Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

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

net/publication/339301835

Analytical Modeling and Design of Negative Stiffness Honeycombs

Article in Smart Materials and Structures · February 2020


DOI: 10.1088/1361-665X/ab773a

CITATIONS READS
30 1,200

3 authors:

Altay Zhakatayev Zhanat Kappassov


Nazarbayev University Nazarbayev University and Bentley University
25 PUBLICATIONS 212 CITATIONS 30 PUBLICATIONS 895 CITATIONS

SEE PROFILE SEE PROFILE

Atakan Varol
Nazarbayev University
156 PUBLICATIONS 3,607 CITATIONS

SEE PROFILE

All content following this page was uploaded by Zhanat Kappassov on 17 February 2020.

The user has requested enhancement of the downloaded file.


Analytical Modeling and Design of Negative
Stiffness Honeycombs
Altay Zhakatayev, Zhanat Kappassov, Huseyin Atakan Varol
E-mail: azhakatayev@nu.edu.kz, zhkappassov@nu.edu.kz, ahvarol@nu.edu.kz

July 2019

Abstract. Negative stiffness honeycombs (NSHs) have multiple advantages


compared to traditional honeycomb structures. These advantages include recoverable
elastic buckling, shock absorption, and vibration isolation. As a result, NSHs have
great potential in applications such as acoustic wave guiding, shape morphing, design
of impact-protection devices and robotic grippers. In this paper, we present a design
methodology for metamaterial consisting of negative stiffness beams assembled in a
honeycomb structure. Based on analytical results, our methodology allows designing
NSH mesostructures with predetermined buckling sequence. An NSH prototype was
designed based on our algorithm and fabricated using a 3D printer with a nylon
filament. The validity of our approach was experimentally verified by performing
displacement controlled compression force measurements. Our methodology gives the
ability to design NSH mesostructures with desired force-displacement characteristics.

Keywords: negative stiffness honeycomb, beam buckling, design methodology,


predetermined buckling sequence.
Sym. Description S. Description S. Description
p
l -horizontal beam length w(x) -beam shape λ = EI
t -thickness of the beam w0 (x) -initial beam shape N = λl
b -depth of the beam E -Young’s modulus Nj -eigenvalues of (11)
h -height of the middle I -second moment of area eb -bending energy
point of the beam of the beams’ cross section ec -compression energy
s -beam arc length p -axial compressive force ef -work of external force
eb l 3
s0 -beam’s initial s fmax -buckling threshold force Eb = EIh 2
ec l 3
X = xl d -deflection of beam’s center Ec = EIh2
e l3
S = hsl2 W (X) = w(x)
h Ef f
= EIh 2
f l3 1 h2
S0 = sh02l F = EIh α = 1 + 16 2
l 2 N1
2
Q = ht D = hd β = 3Q

ζ(γ) -function of γ fmin -minimum force γ -coefficient
Zc -NSH column number Zr -NSH row number Aj -weight of the modes

1. Introduction

Negative stiffness behavior is described as the decrease in the force when the
displacement increases. Simple case of negative stiffness happens when a beam is
2

subject to a compressive force, which causes the beam to buckle. The buckled beam
has two stable configurations, as shown in figure 1a. If a force perpendicular to the
beam is applied, then the beam can transition from one stable configuration into
another. The maximum force supported by the beam before it snaps into alternative
stable position is called the force threshold. A buckled beam is also called a bistable
mechanism, which has two stable equilibrium configurations. The transformation from
one equilibrium configuration into another requires external input or energy. On the
other hand, a metastable mechanism has only one stable equilibrium and one quasi-
stable equilibrium. Therefore, it transforms from stable into metastable configuration
under external input, and if the input is removed, metastable mechanism returns to its
original stable configuration.

Figure 1: a) Buckled simple beam in the first stable configuration (solid line) and in
the second stable configuration (dashed line), b) negative stiffness honeycomb with two
columns and rows, and c) regular hexagonal honeycomb.

