fig. 4 shows the secondary electron image of the lattice metal strain gauge with sensitivity coefficient of 2.069 was used in the
specimens. The rod appears to deviate from the designed cylindrical incident bar to record the incident wave and reflected wave, and the
shape due to some non-molten residual powder particles on the rod semiconductor strain gauge with sensitivity coefficient of 110 was
surface (Fig. 4(a)). Moreover, the lattice material prepared by SLM is used in the transmission bar to record the relatively weak transmitted
wave signal. The Split Hopkinson bar system and the image system
were connected as shown in Fig. 5. All specimens were loaded with
strain rate control of about 1000 s−1.

3. Numerical simulation

3.1. Finite element model

The numerical simulation of the dynamic testing was composed of


solid structure modeling, meshing, pre-processing, solution and verifi-
cation of numerical results. The three dimensional models of the lattice
structures were developed in Solidworks software, and then trans-
formed into the ANSA software to generate the finite element model.
Fig. 7. The schematic of boundary and loading conditions applied to the FE model. The lattice materials were modeled with SOLID164 element sizes
N. Jin et al. / Materials and Design 169 (2019) 107655 5

Fig. 8. FEM model of the base material under dynamic compression test.

between 0.15 mm and 0.20 mm. As can be seen from Fig. 6, a uniform and experimental data are shown in Fig. 9. The parameters used in the
mesh of hexahedral elements of BCC and FCC structure is obtained. model are listed in Table 3.
However, there are inevitably a small number of tetrahedral elements
at the joints of the Dfcc and Dhex structure. For numerical analysis, 4. Results and discussions
the finite element models of the four lattice structures with different
relative density have been obtained by changing their rod diameter, as 4.1. Effect of cell topology on the dynamic behavior of the lattice structures
are shown in Table 2.
Fig. 10 shows the experimental compression behavior of the se-
lected lattice structures at a strain rate of about 1000 s−1 (EX) and the
3.2. SHPB simulation corresponding numerical predictions (SM). As can be seen, the numer-
ical predictions are matched closely by the experimental results with
The commercial software LS-DYNA was used to simulate the lattice less than 15% deviation. The differences are mainly caused by the
structures subjected to uniaxial dynamic compression. As is shown in manufacturing defects of the specimen [30–32] and the ideal boundary
Fig. 7, a moving rigid wall was defined on one side of the lattice struc- and loading conditions assumed in the numerical simulation. Moreover,
ture and a fixed rigid wall on the opposite one. The velocity of the mov- although the constitutive parameters of the applied material in the sim-
ing rigid wall was kept constant to ensure a constant strain rate during ulation are verified and fitted from the bulk material manufactured by
the loading process, and it can be determined by v ¼ ε_  h, in which ε_ is using the same SLM process, there may still be some differences with
the strain rate, h is the height of the lattice model. The stress and strain the lattice material used in the experiment because of the size effect
can be calculated by σ = F/A and ε ¼ ε_  t, in which F is the interaction and forming direction [33,34]. Nevertheless, from the failure character-
force between the fixed rigid wall and the lattice model, A is the original istics shown in the images and the shape of the stress-strain curves, it is
surface area of the lattice model and t is the loading time. Moreover, the feasible to assume that the simulation results are reliable and can be
contact between the rigid wall and the specimen is defined as used to make useful predictions. In general, the stress-strain curves
*AUTOMATIC_SURFACE_TO_SURFAC. show an initial linear response that is followed by the subsequent fluc-
The Johnson-Cook material model was employed to the lattice struc- tuations. The slope of the initial linear elastic region is defined as the
tures. This model has been widely used to describe the mechanical be- modulus E, the stress value at the highest peak is defined as the strength
havior of materials especially for metals subjected to dynamic loading σb, and the area under the stress-strain curve from the beginning to the
[28]. The stress-strain curves of the base material manufactured by strain of 0.2 is defined as the energy absorption per unit volume Ev. Di-
using the same SLM process under different strain rates were tested viding these mechanical properties by the apparent density ρ of the lat-
through SHPB apparatus. The compression direction of the specimen tice structures gives the specific modulus, specific strength and energy
was the printing direction, which was also consistent with the compres- absorption per unit mass. All the mechanical properties of the analyzed
sion direction of the lattice material. The initial guess for the material lattice structures obtained from numerical simulation are plotted in
constants of the Johnson-Cook model was obtained from [29]. Then Fig. 11.
the whole stress-strain curve for the base material under compression The results indicate the difference between the lattice structures
was simulated, and the FEM model is shown in Fig. 8. Set the bullet to with different types of cell topology. Not only do the mechanical prop-
impact the incident bar at different speeds to obtain the stress-strain erties of the four lattice structures differ prominently, but also the defor-
curve of the base material at different strain rates. Adjustments of the mation process and the failure mechanism change as well. These
material constants were made until the simulated result matched variations are reflected in the different shapes of stress-strain curves.
closely with the experimental data. Comparisons of the fitted results It can be observed that the Dfcc-0.8 and Dhex-0.8 that include vertical
rods exhibit a different failure mechanism, compared to the FCC-0.8
and BCC-0.8. In Dfcc-0.8 and Dhex-0.8, failure initiates in the connection
point between rods in the middle of the structure at point 1, shown in
Fig. 10(a) and (b), causing a sudden drop of the measured stress
value. Once that one of the connection points, which is the weakest
link of the structure, collapsed, the other parts of the structure take
over the load-bearing function and results in the stress increases.
When a new link breaks, the curve declines again and follows the
same pattern. However, in FCC-0.8, failure initiates in the face-truss
junction at point 1 shown in Fig. 10(d), and as the shape of the structure
is more complex and each of the rods provides a stronger support to
each other since there are more rods connected at the same joints, the
stress-strain curve is relatively smooth. After the failure point, the

