Vous êtes sur la page 1sur 20

Cubic negative stiffness lattice structure for energy absorption: numerical and experimental studies

Accepted Manuscript

Cubic negative stiffness lattice structure for energy absorption:


numerical and experimental studies

Chan Soo Ha, Roderic S. Lakes, Michael E. Plesha

PII: S0020-7683(19)30309-9
DOI: https://doi.org/10.1016/j.ijsolstr.2019.06.024
Reference: SAS 10414

To appear in: International Journal of Solids and Structures

Received date: 3 October 2018


Revised date: 19 June 2019
Accepted date: 25 June 2019

Please cite this article as: Chan Soo Ha, Roderic S. Lakes, Michael E. Plesha, Cubic negative stiffness
lattice structure for energy absorption: numerical and experimental studies, International Journal of
Solids and Structures (2019), doi: https://doi.org/10.1016/j.ijsolstr.2019.06.024

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service
to our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and
all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Cubic negative stiffness lattice structure for energy


absorption: numerical and experimental studies
Chan Soo Ha1,* , Roderic S. Lakes1 , and Michael E. Plesha1

T
1
Department of Engineering Physics, University of Wisconsin - Madison

IP
*
Corresponding author: ha3@wisc.edu Phone: +01 608 890 0261

CR
June 26, 2019

Abstract
US
We present a new type of a truly three-dimensional cubic negative stiffness lattice structure
that can achieve energy absorption and recover its original configuration under cyclic loading
AN
in excess of a strain of approximately 20 %. This structure was composed of multiple unit
cells exploiting negative stiffness from controlled elastic buckling. Structural properties were
designed to be tuned by adjusting geometry. The effective Poisson’s ratio was equal to zero
regardless of the constituent material. Geometric parameters that can lead to the desired
M

energy absorption without yielding the structure were determined by finite element analysis.
We then fabricated samples representing the lattice structure with different sizes through
additive manufacturing and performed cyclic loading experiments to capture stress-strain
ED

hysteresis loops. Results clearly showed that the designed structure was capable of absorbing
mechanical energy effectively with a full recovery of geometry in three principal directions
and that the amount of energy absorbed during cyclic loading increases with its size.
PT

1 Introduction
CE

Recently, carefully designed cellular structures have been of considerable interest in various
engineering applications due to their unique and beneficial properties such as light-weight
[1] [2] [3], high stiffness [3] [4], high damping [4] [5], tunable Poisson’s ratio, including neg-
AC

ative values [6] [7] [8], controllable thermal expansion [9] [10] [11], and negative stiffness [4]
[5] [12] [13]. These properties are not achievable directly from natural materials and can
provide access to previously unoccupied material design space. In particular, mechanical
energy absorption (or damping) recently has become an important aspect in many daily life
applications such as rubber shoe soles, car bumpers, and bicycle helmets [14]. For example,
the shoe soles absorb energy via viscoelastic effects; stress and strain of viscoelastic mate-
rials depend on time [15]. Car bumpers and bicycle helmets absorb large impact energies
during a short period of time through destructive deformations of foams or architected mate-
rials made of plastics or metals. Conceptually, an ideal energy absorbing material would be
1
ACCEPTED MANUSCRIPT

lightweight, tailorable, and capable of desired energy absorption in a structurally repeatable


manner without time-dependency of stress and strain.
Although a mechanical instability such as buckling and negative stiffness has been con-
sidered as an unwanted behavior or a weakness, it has been investigated that architected
materials incorporating such an instability can give rise to energy absorption [16] [17] [18]
[19] [20] [21] [34] [22] [23] [24] [25]. The idea of negative stiffness has been also exploited
for similar engineering purposes such as shock/vibration isolation and damping [26] [27]
[28] [29] [30] and sensitivity enhancement [31] [32]. However, in most of these studies, the
ability of the material design for the purpose of energy dissipation is somewhat restricted

T
with respect to the allowable directions and/or the magnitude of deformation. For example,

IP
honeycomb materials with negative stiffness and tailored buckling microlattices can only
deform in a single direction under loading and hence are not pertinent for multi-axial energy

CR
absorption. In recent years, a number of studies has proposed material designs that can
achieve multi-axial energy absorption [31] [33] [34] [35]. Other beneficial properties utiliz-
ing the mechanical instability include pattern transformation [36], auxetic deformation [37],
tailorable metamterials with multi-functionalities [16], origami folding with bistability [38],

US
and shape-reconfigurable materials with multistability [17] [18] [35].
The current work builds upon our previous study [25] that utilizes negative stiffness
components generated by geometric nonlinearity to achieve an uniaxial energy absorption.
AN
In this work, we present a novel design of a truly three-dimensional cubic lattice structure
comprising multiple negative stiffness (NS) unit cells for the purpose of energy dissipation
in three principal directions. The NS unit cell was designed with controlled elastic buckling
in the constituent beams caused by geometric nonlinearity, resulting in energy absorption
M

with self-recovery via snap-through response. Finite element (FE) analysis of the NS unit
cell was performed to study the effects of geometric parameters. Several lattice structures
composed of multiple NS unit cells were fabricated by using selective laser sintering (SLS)
ED

for the proof-of-principle purpose. Laboratory experiments were conducted to investigate


their energy absorption capability in response to loading-unloading cycle, and the amount
of absorbed energy ∆E and the loss coefficient Ψ were quantified from load-displacement
PT

curves.