Multiple meta- and/or bistable negative stiffness beams can be grouped in series
and in parallel to form negative stiffness honeycomb (NSH) structures. NSHs, illustrated
in figure 1b, have several advantages compared to traditional materials. Firstly, they
exhibit relatively long plateau in the force-displacement graph, which makes them
suitable for energy absorption and shock isolation [1]. Secondly, their deformation is
fully recoverable, i.e. their energy absorption and shock isolation capabilities are not
restricted to a single use [1]. Traditional honeycomb structures (figure 1c) also have
plateau region in the stress-strain diagram. However, NSHs deform in an elastic mode,
while traditional honeycomb structures deform in a plastic mode. Therefore, traditional
honeycombs can be utilized only once for shock isolation. Afterward, they need to
be replaced, which is not convenient and economically feasible. On the other hand,
NSHs can be used multiple times. Therefore, NSHs are useful for cases when repeated
loading and unloading are present. Traditional honeycombs can also be designed
to have recoverable elastic deformation, but it requires low overall material density.
As a result, stiffness and strength per unit volume of elastic traditional honeycombs
are inconveniently low for real-world applications. Thirdly, NSHs exhibit high initial
stiffness at low material density [2]. Contrarily, high initial stiffness in conventional
honeycombs can only be obtained with relatively high material density, which causes
unrecoverable plastic deformations. The final advantage of NSHs is their ability to
change mechanical properties (e.g. stiffness) by varying the deformation level [3]. This
3

is useful for material property modulation, which facilitates the design of materials with
large controllable variations in macroscopic response. Potential applications of NSH
structures include acoustic wave guiding, filtering, shape morphing, design of impact-
protection devices, controllable switches, robot grippers, and tactile sensors [4, 5].
Now, we will briefly review the literature on NSHs. Experimental investigation
of force-displacement function of NSHs was performed in [2]. Through finite element
analysis (FEA), it was demonstrated that NSHs can approach the energy-absorbing
capacities of the traditional hexagonal honeycombs. Design of NSHs was considered
in [6]. FEA was performed to assess the trade-offs between stress threshold, relative
density and absorption energy of NSHs. Specifically, it was shown that for a fixed
stress threshold, absorption energy per unit cell decreases when the relative density of
the honeycomb is increased. This effect is more pronounced for higher values of the
stress threshold. In traditional hexagonal honeycombs absorption energy increases with
relative density.
Impact behavior of NSH structures was studied in [1]. It was shown that NSH
can absorb energy during impacts at a constant and low force threshold. This is due
to the temporal extension of impact duration and the reduction of peak acceleration
during the impact. By performing FEA and experiments with aluminum and nylon
NSHs, it was demonstrated that the force threshold is proportional to the number of
negative stiffness cell columns, while the energy absorption capability is proportional to
the number of negative stiffness cell rows in the NSH structure. Utilization of negative
stiffness elements, in parallel with linear positive stiffness springs, as vibration and
shock isolators with quasizero stiffness was investigated in [7]. Vibration transmissibility
of negative stiffness beams and their fabrication with selective laser sintering were
studied in [8]. Metastable negative stiffness beams were used to design microlattices
as light-weight shock absorbers [9]. Utilization of bistable negative stiffness elements
to design and manufacture morphing structures was proposed in [10]. In that paper,
the authors investigated the influence of beam parameters (thickness, height, and span)
on its bistable behavior. Buckling, as well as folding, is found to be the most efficient
method of deformation with energy absorption, and the corresponding NSH was applied
to design energy-absorbing bumpers for cars [11].
Metamaterials are man-made materials engineered to achieve unique physical
properties. NSHs have great potential as metamaterials [12]. Cellular materials with
unit cells consisting of compliant bistable or metastable mechanisms were considered
from the perspective of phase transformations in [4]. The authors defined the phase
transformation in a cellular material as the change in geometry of its unit cell while it
transforms from one stable configuration into another with the same topology. Force-
displacement relationship of the negative stiffness structure was approximated by a
piece-wise linear function with three segments. Researchers derived the equations to
find coefficients of these linear functions as a function of system parameters. It was
noted that the metastable behavior of NSHs resembles the pseudo-elasticity exhibited
by superelastic shape memory alloys [4]. Similarly, NSHs were analyzed as tunable
4