Table 3
Material parameters used in the Johnson-Cook model [29].

A (MPa) B (MPa) C m n D1 D2 D3 D4 D5

913 250 0.032 1.1 0.2 0 0.23 0.48 0.04 3.9


Fig. 9. Comparison of the experimental data (EX) with simulation result (SM).
6 N. Jin et al. / Materials and Design 169 (2019) 107655

Fig. 10. The stress-strain curves of the lattice structures with the same rod diameter of 0.8 mm (Dfcc-0.8, Dhex-0.8, BCC-0.8, FCC-0.8): experimental (EX) and simulation (SM) results.

structure deforms progressively and it presents a clear rapture along 45° remains almost unchanged until point 2 when the junction in the mid-
angle. In BCC-0.8, the stress concentration points are the face-truss junc- dle layer breaks. From Fig. 11, it is evident that FCC-0.8 has the highest
tion and the junction of the two intermediate rods where the failure of specific strength, specific modulus and energy absorption, which is
the material starts. As can be seen from Fig. 10(c), the face-truss junc- mainly due to its complex structure and stretching-dominated defor-
tion of the structure breaks at point 1. After that, the rods in the middle mation mode. The stress distribution in the FCC-0.8 is more uniform
of the structure undergo large bending deformation and the stress value than Dfcc-0.8, Dhex-0.8 and BCC-0.8, in which their stress often concen-
trated in the joint points between rods in the middle layer. It should also
be noted that the specific strength, specific modulus, and energy ab-
sorption of Dhex-0.8 are higher than Dfcc-0.8 by 81.48%, 109.98%, and
84.38% respectively. This is explained by the fact that although the de-
formation process and failure mechanism of the two structures are sim-
ilar, the density of Dhex-0.8 is much lower than Dfcc-0.8.

4.2. Effect of relative density on the dynamic behavior of the lattice


structures

The simulated stress-strain curves of the four lattice structures with


different rod diameter under dynamic loading at a strain rate of
1000 s−1 are shown in Fig. 12. As can be seen, the failure mode of the
Dfcc and Dhex does not change as the rod diameter increases. In
Fig. 12(a) and (b), when the diameter increased to 2.2 mm, there is a
sudden stress release at the failure point, causing the sudden drop in
the stress-stain curves. However in BCC, the initial deformation zone
of the structure switches from the intermediate layer to the lower and
upper layer as the rod diameter is larger than 1 mm. Moreover, the
Fig. 11. Compression behavior of the four lattice structures with the same rod diameter of stress distribution of the lattice structure becomes gradually uniform.
0.8 mm (simulation results). In FCC, when the rod diameter is smaller than 0.6 mm, the initial
N. Jin et al. / Materials and Design 169 (2019) 107655 7

Fig. 12. Stress-strain curves of the lattice material with different rod diameter: (a) Dfcc, (b) Dhex, (c) BCC, (d) FCC.