2 Design
CE

2.1 Negative Stiffness Unit cell


AC

Negative stiffness (NS) unit cell having a cubic symmetry was developed, as illustrated in
Figure 1. This unit cell is comprised of three constituents: namely, frame, beam, and plate.
The frame refers to the cube shaped set of ribs. The plate refers to the base of the cylindrical
stalks. The beams connect the frame corners with the plate at the base of the stalks. Hence,
one unit cell is comprised of 24 frame members, 6 plates, and 24 beams. Both the frame
and plate were designed with comparably larger dimensions than those of the beams. This
results in most of deformation to occur within the beams and prevents undesirable transverse
expansion of the NS unit cell upon application of an orthogonally oriented compressive load.
Hence, this configuration helps to avoid other unwanted deformations to occur prior to the

2
ACCEPTED MANUSCRIPT

desired buckling of the beams, and negative stiffness can be developed as a result of snap-
through response. Without these treatments, the beams would not experience enough axial
force to initiate buckling, and negative stiffness behavior would be lost. Although the design
of the NS unit cell was inspired by bistable mechanisms, we designed such a unit cell to
exhibit a full recovery with a single stable state. This allows a creation of lattice structures
having multiple NS unit cells that can recover its initial configurations without requiring
an externally applied load. Furthermore, rigid (i.e., fully fixed) connections between all
constituents of the NS unit cell were considered for purposes of 3D printing fabrication. It
is worth noting that the NS unit cell is cubic symmetric, so it is expected to generate a

T
material with nearly cubic elastic behavior [39].

IP
aframe - 2bbeam
a [001]
b

CR
aframe
Plate Lstalk

tframe

Frame USdstalk aplate


AN
bbeam

aframe + tanα(aframe - aplate) + tplate


M

[010]

Stalk
c Lbeam
2 aplate - bbeam
hbeam Lbeamsinα
ED

[100] α
Beam
2 aframe - bbeam

Figure 1: Schematics of the negative stiffness unit cell shown with geometric parameters.
PT

(a) Isometric view. (b) Side view. (c) View of one of the faces in the [110] direction.
CE

Geometric parameters defining the NS unit cell include the height of the beam hbeam , the
width of the beam bbeam , the inclined angle α, the length of the beam Lbeam , the inclined
height of the beam h, the side length of the plate aplate , the thickness of the plate tplate , the
side length of the frame aframe , and the thickness of the frame tframe , as illustrated in Figure
AC

1.
From geometry, it is straightforward to determine the following relationships; Lbeam =
aframe −aplate

2 cos(α)
, h = Lbeam sin(α), Lcell = aframe + √52 tan(α)(aframe − aplate ) + tplate . In addition,
it is possible to derive an analytical expression for the relative density of the NS unit cell
consisting of 24 beams, 24 frames, 6 plates, and 6 stalks. The relative density can be
expressed as
24Vbeam + 24Vframe + 6Vplate + 6Vstalk
ρ̄ = (1)
[aframe + √52 tan(α)(aframe − aplate ) + tplate ]3
3
ACCEPTED MANUSCRIPT

bbeam hbeam (aframe −aplate ) √


where Vbeam = √
2 cos(α)
, Vframe = [a frame + (1 − 2)bbeam ]t2frame , Vplate = (a2plate −

bbeam )hbeam , and Vstalk = 3π 8tan(α)

2
(aframe − aplate )( 2aplate − bbeam )2 . Note that mass accumu-
lation at the nodes is ignored in Equation 1.
In theory, the mechanical response of the NS unit cell is expected to depend on geometric
parameters of the constituent beams because the resulting deformation is designed to occur
mainly within these members during the process of snap-through response. In other words,
a load-displacement curve of the NS unit cell can be tailored by tuning beam geometries
(bbeam , hbeam , Lbeam , and α). Hence, tuning of an aspect ratio of the beam λ = √ KLbeam
Ibeam /Abeam

T
and the inclined angle α will be emphasized in this work for the determination of a geometric

IP
parameter set that can lead to designing optimally efficient NS lattice structures. Note that
1
K = 1/2 (for fixed connections), Ibeam = 12 bbeam h3beam , and Abeam = bbeam hbeam .

CR
2.2 Negative Stiffness Lattice Structure
To create lattice structures capable of energy absorption, a series of identical NS unit cells

US
in three principal directions were stacked together by utilizing stalks having a circular cross
section, as shown in Figure 2. The addition of the stalks provides adequate space between
adjacent unit cells when the structure reaches its fully reversed configuration during the
process of snap-through behavior in response to loading-unloading cycle. Furthermore, each
AN
stalk was modeled as a relatively rigid member compared to the beams of the NS unit
cell so as to transmit only loads to adjacent unit cells and√ not to undergo any appreciable
deformations. The radius of the stalk rstalk was taken as 22 aplate − 12 bbeam , and the length of
M

the stalk Lstalk was equal to 32 h. This led to the ratio of a cross sectional area of the stalk
to the area of the plate of the NS unit cell equal to approximately 87.5 %, hence it was
reasonable to ignore deformation in the constituent plate in analysis.
ED
PT
CE
AC

4
ACCEPTED MANUSCRIPT

a b

T
IP
z
z

CR
y x, y
x
US
Figure 2: Schematics of the negative stiffness lattice structure consisting of three unit cells
per side. The lattice structure is created by assembling the cells in the three principal
directions.
AN

3 Method and Fabrication


M

3.1 Finite Element Analysis


Negative stiffness (NS) lattice structures presented in this work experience buckling and post-
ED

buckling behavior when subjected to a vertically oriented compressive load. This buckling
behavior gives rise to snap-through response and hence provides energy absorption capability.
To analyze such response, geometric nonlinearity for large deformations and instability must
PT

be considered in FE analysis. In the present study, a commercial FE code ANSYS APDL


19.0 was used, and the formulation of the FE modeling are as follows.
CE

3.1.1 Analysis Type


Analysis type was static, and large deflection effects were included. The latter accounts for
modeling of geometric nonlinearities and stress stiffening effects, if present. The sparse direct
AC

solver was used for most cases, while the full Newton-Raphson iteration method without the
adaptive descent technique was adopted if there is nonlinearity present. Automatic pseudo-
time stepping was considered to ensure that the pseudo-time step variation was neither too
aggressive nor too conservative. The default setting for convergence criteria was used with
line search option for better convergence.