elastic metamaterials with variable acoustic wave propagation behavior [3]. FEA was
utilized to model the static deformation and the subsequent linear wave motion at the
predeformed state. It was shown that NSHs exhibit significant tunability of mechanical
properties and a high degree of anisotropy. In other words, NSH structures might be
efficient for applications requiring mechanical impedance-matching with the surrounding
environment. Multidirectional metamaterials based on negative stiffness beams were
designed in [13]. Specifically, force-displacement characteristics of the single negative
stiffness beam as a function of beam thickness, length, and apex height were studied
using theory and simulations. Authors designed uni-, bi- and tri-directional NSHs with
identical beam parameters and investigated their static force-displacement and dynamic
transient properties due to multiaxial loading. However, in this study, the influence
of beam parameters on the buckling sequence was not explored. Negative stiffness
metamaterials under tension, which can be tuned by harnessing the snap-through
instabilities, were investigated in [14]. Dynamic analysis of phase transforming cellular
structures based on negative stiffness beams was investigated in [15]. Metamaterials
with negative stiffness and Poisson’s ratio were implemented in [16]. The metamaterial
was designed using regular array of interlocked rigid hexagonal subunits, which could
slide relative to each other only along certain predetermined directions. This feature
endows the metamaterial with exotic behavior, such as negative Poisson’s ratio. Buckled
beams, magnets and polymethacrylimide foams were considered as connection elements
between hexagonal subunits. However, the paper mostly describes the relation between
geometry of hexagonal subunits and the metamaterial properties, while the design of
buckling beams is not discussed. Similarly, metamaterial with negative Poisson’s ratio,
called auxetic material, was designed in [17].
Static force-displacement, equilibrium, and stability analysis of discrete chains of bi-
stable elements were performed in [18]. Nonlinear dynamic analysis of one-dimensional
chains of bistable elements demonstrated rich dynamic response with three different
acoustic wave propagation regimes depending on wave amplitude [19]. It was shown
that acoustic wave propagation characteristics of the chain can be controlled by fine-
tuning the geometry of the bistable elements. Tilted elastic beams were shown to possess
bistable and metastable behavior [20], and energy-trapping materials were developed
based on locking-in of strain energy in them. Bi-stable hinge mechanism was used to
design several types of multistable 2D and 3D structures [21]. Authors demonstrated
that reducing unit cell size in these materials increases their strength and elasticity.
Different values of the external force, at which a single row of the NSH collapses,
were attributed to small imperfections in the manufactured NSHs [1, 4, 6, 20, 21]
and to non-ideal boundary conditions [22, 23]. These imperfections trigger the rows
at slightly different external force values. A design methodology for honeycombs
consisting of bistable beams with a predetermined deformation sequence was considered
in [23]. However, variations of only two factors were investigated analytically and
experimentally. Firstly, the thickness of the curved beam was varied to obtain
the deterministic deformation sequence. Secondly, the effect of mode imperfections
5

proportional to the third buckling mode was studied. Contribution of deformation of a


single bistable cell to the overall honeycomb deformation was obtained by formulating
and solving an optimization problem. FEA simulations matched the experimental
results closely, while theoretical analysis correctly described the deformation sequence.
Despite the progress in NSH research, authors are not aware of a formal method to design
NSHs with a predetermined snapping force values taking into account all geometric and
material properties.
In traditional honeycombs, compressive strength, and energy absorption capabilities
are influenced by cell shape and density [6]. Similarly, compressive strength and energy
absorption properties of NSHs depend on unit cell geometry, material, and honeycomb
structure. Therefore, selecting optimal values for dimensional parameters of NSH cells
is crucial for their properties and performance. Based on an earlier analysis of force-
displacement characteristics of a double beam in [24], we present a design methodology
for NSHs with predetermined buckling sequence. We expand their work by deriving
explicit and implicit equations for beam’s dimensional parameters under a certain class
of force-displacement behaviors. The main contribution of our work is to offer a design
methodology, which can be utilized to construct NSHs with desired force threshold
and force-displacement properties. As a result, it becomes possible to synthesize NSHs
with unique and predetermined mechanical properties. We verify our method through
experiments with 3D-printed NSHs.
The remainder of the paper is structured as follows: Analysis of the negative
stiffness beam and the design algorithm for NSH are given in section 2. Experimental
procedure is described in section 3, which is followed by presentation of simulation and
experimental results in section 4. The latter also contains discussion of the obtained
results. Finally, the paper is concluded in section 5.

2. Analysis of the Negative Stiffness Beam

2.1. Theoretical Analysis of the Buckled Beam


Mathematical analysis of negative stiffness beams was performed in [24]. It was shown
that prefabricated curved beams behave similarly to the straight buckled beams under
axially compressive force [22, 24]. In order to be concise, the nonlinear force response
for a single beam is derived analytically in appendix 7. In the derivation, we start
with the description of a buckling beam, figure 2a. Then, we present equations of
three types of work when the beam deforms, introduce normalized variables, represent
the three equations of work using these, and derive buckling modes based on energy
conservation. Finally, we obtain a force-displacement equation using normalized and
physical variables.
The second mode buckling lowers initial stiffness and threshold force of the
honeycomb, and also it prevents the beam from exhibiting negative stiffness
behavior [22]. In order to suppress the second mode buckling, one customary solution
6
f
y a) b)
s0

t w0(x) h b
x
l

Figure 2: Geometry of: a) the single negative stiffness beam and b) double beam
connected at the center.