deformation zone of the structure gradually transitions from the top in which the specimen are obtained by using a limit ring during the ex-
layer to the bottom layer. As the rod diameter increases, the structures perimental tests.
are gradually observed to have a clear rupture along 45°. Strength and modulus of different porous metal materials are plot-
As expected, each of the compressive properties of the lattice mate- ted in Fig. 17. When comparing the four lattice structures investigated
rial increases with the increase in relative density. In order to analyze in this work with the foam [37–39], honeycomb [40] and other lattice
the compressive properties of the lattice material more systematically, structures [9,10,19] from the literature, it is found that lattice structures
power laws, which relate structure mechanical properties to different have outstanding mechanical properties especially in the lower relative
relative density, are fitted to the numerical analysis data: density range. For compressive strength and modulus, Dhex out-
performs all commercially available foam and honeycomb materials,
X ¼ Kρm ð1Þ as well as the Dfcc, FCC, BCC, pyramid, cubic lattice structures. Moreover,
the Dfcc, FCC and BCC display a similar strength to the Ti honeycomb
and Ti foam, but have a higher modulus for the same relative density.
where X is the mechanical properties of the lattice structures and ρ is the The figure highlights the influence of structure on the mechanical prop-
relative density. The parameters K and m are depend on the types of cell erties of porous materials. It is shown that, as the relative density of the
topology. The exponent of the power law fitted to the numerical data material decreases, the strength and modulus of lattice materials de-
points (Figs. 13–15) varies between 1.32 (FCC) and 1.99 (BCC) for crease more slowly than other porous materials. Moreover, for the lat-
strength, between 1.31 (FCC) and 1.99 (BCC) for modulus, between tice structures with the same relative density, Dhex shows relatively
1.49 (Dhex) and 2.17 (BCC) for energy absorption. From the literature, good mechanical properties as its strength is comparative to FCC and
the strength and stiffness of an ideal stretching-dominated structure its modulus is higher than FCC. The strength of Dfcc is lower than FCC,
scale linearly with relative density as σ ∝ρ and E∝ρ. This contrast with whereas its modulus is similar to FCC. Although materials with diamond
bending-dominated structure with σ∝ρ2 and E∝ρ2 , or stochastic foams crystal structure often exhibits better mechanical properties than mate-
with σ ∝ρ3 and E∝ρ3 scaling [2,13,35,36]. So the FCC is dominated by rials with FCC and BCC crystal structures, it is not the case for lattice ma-
stretching of the rods, BCC is dominated by bending of the rods, whereas terials. Therefore, it is considered that there is no obvious
the Dfcc and Dhex are a mixture of the two deformation modes. This re- correspondence between the mechanical properties of lattice materials
sult is consistent with the deformation characteristics of the lattice and the mechanical properties of materials with the same crystal
structures observed during the loading process, as depicted in Fig. 16, structures.
8 N. Jin et al. / Materials and Design 169 (2019) 107655

(a) (b)

(c) (d)

Fig. 13. Relationship between strength and relative density for: (a) Dfcc, (b) Dhex, (c) BCC, (d) FCC.

(a) (b)

(c) (d)

Fig. 14. Relationship between modulus and relative density for: (a) Dfcc, (b) Dhex, (c) BCC, (d) FCC.
N. Jin et al. / Materials and Design 169 (2019) 107655 9

(a) (b)

(c) (d)

Fig. 15. Relationship between energy absorption per unit volume and relative density for: (a) Dfcc, (b) Dhex, (c) BCC, (d) FCC.

5. Conclusions in FCC and BCC, failure initiates in the face-truss junction. The results
show that the mechanical properties of lattice structures with different
The failure and energy absorption characteristics of four lattice relative density can be described by a power law function. The exponent
structures were studied by means of numerical simulation and the re- of the power law fitted to the numerical data points varies between 1.32
sults were consistent well with the experimental ones. It is observed (FCC) and 1.99 (BCC) for strength, between 1.31 (FCC) and 1.99 (BCC)
that the Dfcc and Dhex that include vertical rods exhibit a different fail- for modulus, between 1.49 (Dhex) and 2.17 (BCC) for energy absorp-
ure mechanism as compared to the FCC and BCC. In Dfcc and Dhex, fail- tion. Generally, the FCC and BCC are dominated by stretching and bend-
ure initiates in the connection points between rods in the middle of the ing of the rods respectively, whereas the Dfcc and Dhex are a mixture of
structure, causing a sudden drop of the measured stress value. However the two deformation modes. Moreover, lattice structures with the

Fig. 16. Deformation characteristics at the strain of 0.3: (a) BCC, (b) Dfcc, (c) FCC, (d) Dhex.
10 N. Jin et al. / Materials and Design 169 (2019) 107655

Fig. 17. Comparison of different porous metal materials: (a) strength, (b) modulus.