5
ACCEPTED MANUSCRIPT

3.1.2 Finite Element Modeling of Negative Stiffness Lattice Structure


FE models representing the NS unit cell is shown in Figure 3. The particular finite elements
used were the BEAM189 element to mesh both beam and frame members, and SHELL181
was used for plates. These are flexural finite elements having 6 degrees of freedom (DOF)
per node and are capable of linear, large rotation, stress stiffening, and large strain nonlinear
responses. Furthermore, having 6 DOFs at each node enables a smaller overall number of
elements in a structure as compared to employ continuum finite elements having 3 DOFs
per node.

T
With reference to Figure 3(a), a FE model representing the NS unit cell consists of 24
beam members, 24 frame members, 6 plate members, and affixed to each plate is a short

IP
stalk member (6 in total). Each beam member was modeled using four beam finite elements,
each frame member was modeled using three beam finite elements, each plate member was

CR
modeled using 24 shell finite elements, and each stalk was modeled using one beam finite
element. Hence, the finite element model for a single unit cell consists of 174 beam finite
elements and 144 shell finite elements. The FE models employed a linear elastic material

US
model with the measured material properties of PA as base material (E = 0.717 GPa, σys =
31.6 MPa, and ν = 0.24; please see Section 3.2 for details). It is worth noting that the use
of linear elastic material model in FE models would not affect a detection of snap-through
response caused by negative stiffness because such nonlinear behavior even occurs if the
AN
constituent material is elastically linear.
To represent an infinite lattice structure in which each NS unit cell has exactly the
same deformed shape, the following boundary conditions (as shown in Figure 3(b)) were
applied. A center-node of a stalk of the unit cell on +z face was subjected to a downward
M

displacement δ and was constrained in x and y translations and rotations about all directions;
an opposite center-node on −z face was fixed; all other nodes were unconstrained. Such
ED

boundary conditions simulate a fully bonded condition in conjunction with the prescribed
displacement, which would be observed in laboratory experiments exciting loading-unloading
cycle. Moreover, displacement control compression (i.e., the prescribed displacement) was
utilized to capture negative stiffness (i.e., negative slope portion of a load-displacement
PT

curve), and such stiffness most likely would not be displayed if force-controlled loading was
applied; the latter loading configuration would show a plateau region between peak loads
(i.e., points of zero slope in the load-displacement curve).
CE
AC

6
ACCEPTED MANUSCRIPT

δ
a b

T
IP
z

CR
z
y
x x, y
US
Figure 3: (a) Finite element model of the negative stiffness unit cell. (b) Load and boundary
AN
conditions applied to finite element model.

3.2 3D Printing Fabrication


M

For purposes of conducting loading-unloading cyclic compression experiments, several phys-


ical models representing the NS lattice structures were fabricated using a DTM 2500 Plus
performing SLS. This 3D printer provides a typical layer thickness of 0.1016 mm with the
ED

diameter of a laser (power source to sinter powdered material) equal to 0.254 mm, tolerances
of ±0.127 mm for the first 25.4 mm and ±0.0508 mm for each additional 25.4 mm. A base
material used for fabrication was polyamide (PA), and this is a typical material for SLS
PT

3D printing method. Material properties of the base material were measured from a tensile
test according to ASTM-D638-14 using a screw-driven test frame (Sintech 10/GL) utilizing
a load cell with 10,000 lb capacity. Overall dimensions of a standard dumbbell-shaped test
specimen were 19 mm × 165 mm × 2 mm. At room temperature, the measured material
CE

properties of PA were E = 0.717 GPa, σys = 31.6 MPa, ν = 0.24, and ρ = 0.962 g/cm3 .
Figure 4 shows a SLS printed NS lattice structure with three unit cells per side.
AC

7
ACCEPTED MANUSCRIPT

a b

T
IP
CR
10 mm 10 mm

US
Figure 4: Photographs of 3×3×3 SLS printed negative stiffness lattice structure with a side
length of approximately 80 mm. (a) Isometric view. (b) Side view.
AN
3.3 Experiment
All loading-unloading cyclic compression tests were performed with a universal testing frame
M

(MTS QTest/5) utilizing a MTS load cell (Model #: 4501009-B) with a load capacity of
500 N in order to reveal stress-strain hysteresis loops of the fabricated NS lattice structures.
The load cell has a load precision of ±0.001 N, and was calibrated on Apr. 15th, 2016. The
ED

loading-unloading cycle was enabled by periodically applying loading and unloading. Each
lattice was subjected to the maximum displacement equal to 4Nlayer h, which gives rise to a
fully reversed configuration without yielding the lattice. For each test, a preload of 0.1N was
PT

applied to all lattices, the load cell moved at constant testing speed of 4 mm/min, and a data
acquisition rate was 10 Hz. Load-displacement data was recorded for one complete loading-
unloading cycle. Engineering strain  and stress σ were computed as  = ∆L/Lunit cell ,
where ∆L is the displacement measured by the load cell, and σ = F/Aeff , where F is the
CE

load measured by the load cell and Aeff is the effective cross section area of the lattice
(Aeff = (Ncell Lcell )2 ). All lattices were tested without external supports (i.e., grips), because
their original configuration were designed to self-recover from the fully reversed configuration
AC

upon the removal of applied load.