is to connect two beams in the middle [3,6,13], as shown in figure 2b. For simplicity, we
now assume that only the first buckling mode is present, i.e. Aj = 0, j 6= 1, where Aj are
the weight coefficients of the corresponding buckling modes. We encourage the readers
to go over the derivations in appendix 7, where most of the variables are introduced.
Additionally, the symbols and their meanings are summarized in the table on the first
page. If equations (26), (27) and (32) are utilized from appendix 7, then the following
force-displacement equation is obtained

N14 1+β
F (D) = β(D3 − 3D2 + 2 D), (1)
16 β
2
where β = 3Q2α
. Equation (1) is cubic and F (0) = 0. Extrema of the equation (1) can
dF
be found from dD = 0 as
s
2 1+β
D1,2 = 1 ∓ 1 − ( ) (2)
3 β
For two roots in equation (2) to be real, the condition β > 2 should be satisfied. It can
d2 F d2 F
be shown that dD 2 (D1 ) > 0 and dD 2 (D2 ) < 0 and so F (D1 ) and F (D2 ) correspond to

local maximum and minimum, respectively, see figure 3. Another condition that we can
impose on equation (1) is the requirement that the local maximum force is equal to some
fraction of the force when the beam is fully compressed (buckled), i.e. F (D1 ) = γF (2),
where γ is coefficient. If we choose F (D1 ) > F (2) (γ > 1), then during the force-
controlled deformation, the honeycomb will continue buckling until the force reaches
the pre-buckling level F (D1 ). However, this will happen at D > 2 and the symmetry of
the system will be broken. Alternatively, if there is a hard stop at D = 2, then the force
will be restored to the value F (D1 ) at D = 2. If, on the other hand, F (D1 ) < F (2) (i.e.
γ < 1) is chosen, then during the force-controlled deformation the beam will stop before
it reaches the symmetric position at D = 2. Contrarily, if the condition F (D1 ) = F (2)
(γ = 1) is imposed, then during the force controlled deformation the beam will buckle
exactly until it reaches D = 2. From the condition F (D1 ) = γF (2), we arrive at the
7

following relation s
2 β−2 2 1+β
=1+ 1− ( ), (3)
γ 3 3 β
which needs to be satisfied. By performing some algebraic manipulations on equation
(3), we arrive at the following cubic polynomial

3 2
 2 2

β − 6β + 12 − 27( − 1) β − 8 = 0. (4)
γ
From this polynomial we can find the necessary values of β. Out of the three roots
of the polynomial (4), only one is real, while the other two are in general complex
numbers. Since all variables utilized for the NSH are real, the real root is selected.
p
3

Using ζ(γ) = (γ − 2)2 (γ + 2 γ − 1), the real root can be written as

3ζ(γ) 3(γ − 2)2


β= + + 2. (5)
γ γζ(γ)
Now, we will switch back to the physical variables using equation (18). The
maximum fmax and minimum fmin forces, obtained at d = D1 h and d = D2 h,
respectively, can be written as
s
EIh N14  β−2 2 1+β 
fmax,min = 3 1± 1− ( ) . (6)
l 8 3 3 β

These two forces are selected as design variables, and they correspond to non-
dimensional forces F (D1 ) and F (D2 ), figure 3. By taking the ratio of fmax to fmin ,
the following expression is obtained
fmin
γ =1+ . (7)
fmax

700

600

500

400 F(D1) or fmax


300

200

100

-100
F(D2) or fmin
-200

-300
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 3: Nondimensional force F as a function of displacement D for three cases:


metastable, bistable and critically stable.
8
3Q2
From the definition of β = 2α
, we can find the relationship between t and h as
s
3
t=h 1 h2
. (8)
2β(1 + 16 l2
N12 )

If equation (8) is substituted into fmax expression in (6), then we obtain the following
nonlinear equation for h
s
2 3 4
 2β 1 h 2 2 EbN1 4
  β−2 2 1+β 
8fmax (1 + N ) − h 1+ 1− ( ) = 0. (9)
3 16 l2 1 12l3 3 3 β