stretching-dominated deformation mode are more likely to exhibit bet- [14] S.M. Ahmadi, S.A. Yavari, R. Wauthle, et al., Additively manufactured open-cell po-
rous biomaterials made from six different space-filling unit cells: the mechanical
ter mechanical properties under dynamic loading. and morphological properties, Materials 8 (4) (2015) 1871–1896.
[15] Y. Wang, J.M. Chen, Y.P. Yuan, Influence of the unit cell geometrical parameter to the
mechanical properties of Ti6Al4V open-porous scaffolds manufactured by selective
CRediT authorship contribution statement laser melting, Appl. Mech. Mater. 851 (2016) 201–210.
[16] T.D. Thomas, A.B. Spierings, D. Mohr, Addictively-manufactured metallic micro-
lattice materials for high specific energy absorption under static and dynamic load-
Nan Jin: Methodology, Software, Formal analysis, Investigation, Writing - ing, Acta Mater. 116 (2016) 14–28.
original draft, Visualization. Fuchi Wang: Conceptualization, Data [17] S.T. Hong, J. Pan, T. Tyan, et al., Quasi-static crush behavior of aluminum honeycomb
specimens under compression dominant combined loads, Int. J. Plast. 22 (1) (2006)
curation, Supervision, Resources, Project administration, Funding acquisi-
73–109.
tion. Yangwei Wang: Conceptualization, Validation, Resources, Project [18] S. Mckown, Y. Shen, W.K. Brookes, et al., The quasi-static and blast loading response
administration, Data curation, Writing - review & editing, Supervision, of lattice structures, Int. J. Impact Eng. 35 (8) (2008) 795–810.
Funding acquisition. Bowen Zhang: Methodology, Resources, Validation. [19] J.G. Liu, S. Pattofatto, D.N. Fang, et al., Impact strength enhancement of aluminum
tetrahedral lattice truss core structures, Int. J. Impact Eng. 79 (2015) 3–13.
Huanwu Cheng: Validation, Writing - review & editing. Hongmei Zhang: [20] Z. Zheng, J. Yu, J. Li, Dynamic crushing of 2D cellular structures: a finite element
Validation, Writing - review & editing. study, Int. J. Impact Eng. 32 (1) (2005) 650–664.
[21] H. Zhao, G. Gary, Crushing behavior of aluminum honeycombs under impact load-
ing, Int. J. Impact Eng. 21 (10) (1998) 827–836.
[22] X. Tang, V. Prakash, J.J. Lewandowski, et al., Inertial stabilization of buckling at high
Acknowledgement rates of loading and low test temperatures: implications for dynamic crush resis-
tance of aluminum-alloy-based sandwich plates with lattice core, Acta Mater. 55
The authors acknowledge the support of the National Key Research (8) (2007) 2829–2840.
[23] M. Pham, C. Liu, L. Todd, et al., Damage-tolerant architected materials inspired by
and Development Program of China (Grant No. 2016YFB0700502).
crystal microstructure, Nature 565 (2019) 305–311.
[24] A. Banerjee, D. Bernoulli, H. Zhang, et al., Ultralarge elastic deformation of nanoscale
References diamond, Science 360 (2018) 300–302.
[25] Z.C. Pan, Sun Hong, Zhang Yi, et al., Harder than diamond: superior indentation
[1] H. Mozafari, S. Khatami, H. Molate, Out of plane crushing and local stiffness determi- strength of wurtzite BN and lonsdaleite, Phys. Rev. Lett. 102 (5) (2009), 055503. .
nation of proposed foam filled sandwich panel for Korean tilting train express- [26] L.J. Xiao, W.D. Song, C. Wang, et al., Mechanical properties of open-cell rhombic do-
numerical study, Mater. Des. 66 (2015) 400–411. decahedron titanium alloy lattice structure manufactured using electron beam
[2] T.A. Schaedler, A.J. Jacobsen, A. Torrents, et al., Ultralight metallic microlattices, Sci- melting under dynamic loading, Int. J. Impact Eng. 100 (2017) 75–89.
ence 334 (6058) (2011) 962–965. [27] J.L. Yu, J.R. Li, S.S. Hu, Strain-rate effect and micro-structural optimization of cellular
[3] S.L. Lu, M. Qian, H.P. Tang, et al., Massive transformation in Ti-6Al-4V additively metals, Mech. Mater. 38 (1) (2006) 160–170.
manufactured by selective electron beam melting, Acta Mater. 104 (2016) 303–311. [28] J. Sun, Y.B. Guo, Material flow stress and failure in multiscale machining titanium