The hysteresis loop enclosed by different loading and unloading paths in load-displacement
curves represents the amount of dissipated mechanical energy (in Joule) by the NS lattice
structures; one can obtain a similar quantity from the stress-strain data, and this provides
the dissipated energy per unit volume. In this work, we investigated the former quantity
by numerically integrating load-displacement curves using trapezoidal integration in Python
and adopted the loss coefficient, defined as Ψ = ∆E/E, for evaluating the performance of
energy absorption of the structure, where E and ∆E are the stored and dissipated energy,
respectively.
8
ACCEPTED MANUSCRIPT

4 Results
4.1 Determination of Geometric Parameters of Negative Stiffness
Unit Cell
A determination of optimized geometric parameters for the negative stiffness (NS) unit cell
was first performed by FE analysis. As discussed in Section 2.1, the mechanical performance
of the NS unit cell thoroughly depends on the beam aspect ratio λ and the inclined angle α,
because these variables essentially control the buckling and post-buckling behavior during

T
snap-through process. Hence, a study on the effect of these beam design parameters was

IP
emphasized here to determine the optimized geometric parameters.
It is worth noting that there are no theoretical limitations on choosing the geometric
parameters of the NS unit cell to exploit snap-through response leading to energy absorption.

CR
However, due to fabrication limitations (such as print resolution and tolerances) of SLS,
which was employed in the present study to build physical models for laboratory experiments,
we were limited when selecting geometric parameters. Consequently, for the purpose of

US
fabricating a minimal structure, these parameters were chosen based on the smallest size
feature of the NS unit cell which was the height of the beam hbeam . Since the minimum
resolution of the printer is approximately 0.4 mm, we set the minima of hbeam as 0.5 mm
AN
and bbeam as 2 mm to ensure a reasonable margin for printing quality in the physical models.
Other parameters such as tplate and tframe were chosen so that these structural elements are
relatively rigid compared to the constituent beams; that is, tplate = 0.5 mm and tframe = 1
mm. The rationale is the frames and plates are only for preventing unwanted expansion of
M

the NS unit cell in response to an orthogonally oriented compressive load. Although there
are perhaps substantial sets of geometric parameters, we have limited our study to aplate
= 4 mm, and have varied aframe from 10 mm to 18 mm to probe the effect of the beam
ED

aspect ratio λ of the constituent beams on the performance of energy dissipation, and this
consequently led to λ varying from 15 to 35.
We specifically explored geometric parameters (Figures 5(a)-5(e)) ranging from α = 1◦
to 10◦ and λ = 15 to 35 with an increment of 5. Each NS unit cell was compressed up
PT

to 4h. This displacement gives rise to a fully reversed configuration of the NS unit cell
and is equivalent to approximately a strain of 20 %. The x axis represents the normalized
displacement defined as the ratio of applied displacements d to dmax (=4h), while the y axis
CE

denotes load measured by the load cell of the test frame. Curves with blue highlight denote
sets of geometric parameters that can give rise to negative stiffness (i.e., negative slope)
during loading. Figure 5(f) illustrates a map of relative density of the NS unit cell in terms
AC

of the beam aspect ratio and the inclined angle. Relative density was found to be inversely
proportional to both the inclined angle and the beam aspect ratio.

9
ACCEPTED MANUSCRIPT

a b c
Load F [N]

T
10

IP
d e f

Inclined angle α [degree]


8

CR
Load F [N]

US 2

15 20 25 30 35
AN
Normalized displacement d/dmax [-] Beam aspect ratio λ [-]

Figure 5: Load displacement relationships of negative stiffness unit cells with (a) λ = 15,
(b) λ = 20, (c) λ = 25, (d) λ = 30, and (e) λ = 35, which are obtained from finite element
M

simulations. Curves highlighted by blue color denote sets of geometric parameters resulting
in the desired negative stiffness response in the unit cell. The inclined angle varies from 1◦
to 10◦ . (f) Relative density map in terms of λ and α.
ED

According to FE simulations, a sufficiently large inclined angle was necessary to develop


both snap-through response and negative slope in these curves; for example, for λ = 35,
inclined angles larger than 5◦ exhibited snap-through as well as negative stiffness. Adequate
PT

values of λ were also needed to produce considerable peak-to-peak difference in load magni-
tude (i.e., difference in load between zero-slope points), and an increase in this difference was
CE

found to be proportional to λ. These findings physically make sense since a higher λ results
in an increase in Lbeam while holding bbeam and hbeam constant, and hence the beam becomes
capable of a larger elastic deformation. In addition, it was observed that the NS unit cell
produced negligible transverse deformation in response to orthogonally oriented compres-
AC

sive load; the ratio of the transverse to compressive deformation was about one hundredth.
Note that this was expected since the cell was designed to de-couple displacements in three
principal directions. As a result, the resulting Poisson’s ratio was found to be nearly zero
and independent of strains regardless of the constituent material. We also discovered that
load-displacement curves from either loading or unloading were identical for the unit cells.
This is because FE models use a material model that is perfectly homogeneous, linear, and
isotropic, and because the applied boundary condition represents perfectly balanced fixed
constraint. Nonetheless, when physical models of the lattice structure comprising multiple
NS unit cells are subjected to loading-unloading cycle, two distinct loading and unloading
10
ACCEPTED MANUSCRIPT

curves are expected to appear due to instability of the constituent beams of the cells in
the lattice structure. This causes marginal differences in buckling strength of the individual
beams and hence creates minute differences between beam buckling. Details will be discussed
later (see Section 4.2).
To choose the optimized geometric parameters of the NS unit cell resulting in desirable
energy absorption capability, the maximum von Mises stress of the NS unit cell at its fully
reversed state and the amount of energy produced during loading were computed and plotted
in Figure 6. Both of these quantities increased with the inclined angle but not in a linear
relation. Although various sets of geometric parameters can be chosen for the NS unit

T
cell capable of energy absorption without yielding, we have chosen λ = 35 and α = 8◦

IP
(resulting in ρ̄ = 0.0497) to fabricate a minimal structure. This ensures that fabricated
physical models do not suffer from the possibility of having random porosity and unexpected

CR
dimensional inconsistency that may occur during fabrication process, which would result in
unwanted damage to the models during deformation. In addition, for the selected geometric
parameters, the corresponding σmax /σys was approximately 0.6.

a 10 σmax/σys [-]
1
US b 10 Eloading [mJ]

6
Inclined angle α [degree]

snap-through
Inclined angle α [degree]

AN
8 8
0.5

6 6 4
M

4 4
ED

2
2 no snap-through 0.1 2

15 20 25 30 35 15 20 25 30 35
PT

Beam aspect ratio λ [-] Beam aspect ratio λ [-]

Figure 6: Effect of the inclined angle α and the beam aspect ratio λ = √KLbeam (where
Ibeam /A
CE

K = 1/2) on the normalized stress (σmax /σys ) and stored energy by the negative stiffness
unit cell obtained from finite element simulations. (a) Snap-through map represented by the
ratio of the maximum von Mises stress of the unit cell at its fully reversed state to the yield
AC

stress of base material in terms of λ and α. The black dashed line represents the numerically
measured transition between the geometries that give rise to snap-through response and
those that do not result in snap-through instability. (b) Map of the amount of stored energy
during loading in terms of λ and α.