As mentioned above, for negative stiffness behavior of the beam, the condition
β > 2 should be satisfied. Utilizing equation (3), γ can be plotted as a function of β
(figure 4a). The inverse function can be plotted based on equation (5). As it can be
observed, the range of 2 < β < ∞ corresponds to the range 2 > γ > 0, (e.i. when β = 2,
then γ = 2). Dependence of F (D1 ) and F (D2 ) on γ is shown in figure 4b, where it can
be concluded that these forces are uniquely determined by γ (single valued functions).
Therefore either of γ or a pair of F (D1 ) and F (D2 ) can be used as design variables. The
following important observations can be made from the figure 4b. F (D2 ) > 0 within
the range 1 < γ < 2, which means that the beam is metastable (if the external force
is removed after buckling, the beam returns to its initial position). While F (D2 ) < 0
within the range 0 < γ < 1, which means that the beam is bistable (if the external force
is removed after buckling, the beam stays in its second equilibrium state). The point
γ = 1 is boundary point which separates metastable from bistable region, which can be
called as critically stable configuration, and in this special case the following equations
are valid
N14 3 9 1 3 N4
F = (D −3D2 + D), D1 = , D2 = , F (D1 ) = F (2) = 1 , F (D2 ) = 0. (10)
2 4 2 2 4

2
1000

1.5
500

1
0

0.5 -500

a) b)
0 -1000
0 50 100 150 200 0 0.5 1 1.5 2

Figure 4: a) The relation between γ and β. b) The relation between F and γ.


9

2.2. Design of NSHs with Predetermined Buckling Sequence and Forces


There are four-dimensional parameters of buckling honeycombs: l, b, h, and t. How do
we choose these parameters for a NSH with predetermined buckling sequence? Let us
assume that the forces at which NSHs should buckle are given as a sequence (vector)
of forces fmax . Additionally, length l and depth b are external dimensions of the beam,
which affect the overall dimensions of the NSH. In other words, if the length and depth
of beams are varied, then it would be more challenging to obtain the NSH within a
predetermined volumetric envelope. In general, the overall size of the NSH is fixed or
known in advance. Therefore, it is more convenient to fix the length and depth of the
beam and instead, vary its thickness t and profile height h. We can devise the following
procedure to design NSHs with pre-determined buckling sequence (algorithm 1).

Algorithm 1 Design of NSH with a predetermined buckling sequence


input: NSH material with Young’s modulus E, number of columns Zc , number of
rows Zr , set of values for length of a single beam li , depth of a single beam bi , maximum
force at which NSH should buckle fmax,i , and minimum force which is achieved in the
beam’s force-displacement characteristics fmin,i : 0 < |fmin,i | < |fmax,i |.
for all i ∈ [1, 2, ..., Zr ] do
γi ← Equation (7) given fmax,i and fmin,i ;
βi ← Equation (5) given γi ;
hi ← Equation (9) using a nonlinear root finder given li , bi , βi and fmax,i ;
ti ← Equation (8) using li , hi and βi .
end for
return Values for beam thickness and apex height of the middle point

Please note that sub-index i in the parameters mentioned in the algorithm denotes
only the fact that each row of the NSH can have beams with different physical
dimensions. However, all equations prior (9) do not contain the sub-index i. The
aforementioned algorithm was devised with the assumption that beam parameters are
the same for a given row, while they differ from row to row. Another point is that the
net force threshold of the NSH is Zc fm , while total energy absorbed is correlated to Zr .
The final point is that the condition 0 < γ < 2 translates to the following condition in
terms of the minimum force 0 < |fmin,i | < |fmax,i |.
The proposed NSH design algorithm has the following advantages compared to
the existing alternatives. Firstly, unlike the design methodologies presented in [4, 24],
our method does not use a piecewise linear approximation of force-displacement
characteristics. In other words, the full cubic representation of the F (D) is utilized.
The second advantage is that the proposed methodology allows designing all NSH
types: bistable, metastable and critically stable (just at the boundary of bistable and
metastable) NSHs. Therefore, a designer is free to design NSHs with predetermined
buckling sequence. Additionally, NSHs with interchanging rows of bistable, metastable
and critically stables behaviors can be designed.
10

Some conditions for force-displacement characteristics of the NSHs are unattain-


able. For example, we can try to apply the condition that the local maximum force is
equal to the negative value of the local minimum force, i.e. F (D1 ) = −F (D2 ). This
would ideally result in NSH structure with symmetric bistable behavior of beams. If
this condition is applied to (1), then the only case when it is satisfied corresponds to
β = 2. However, this value of β violates the requirement (β > 2) for the existence of two
extrema of F (D). As a result, F (D) curve will have only a single extremum which also
means that the beams will not possess negative stiffness behavior. Therefore, this case
can not be implemented. Yet another case, where F (0) = F (2) = 0, can be considered.
However, again this condition, when applied to (1), can not be satisfied for any values
of β, and is thus impossible to implement.
After designing the NSH with specific parameters (li , bi , hi and ti , i = 1, .., Zr ), let
us describe the procedure how to obtain its overall force-displacement curve. Force-
displacement characteristic of each beam in the NSH structure is obtained from
equation (10). Then, the overall system behavior depends on whether force- or
displacement-controlled deformation is applied for the NSH [10]. If the force-controlled
deformation takes place, then the NSH gradually deforms until the first (lowest) force
threshold is achieved for one layer of the NSH. At that moment the specific layer
suddenly buckles and the buckled layer can be considered as non-present for further
process. As the force is increased further, the remaining rows in the NSH will continue
gradual deformation until the force threshold of the second layer is achieved. At that
moment the second row collapses and the process continues until all of the remaining
rows of the NSH buckle. If all rows are compressed, then NSH behaves as a rigid
material. The situation is different for displacement controlled deformation, during
which the position of the NSH’s position is set. In other words, in this case, a necessary
external force (for example, due to a robot manipulator) is required to satisfy equilibrium
condition at any compressed position of the NSH. During the position controller
displacement, NSH can demonstrate troughs in the force-displacement characteristics.