[4] D. Cormier, O. Harrysson, H. West, Characterization of H13 steel produced via elec- alloy Ti-6Al-4V, Int. J. Adv. Manuf. Technol. 41 (7–8) (2009) 651–659.
tron beam melting, Rapid Prototyp. J. 10 (1) (2013) 35–41. [29] Tao Huang, Weiguo Wu, Xiaobin Li, et al., Numerical investigation on truncated
[5] A. Zargarian, M. Esfahanian, J. Kadkhodapour, et al., Numerical simulation of the fa- cylindroconical projectile penetrating thin target, Chin. J. Ship Res. 4 (2) (2009)
tigue behavior of additive manufactured titanium porous lattice structures, Mater. 48–52.
Sci. Eng. C 60 (2016) 339–347. [30] N. Meisel, C. Williams, A. Druschitz, Lightweight metal cellular structure via indirect
[6] G. Campoli, M.S. Borleffs, S.A. Yavari, et al., Mechanical properties of open-cell metal- 3D printing and casting, 24th International Solid Freeform Fabrication Symposium,
lic biomaterials manufactured using additive manufacturing, Mater. Des. 49 (2013) 2013.
957–965. [31] E. Hernandez-Nava, C.J. Smith, F. Derguti, et al., The effect of defects on the mechan-
[7] J. Yang, H. Yu, J. Yin, et al., Formation and control of martensite in Ti-6Al-4V alloy ical response of Ti-6Al-4V cubic lattice structures fabricated by electron beam melt-
produced by selective laser melting, Mater. Des. 108 (2016) 308–318. ing, Acta Mater. 108 (2016) 279–292.
[8] E. Chauvet, C. Tassin, J.J. Blandin, et al., Producing Ni-base superalloys single crystal [32] L. Liu, P. Kamm, F. Garcia-Moreno, et al., Elastic and failure response of imperfect
by selective electron beam melting, Scr. Mater. 152 (2018) 15–19. three-dimensional metallic lattices: the role of geometric defects induced by selec-
[9] R. Guo, R. Liu, W. Jiang, et al., Numerical analysis on static mechanical properties of tive laser melting, J. Mech. Phys. Solids 107 (2017) 160–184.
the periodic multilayer lattice material, Engineering 3 (2011) 1149–1154. [33] N. Hrabe, T. Quinn, Effects of processing on microstructure and mechanical proper-
[10] J.A. Harris, R.E. Winter, G.J. Mcshane, Impact response of additively manufactured ties of a titanium alloy (Ti-6AI-4V) fabricated using electron beam melting (EBM),
metallic hybrid lattice materials, Int. J. Impact Eng. 104 (2017) 177–191. part 1: distance from build plate and part size, Mater. Sci. Eng. A 573 (3) (2013)
[11] I. Ullah, M. Brandt, S. Feih, Failure and energy absorption characteristics of advanced 264–270.
3D truss core structures, Mater. Des. 92 (2016) 937–948. [34] N. Hrabe, T. Quinn, Effects of processing on microstructure and mechanical proper-
[12] R. Hasan, R.A.W. Mines, E. Shen, et al., Comparison of the drop weight impact perfor- ties of a titanium alloy (Ti–6Al–4V) fabricated using electron beam melting (EBM),
mance of sandwich panels with aluminum honeycomb and titanium alloy micro lat- part 2: energy input, orientation, and location, Mater. Sci. Eng. A 573 (3) (2013)
tice cores, Appl. Mech. Mater. 24-25 (2010) 413–418. 271–277.
[13] L.R. Meza, S. Das, J.R. Greer, Strong, lightweight, and recoverable three-dimensional [35] V.S. Deshpande, N.A. Fleck, M.F. Ashby, Effective properties of the octet-truss lattice
ceramic nanolattices, Science 345 (6202) (2014) 1322–1326. material, J. Mech. Phys. Solids 49 (8) (2001) 1747–1769.
N. Jin et al. / Materials and Design 169 (2019) 107655 11

[36] V.S. Deshpande, M.F. Ashby, N.A. Fleck, Foam topology: bending versus stretching [39] C.E. Wen, Y. Yamada, K. Shimojima, et al., Novel titanium foam for bone tissue engi-
dominated architectures, Acta Mater. 49 (6) (2001) 1035–1040. neering, J. Mater. Res. 17 (10) (2002) 2633–2639.
[37] N. Bekoz, E. Oktay, Effects of carbamide shape and content on processing and prop- [40] H. Wang, Qingfen Li, et al., Fabrication and compression properties of Ti honeycomb
erties of steel foams, J. Mater. Process. Technol. 212 (10) (2012) 2109–2116. [J], Aerospace Mater. Technol. 37 (4) (2007) 42–45.
[38] N. Bekoz, E. Oktay, High temperature mechanical properties of low alloy steel foams
produced by powder metallurgy, Mater. Des. 53 (1) (2014) 482–489.

You might also like