4.2 Experiment
Although lattice structures consisting of multiple NS unit cells were designed to generate
nearly cubic elastic behavior due to their cubic symmetry, it was seen that physical mod-
11
ACCEPTED MANUSCRIPT

els fabricated by SLS did not manifest such a behavior due to nonuniform print resolution
in different directions. This difference in the print resolutions most likely gives rise to ei-
ther orthotropic or transversely isotropic structural properties in different loading directions,
assuming a print resolution in transverse directions is different from that in printing direc-
tion. Hence, loading-unloading cyclic compression tests over the complete cycle in these
two different directions (i.e., the printing and transverse directions) were performed, and the
corresponding stress-strain relationships were measured, as depicted in Figure 7.
NS lattice structures successfully produced the hysteresis loop developed by different
loading and unloading paths in stress-strain responses. For all physical models, the enclosed

T
area representing the energy absorbed by the NS lattice structure over the complete cyclic

IP
of loading and unloading was computed, as summarized in Table 1. As illustrated in Figures
7(a) and 7(b), it was observed that a FE model of a single unit cell slightly overestimated

CR
experimental measurements of a similar topology. Again, the authors believe that this is
mainly due to the fabrication imperfection such as dimensional inconsistency. Nonetheless,
it is reasonable to state that the FE simulations made a good agreement with results ob-
tained from the experiment because a similar trend between these two approaches was seen.

US
All physical models produced the maximum allowable strain approximately equal to 20 %
at their fully reversed state and recovered to their original configuration without the need of
external load, as depicted in Figures 7(b)-7(f). Moreover, we discovered a marginal steepen-
AN
ing behavior of the stress-strain curves for samples with Ncell > 2 (i.e., an increase in local
maxima of the curve), which is commonly revealed in structures with dimensional gradients
[24]. It was observed that the constituent beams of the NS unit cells buckle sequentially
but not in a particular order. This seems to be affected by relative imperfections from 3D
M

printing fabrication in one or more beams of the NS unit cells, leading them to buckle under
slightly less force thresholds than those in other layers. Furthermore, there was no difference
in a buckling pattern in the lattice structures with a different number of the cells. For all
ED

samples, a layer of the cells close to either the load cell on the top boundary or a support on
the bottom boundary appears to deform first, and beam buckling continues to move toward
the center of the lattice. However, within a layer, the cells deforms randomly (not in a specific
PT

order). This is mainly due to instability caused by buckling and fabrication imperfections as
discussed previously. In experiments, we applied a cyclic compression loading to the samples
with a specified strain so that all the constituent beams can achieve their elastic post-buckle
conditions locally. Moreover, the samples were tested to exert buckling without experienc-
CE

ing material yielding for the purpose of an energy dissipation with an elastic full recovery,
and thus a global buckling (i.e., large scale buckling) was not observed. If a strain applied
to the samples was exceeded the maximum allowable displacement of 4Nlayer h designed for
AC

the fully elastic recovery, frame members aligning with the direction of an applied loading
would buckle due to contacts between adjacent layers (similar with densification) while the
constituent beams still remain at their elastic post-buckling state, and this phenomenon is
still in a category of (small scale) local buckling.
As the size of the NS lattice structure increases, the performance of energy dissipation of
the NS lattice structure appears to be improved. Although we have limited the size of the NS
lattice structure up to three unit cells per side, this finding is sufficiently verified in a sense
that the amount of energy absorbed increases with its size; Ψ ∼ 0.40 for Ncell = 1, Ψ ∼ 0.51
for Ncell = 2, and Ψ ∼ 0.62 for Ncell = 3. Moreover, an ideal number of snap-through
12
ACCEPTED MANUSCRIPT

responses would be equal to 2Nlayer for one complete loading-unloading cycle, however a
deviation from this case was seenin stress-strain responses due to a fact that some beams
deformed simultaneously while others buckled at different times. In addition, the physical
models loaded in the print direction achieve a slightly higher stress than those loaded in the
transverse direction, but the difference between these two cases was negligible. The amount
of absorbed energy, by contrast, displayed slightly larger values for NS structures loaded in
the transverse direction as compared to those tested in the print direction except for Ncell =1.
Therefore, there is modest mechanical orthotropy as a result of 3D printing of a structure
with designed cubic symmetry.

T
The effects of contact conditions were also studied by considering two different end con-

IP
ditions of the lattice structure samples fabricated by SLS: namely, flat and half-sphere ends.
Whereas the flat ends simulate built-in constraints (no rotations at contacts), the half-sphere

CR
ends behave similar to pinned constraints where rotations are allowed to occur. Our exper-
imental results showed that the differences in the stress-strain responses between these two
contact conditions were negligible. It seems that end tilt generated by the flat ends is coun-
teracted by random buckling of the constituent beams of the NS lattice structure. The FE

US
simulations performed in this work, however, do not capture the tilt of the flat ends as the
beams buckle. To capture this behavior, FE models most likely require to entail more DOFs
than that of a practically given the number of DOF already in the present FE models. How-
AN
ever, this tilt is in different directions. If these are averaged, the simulation will be more
similar to the experiment. For this reason, our FE simulations used built-in conditions as
described in Section 3.1.2; that is, all stalks on the bottom of the NS unit cells in the lattice
structure have all six DOF constrained, while all stalks on the top of the structure have five
M

DOF constrained plus a prescribed displacement for the remaining DOF.