2.3. Analysis of NSHs with the Given Geometry


Let us remind that given the desired buckling sequence or overall force-displacement
relation, the algorithm 1 allows us to find the necessary geometry of the beams. Using
the aforementioned analysis, we can deduce the following procedure to analyze the
force-displacement behavior of NSH with fixed (given) geometry. In other words, next
we present the algorithm, which would allow us to find the overall force-displacement
relation of the NSH given the geometry of the beams (algorithm 2). This situation
might arise when, for example, the beam is already manufactured, and one needs to
find its buckling sequence or overall force-displacement relation.
11

Algorithm 2 Analysis of NSH with a predetermined geometry


input: NSH material with Young’s modulus E, number of columns Zc , number of
rows Zr , set of values for length li , depth bi , thickness ti and height hi of a single
beam.
for all i ∈ [1, 2, ..., Zr ] do
1 h2
αi ← Equation α = 1 + 16 l2
N12 ;
2
βi ← Equation β = 3Q 2α
;
γi ← Equation (7) given βi ;
fmax,i ← Equation (6) given βi ;
fmin,i ← Equation (6) or (7) given βi or fmax,i and γi ;
end for
return Estimate of the nonlinear reactive force as the function of displacements

3. Case Study

In order to validate our analytical model and the corresponding NSH design
methodology, we fabricated one NSH prototype and characterized it experimentally.
We used our algorithm 1 to design this prototype. First, we set seven input parameters
of the algorithm. These parameters are Young’s modulus E, number of rows Zr ,
number of columns Zc , sequence of beams’ length l, depth b, buckling force fmax ,
and minimum force fmin . Design of NSH with two columns Zc = 2 was selected to
prevent its undesirable bending and rotation. Our goal was to achieve four buckling
thresholds, which results in four rows Zr = 4, that appear at fmax = [4 7 9 17] N
force amplitudes for a corresponding single beam. For a whole NSH structure the
overall buckling force thresholds would be four times of these fmax values (two columns
double the value and double beams for each column again make it double). These
force values were chosen arbitrarily, as well as fmin = [−3 −5 −5 −10] N values. We
intended the buckling sequence to be such that the beams buckle from the top to the
bottom row in sequential order. Even though E = 300 MPa is given by the material
used for prototyping (nylon 12 powder), its real value can vary in the interval between
300 MPa until 2.4 GPa depending on the implementation of NSH using 3D printing
technology [25]. For example, the exact value of Young’s modulus depends on the mass
density of the material, which is in turn affected by sintering degree. Lengths of all
beams were fixed at l = 60 mm, while their depths were chosen as b1 = b2 = 5 mm and
b3 = b4 = 7.5 mm. By feeding these parameters to the algorithm 1, we obtained the
following values for h = [5.0 5.6 5.0 6.0] mm and t = [1.0 1.1 1.3 1.5] mm. The
values of all parameters are summarized in Table 1.
We used double beams connected at the center (figure 5a) to prevent buckling
through the second mode as mentioned in section 2. Every double beam set was then
integrated with a T-shaped indenter and rigid supporting walls. The walls should
be stiff enough to withstand the mechanical stress caused by deformations of the
12

Force
gauge
sensor
Wedge
Honeycomb
grip
Platforms

Wedge
grip

(a) (b)

Figure 5: (a) Dimensions of NSH mesostructure. (b) Overall view of the testing setup
with NSH structure.