Table 1: Experimentally obtained amount of energy absorbed by the negative stiffness lattice
ED

structures with two different end conditions in printing and transverse directions.

End Loading
Size Sample Mass m [g] ∆E [mJ] ∆E/m [J/kg] Ψ [-]
condition direction
PT

1×1×1 A 0.447 Flat Printing 0.98 2.19 0.434


Transverse 0.71 1.48 0.372
CE

2×2×2 B 3.588 Flat Printing 6.91 1.93 0.465


Transverse 6.28 1.75 0.524
C 3.883 Half-sphere Printing 6.22 1.60 0.479
AC

Transverse 8.05 2.07 0.637


3×3×3 D 11.931 Flat Printing 32.23 2.70 0.618
Transverse 38.69 3.24 0.648
E 12.934 Half-sphere Printing 33.02 2.55 0.629
Transverse 34.07 2.63 0.586

13
ACCEPTED MANUSCRIPT

1.5
Stress σ [kPa] a c Printing direction
e Transverse direction

0.75

0
1.5
b d

T
Printing direction Printing direction
f Transverse direction
Stress σ [kPa]

IP
0.75

CR
0
0 10 20 0 10 20 0 10 20

g
Strain ε [%]

h
US
Strain ε [%]

i
Strain ε [%]
AN
M

10 mm
ED
PT

Undeformed Intermediate Fully deformed


CE

Figure 7: Stress-strain hysteresis loops of negative stiffness lattice structures with a different
number of cells per side. (a) FE simulation of a single unit cell; an evolution of deformation
configurations at different strains is also shown. (b) Single unit cell in printing direction
AC

(experiment). (c)-(d) 2×2×2 and 3×3×3 lattices structure in printing direction, respec-
tively (experiment). (e)-(f) 2×2×2 and 3×3×3 lattices structure in transverse direction,
respectively (experiment). Blue and red curves represent responses of samples with the flat
and half-sphere end conditions, respectively. All samples were capable of self-recovery with
a desired energy dissipation after the removal of loading. (g)-(i) Photographs displaying an
evolution of deformations of a 3×3×3 negative stiffness lattice structure.

14
ACCEPTED MANUSCRIPT

5 Discussion
The findings presented in this work provide a guidance for designing a truly three-dimensional
cubic negative stiffness (NS) lattice structure that is capable of multi-directional energy
absorption for various engineering purposes. A typical objective of the present work would
be an isolation from multi-axial dynamic mechanical loading that is unexpected or greater
than a predetermined threshold Our design exhibits mechanical energy dissipations in the
three principal directions (i.e., x, y, and z directions) independently in such a way that
such dissipations can only occur in direction(s) where an external stimulus is present. For

T
example, if the lattice structure is subjected to cyclic compression loading in the x and

IP
y directions, mechanical energy associated in these directions are dissipated while there is
no such a dissipation in the z direction. Thus, our design is capable of energy dissipation
in the direction(s) of interest while providing no sensible deformation in other direction(s).

CR
Moreover, this decoupling can lead to nearly zero Poisson’s ratio if the design is subjected
to a uniaxial loading. In this work, we refer such an energy dissipation phenomenon to a
multi-directional energy absorption. This energy dissipation phenomenon is different from

US
studies proposed in Refs. [22] [24] [25] [28] which are restricted to deform in one direction
and hence is only capable of a uni-directional energy absorption.
We have in fact performed FE simulations for the NS lattice structures (having more
AN
than one cell per side; Ncell > 1) in response to a loading-unloading cyclic compression.
Even though effects of geometric nonlinearity were included in FE simulations, we observed
deviation in the load-displacement relationships between FE and experimental results. This
discrepancy can be explained as follows. 1) We have discovered that the base material (which
M

was PA) used for fabrication of physical models via SLS behaves like a viscoelastic material
[25]. Such a material is sensitive to time (i.e., time-dependent) whereas in our FE simulations
the material is assumed to be elastically linear with no viscoelastic effects. Consequently, this
ED

difference may have led to the deviation in the load-displacement relationships obtained from
these two approaches. 2) Since the base material is viscoelastic, strain rate during loading
and unloading can produce profound differences in the effective stiffness and strength of
the physical models representing the NS lattice structures. Thus, it is sensible to say that
PT

mechanical properties of the fabricated lattice structures depend on the rate of application
of the load; cyclic loading experiments performed in this work, however, employed a single
strain rate. To investigate this, one can consider several strain rates for experiments to
CE

investigate a detailed characterization of viscoelastic behavior of the NS structure. 3) The


quality of physical models fabricated by SLS manifested inconsistent geometric dimensions
and random porosity; in fact this topic has been actively studied recent studies (see Ref. [25]
AC

[40] [41] [42]). Again, an energy absorption mechanism adopted in the NS lattice structure
presented in this work exploits elastic buckling of the constituent beams in which bending
rigidity plays significantly during deformation. Consequently, dimensional inconsistency and
unexpected porosity from fabrication via SLS can directly affect the performance of the
NS lattice structure, and this likely accounts for discrepancies between the FE predictions
and the experiments. By considering these three aspects discussed above, the authors have
decided to present only experimental results for the NS lattice structures.
There are opportunities for future work. One example is to study a size effect of the
NS lattice structure with sufficient numbers of unit cells per side. The structure shows
15
ACCEPTED MANUSCRIPT

nonlinearity with regard to both the maximum stress at its fully reversed configuration
and the amount of absorbed energy. Moreover, this structure is compliant in tension (or
compression) and bending but is resistant to torsion. These are unusual in classical solids. A
more generalized theory such as Cosserat elasticity can perhaps characterize such a structure
better than a classical one, since Cosserat elasticity incorporates a local rotation of points
as well as typical translational degrees of freedom. Another possible future work is to study
quasistatic and impact behavior of the NS lattice structure. This may provide its energy
absorption characteristics (such as thresholding behavior) in response to various loading
conditions.

T
IP
6 Conclusion

CR
In this present work, we have developed and examined a truly three-dimensional cubic neg-
ative stiffness (NS) lattice structure that can achieve energy absorption capability in three
principal directions, using FE analysis and laboratory testing of samples produced by selec-

US
tive laser sintering. The designed NS structure is different from regular lattice structures
(which most likley undergo destructive energy absorption (e.g., honeycombs)) in a sense
that it can absorb mechanical energy with self-recovery utilizing controlled elastic buckling
phenomena with snap-through response. Hence, the NS lattice structure is capable of a full
AN
recovery upon the removal of loading and energy absorption simultaneously with a strain of
approximately 20 %. In addition, it exhibited decoupling of axial deformation from trans-
verse deformation, therefore Poisson’s ratio is equal to zero regardless of the constituent
M

material. Structural properties were designed to be tailored by adjusting geometric param-


eters. Experimental results have clearly displayed that the present structure is capable of
absorbing mechanical energy with a full recovery of geometry upon unloading and that the
ED

amount of absorbed energy increases with the size of the lattice structure. We envision
this newly developed NS structure with energy absorption capability will find applications
that require a full recovery upon successive impact or shock/vibration isolation from multi-
directional unexpected dynamic loading and further can lead to designs of new engineered
PT

materials with multi-functionality.


CE

Acknowledgments
We gratefully acknowledge support of this research by the ACS Petroleum Research Founda-
tion via Grant 52695-ND10 and by ARO U.S. Army Research Office under Grant W911NF-
AC

13-1-0484.

References
[1] R. S. Lakes, Materials with structural hierarchy, Nature, 361 (6412), 511 (1993).

16
ACCEPTED MANUSCRIPT

[2] T. A. Schaedler, A. J. Jacobsen, A. Torrents, A. E. Sorensen, J. Lian, J. R. Greer, L.


Valdevit, W. B. Carter, Ultralight metallic microlattices, Science, 334 (6058), 962-965
(2011).

[3] X. Zheng, H. Lee, T. H. Weisgraber, M. Shusteff, J. DeOtte, E. B. Duoss, J. D. Kuntz,


M. M. Biener, Q. Ge, J. A. Jackson, S. O. Kucheyev, Ultralight, ultrastiff mechanical
metamaterials. Science, 344 (6190), 1373-1377 (2014).

[4] T. Jaglinski, D. Kochmann, D. S. Stone, R. S. Lakes, Composite materials with vis-

T
coelastic stiffness greater than diamond, Science, 315 (5812), 620-622 (2007).

IP
[5] R. S. Lakes, T. Lee, A. Bersie, Y. C. Wang, Extreme damping in composite materials
with negative stiffness inclusions, Nature, 410 (6828), 565-567 (2001).

CR
[6] R. S. Lakes, Foam structures with a negative Poisson’s ratio, Science, 235, 1038-1040
(1987).

[7] U. D. Larsen, O. Sigmund, S. Bouwstra, Design and fabrication of compliant micromech-

US
anisms and structures with negative Poisson’s ratio. Micro Electro Mechanical Systems,
1996, MEMS’96 Proceedings. An Investigation of Micro Structures, Sensors, Actuators,
Machines and Systems. IEEE, The Ninth Annual International Workshop on, 365-371
AN
(1996).

[8] R. H. Baughman, J. M. Shacklette, A. A. Zakhidov, S. Stafstrm, Negative Poisson’s


ratios as a common feature of cubic metals, Nature, 362, 392 (6674) (1998).
M

[9] T. A. Mary, J. S. O. Evans, T. Vogt, A. W. Sleight, Negative thermal expansion from


0.3 to 1050 Kelvin in ZrW2O8, Science, 272 (5258), 90-92 (1996).
ED

[10] R. S. Lakes, Cellular solids with tunable positive or negative thermal expansion of
unbounded magnitude, Applied physics letters, 90 (22), 221905 (2007).
PT

[11] H. Xu, D. Pasini, Structurally efficient three-dimensional metamaterials with control-


lable thermal expansion, Scientific reports, 6, 34924 (2016).

[12] Z. Hashin, S. Shtrikman, A variational approach to the theory of the elastic behaviour
CE

of multiphase materials, Journal of the Mechanics and Physics of Solids, 11 (2), 127-140
(1963).
AC

[13] R.S. Lakes, W.J. Drugan, Dramatically stiffer elastic composite materials due to a
negative stiffness?, Journal of the Mechanics and Physics of Solids, 50 (5), 979-1009
(2002).

[14] J.C. Dixon, The shock absorber handbook, John Wiley & Sons (2008).

[15] R. S. Lakes, Viscoelastic materials, Cambridge University Press (2009).

[16] B. Florijn, C. Coulais, M. van Hecke, Programmable mechanical metamaterials, Physical


review letters, 113(17), 175503 (2014).
17
ACCEPTED MANUSCRIPT

[17] S. Shan, S.H. Kang, J.R. Raney, P. Wang, L. Fang, F. Candido, J.A. Lewis, K. Bertoldi,
Multistable architected materials for trapping elastic strain energy, Advanced Materials,
27(29), 4296-4301 (2015)

[18] H. Yang, L. Ma, Multi-stable mechanical metamaterials with shape-reconfiguration and


zero Poisson’s ratio, Materials & Design, 152, 181-190 (2018).