beam. Otherwise, the walls bend and the buckling effect diminishes since there is no
compression energy variation (equation 14). Indeed, the axial compressive force p (see
section 2) is the reaction force of the supporting walls. The double beam with rigid walls
on both sides acts as a unit cell of NSH structure. These unit cells were attached to
each other both in series and in parallel. Since we are interested in vertical force versus
vertical displacement characteristic of the NSH, beam parameters are the same within
a single row. All geometrical parameters (bi , hi , and ti , i ∈ [1, 2, 3, 4]) are depicted in
figure 5a. The NSH, with parameters from Table 1, was designed in SolidWorks software
and the final sample was fabricated with a selective laser sintering (SLS) process based
on a 3D printer using Nylon 12 powder.
Experiments were conducted in order to obtain the force-displacement characteristic
of the 3D-printed NSH sample using a universal testing machine (Tinius Olsen H25KS),
with maximum force capacity of load cell 25 kN and maximum displacement rate of
1000 mm/min. Quasi-static compression and extension tests were performed in order
to obtain the force-displacement relation for NSH. We set the compression speeds to
60 and 300 mm/min. Triangular side walls made from steel were rigidly attached to
the bottom platform. Their role was to prevent the asymmetric and out-of-the-plane
bending of the NSH, and to enforce a zero-displacement boundary condition at the edges
of the beams. The upper part of the NSH was fixed to the 3D-printed top platform made
out of PLA plastic, which in turn was grasped firmly by the upper wedge grips of the
universal testing machine (figure 5b). The bottom part of the NSH was fixed to the
lower platform (PLA plastic), which was attached to the bottom wedge grips. The
universal testing machine squeezed the sample while its force sensor was simultaneously
13

recording the reaction force at 200 Hz.


We also performed numerical simulations of the vertical compression of the designed
NSH using finite element analysis (FEA) implemented in ANSYS 18.1 Mechanical
Software. In FEA, the bottom part of the NSH was fixed, while the side walls were
constrained from the horizontal motion to prevent lateral expansion of the structure.
An elastic and isotropic constitutive model was utilized in simulations, with Young’s
modulus, Poisson’s ratio and density of the material equal to E = 600 MPa, ν = 0.3 and
ρ = 1100 kg/m3 , respectively. In quasi-static FEA analysis, the vertical displacement
was imposed in the range 0-43.2 mm. Three different mesh sizes (1 mm, 0.75 mm and
0.5 mm) were used to investigate the convergence of FEA simulations. The model was
meshed using tetrahedron elements, patch conforming algorithm and global settings
for element order. Simulations were conducted on HP Z640 Workstation with the
processor Intel(R) Xeon(R) CPU E5-2620. The analytical model, FEA simulations
and experimental results for the given NSH are compared in the next section.

Table 1: Buckling beam parameters for NSH prototype.

Parameter Value Parameter Value


Zr , Zc 4, 2
l 60 mm h [5.0 5.6 5.0 6.0] mm
Output
Input

b [5 5 7.5 7.5] mm t [1.0 1.1 1.3 1.5] mm


fmax [4 7 9 17] N
fmin [−3 −5 −5 −10] N

4. Results

4.1. Simulation Results


One set of simulations, based on equations (8) and (9), was performed in order to
investigate the influence of parameters l and b on h and t. For these simulations, l was
varied from 4 mm until 10 cm with a step size of 4 mm, while b was varied from 2 mm
until 5 cm with a step size of 2 mm. Figures 6a-b depicts dependence of parameters h
and t on l and b for a fixed values of E = 1.2 GPa, γ = 1, β = 8 and fmax = 5 N. Results
show that as l is increased and/or b is decreased, then both h and t increase. The reason
for this behavior might be the following: for the same material, as the beam length is
increased and/or its depth is decreased, then the force threshold decreases (it becomes
easier to snap the beam). Therefore in order to keep the value of the force threshold,
both beam thickness and apex height of its middle point should be increased. Another
set of simulations was obtained by varying E from 100 MPa until 2 GPa with a step size
of 100 MPa and β from 2 until 20 with a step size of 1, figures 6c-d. In this case, both
l = 60 mm, b = 5 mm and fmax = 5 N were fixed. Results demonstrate h increases if
stiffness is decreased and/or β is increased, but t increases if both E and β are decreased.
Finally, simulations were performed for variations of fmax from 1 N until 30 N with step
14

of 1 N and of γ from 0.1 until 1.9 with step size 0.1, figures 6e-f. Results indicate that for
a fixed E, l and b, decreasing γ increases h and decreases t. This can be interpreted as
increase in Q = ht leads to more pronounced buckling behavior. While increasing fmax
causes increase of h and t, which can be explained as increasing middle apex height
and beam thickness causes increase of the force threshold. The results of convergence

c) e)
a)

f)
b) d)

Figure 6: Influence of l and b on: a) h and b) t, and of E and β on: c) h and d) t, and
of fmax and γ on: e) h and f) t.

simulations with three different mesh sizes were the following: the maximum relative
difference for buckling forces obtained from simulations with different mesh sizes was
3%. Therefore, it can be concluded that our FEA simulations converged.