[19] S. Schaare, A.V. Dyskin, Y. Estrin, S. Arndt, E. Pasternak, A. Kanel-Belov, Point


loading of assemblies of interlocked cube-shaped elements, International Journal of En-

T
gineering Science, 46 (12), 1228-1238 (2008).

IP
[20] L. Kashdan, C.C. Seepersad, M. Haberman, P.S. Wilson, Design, fabrication, and eval-
uation of negative stiffness elements using SLS, Rapid Prototyping Journal, 18 (3),

CR
194-200 (2012).

[21] E.B. Duoss, T.H. Weisgraber, K. Hearon, C. Zhu, W. Small, T.R. Metz, J.J. Vericella,
H.D. Barth, J.D. Kuntz, R.S. Maxwell, C.M. Spadaccini, Three-dimensional printing of

als, 24 (31), 49054913 (2014). US


elastomeric, cellular architectures with negative stiffness, Advanced Functional Materi-

[22] D.M. Correa, T. Klatt, S. Cortes, M. Haberman, D. Kovar, C. Seepersad, Negative


AN
stiffness honeycombs for recoverable shock isolation, Rapid Prototyping Journal, 21
(2), 193-200 (2015).

[23] D.M. Correa, C.C. Seepersad, M.R. Haberman, Mechanical design of negative stiffness
M

honeycomb materials, Integrating Materials and Manufacturing Innovation, 4 (1), 1-11


(2015).
ED

[24] T. Frenzel, C. Findeisen, M. Kadic, P. Gumbsch, M. Wegener, Tailored buckling micro-


lattices as reusable lightweight shock absorbers, Advanced Materials, 28(28), 5865-5870
(2016).
PT

[25] C. Ha, R.S. Lakes, M.E. Plesha, Design, fabrication, and analysis of lattice exhibit-
ing energy absorption via snap-through behavior, Materials and Design, 141, 426-437,
(2018).
CE

[26] W.J. Drugan, Wave propagation in elastic and damped structures with stabilized
negative-stiffness components, Journal of the Mechanics and Physics of Solids, 106,
34-45 (2017).
AC

[27] D. Chronopoulos, I. Antoniadis, M. Collet, M. Ichchou, Enhancement of wave damping


within metamaterials having embedded negative stiffness inclusions, Wave Motion, 58,
165-179 (2015).

[28] B.A. Fulcher, D.W. Shahan, M.R. Haberman, C.C. Seepersad, P.S. Wilson, Analytical
and experimental investigation of buckled beams as negative stiffness elements for pas-
sive vibration and shock isolation systems, Journal of Vibration and Acoustics, 136 (3),
031009 (2014).

18
ACCEPTED MANUSCRIPT

[29] C.B. Churchill, D.W. Shahan, S.P. Smith, A.C. Keefe, G.P. McKnight, Dynamically
variable negative stiffness structures, Science advances, 2 (2), e1500778 (2016).

[30] D. A. Debeau, C. C. Seepersad, M. R. Haberman, Impact behavior of negative stiffness


honeycomb materials, Journal of Materials Research, 33 (3), 290-299 (2018).

[31] L.Y. Zhang, G.K. Xu, Negative stiffness behaviors emerging in elastic instabilities of
prismatic tensegrities under torsional loading, International Journal of Mechanical Sci-
ences, 103, 189-198 (2015).

T
[32] H. Kalathur, R.S. Lakes, Enhancement in piezoelectric sensitivity via negative structural

IP
stiffness, Journal of Intelligent Material Systems and Structures, 27 (18), 2568-2573
(2016).

CR
[33] A.G. Evans, M.Y. He, V.S. Deshpande, J.W. Hutchinson, A.J. Jacobsen, W.B. Carter,
Concepts for enhanced energy absorption using hollow micro-lattices, International
Journal of Impact Engineering, 37 (9), 947-959 (2010).

US
[34] L. Salari-Sharif, T.A. Schaedler, L. Valdevit, Energy dissipation mechanisms in hollow
metallic microlattices, Journal of Materials Research, 29 (16), 1755-1770 (2014).
AN
[35] B. Haghpanah, L. SalariSharif, P. Pourrajab, J. Hopkins, L. Valdevit, Multistable
shapereconfigurable architected materials, Advanced Materials, 28(36), 7915-7920
(2016).
M

[36] T. Mullin, S. Deschanel, K. Bertoldi, M.C. Boyce, Pattern transformation triggered by


deformation, Physical review letters, 99(8), 084301 (2007).
ED

[37] K. Bertoldi, P.M. Reis, S. Willshaw, T. Mullin, Negative Poisson’s ratio behavior in-
duced by an elastic instability, Advanced materials, 22(3), 361-366 (2010).

[38] J.L. Silverberg, J.H. Na, A.A. Evans, B. Liu, T.C. Hull, C.D. Santangelo, R.J. Lang,
PT

R.C. Hyward, I. Cohen, Origami structures with a critical transition to bistability arising
from hidden degrees of freedom, Nature materials, 14(4), 389 (2015).

[39] J.D. Renton, Elastic Beams and Frames (Horwood, Chichester, UK, 3d. 2, 2002).
CE

[40] J.M. Williams, A. Adewunmi, R.M. Schek, C.L. Flanagan, P.H. Krebsbach, S. E. Fein-
berg, S. J. Hollister, S. Das, Bone tissue engineering using polycaprolactone scaffolds
AC

fabricated via selective laser sintering. Biomaterials, 26 (23), 4817-4827 (2005).

[41] D.L. Bourell, T.J. Watt, D.K. Leigh, B. Fulcher, Performance limitations in polymer
laser sintering, Physics Procedia, 56, 147-156 (2014).

[42] E.O.T. Olakanmi, R.F. Cochrane, K.M. Dalgarno, A review on selective laser sinter-
ing/melting (SLS/SLM) of aluminium alloy powders: Processing, microstructure, and
properties, Progress in Materials Science, 74, 401-477 (2015).

19

Vous aimerez peut-être aussi