4.2. Experimental results


Analytical model, experimental results and FEA simulations of the force-displacement
behavior of the NSH sample are illustrated in Figure 7. Specifically, the results of
three experiments were utilized to obtain the mean (black solid line for 300 mm/min
and gray solid line for 60 mm/min) of the force-displacement relation for the overall
NSH structure. Experiments at two different compression speeds yield close results.
We remind a reader that for a single beam, the force values are four times less.
The nonlinear force response of the structure is evident as the displacement increases.
There are four negative stiffness regions that start at approximately 11 N, 14 N, 40 N
and 89 N. For a single beam, these forces correspond to values of 2.75 N, 3.5 N,
10 N and 22.25 N, the third of which is close in value to the analytically predicted
buckling force value. These bucklings occur at displacements of around 5 mm, 14 mm,
26 mm, and 36 mm, respectively. The result of FEA simulation (red dashed line),
in general, follows the overall buckling trend of the NSH sample. The predicted non-
15

100
300 mm/min
80 60 mm/min
Theory
60 FEA
Force (N)
40

20

-20

-40
0 10 20 30 40
Displacement (mm)

Figure 7: Analytical solution, numerical simulation and experimental results for force-
displacement relationship of the NSH sample.

(a) (b) (c) (d)

Figure 8: Buckling sequence of the NSH sample. From (a) to (d) are states from the
first buckling to the the last one.

linear force response derived using the proposed design algorithm 1 is also shown (green
dash-dotted line). It can be observed that the overall behavior of the theoretical
force-displacement curve follows the experimental behavior, even though the proposed
algorithm predicts significant negative forces, which are not observed in the experiments.
The discrepancies between the theoretical model and experiments (real-world and FEA)
might be attributed to several sources. One possible source is the simple model of
the material (homogeneous, isotropic and linear elastic) used in the analysis, while
the actual material might be close in behavior to plastic materials. Another potential
source is the simple buckling model used in the analysis, where only the first buckling
mode is considered, while others are ignored. Additional source of discrepancy might
be attributed to imperfections within the material (non-uniform material distribution
throughout its volume due to 3D printing) and to nonlinear properties of the material
(hysteresis). The final source is possibly related to varying geometry (non-constant
thickness and depth) of the beams due to imperfect 3D printing, and to the presence of
significant frictional forces, which will affect the force measurements. Buckling sequence
16

of the NSH mesostructure is depicted in figure 8. From the top to the bottom are rows
from one to four. The beams buckled in the sequential order from top to bottom, which
matches the analytically predicted buckling sequence. As a result, we want to stress that
our proposed design methodology can be used for qualitative design of NSH structures
for a given buckling sequence. Real-world experiments, coupled with iterative design
process, and more accurate model are required for quantitative design of NSH with
actual buckling forces at specific values. The video of the experiment is provided in the
supplemental multimedia material.

5. Conclusion

Theoretical behavior of the negative stiffness beam was investigated based on the
energy method. Specifically, an analytical solution for a metamaterial, which consists of
negative stiffness beams assembled in a honeycomb structure, was derived. As a result,
design methodology, which would allow the design of the NSH with predetermined
buckling sequence, was developed. Additionally, analysis methodology, which allows
investigating the force-displacement behavior of the NSH with the given geometry,
was also presented. In order to verify the theoretical analysis, FEA simulations were
performed on the 3D model of the NSH with four rows and two columns. Moreover,
NSH sample was manufactured with 3D printing technologies out of nylon powder.
The subsequent real-world experiments, under displacement controlled compression,
were performed to determine its force-displacement characteristics. The experimentally
observed and designed buckling sequences are similar. Advantages of the proposed
methodologies include the ability to design NSH structures with predetermined force
thresholds and buckling sequence, and clear understanding of different trade-offs in the
design of NSHs. The proposed methodology gives engineers and scientists a tool to
design and analyze NSH mesostructures. As future work, we plan to refine the model
by including shear and bulk modulus of the material, and the higher-order buckling
modes.

6. Acknowledgments

This work was partially supported by Nazarbayev University Faculty-development


competitive research grants program “Variable Stiffness Tactile Sensor for Robot
Manipulation and Object Exploration” 110119FD45119 and the Ministry of Education
and Science of the Republic of Kazakhstan grant “Methods for Safe Human Robot
Interaction with Variable Impedance Actuated Robots” AP05135733.

7. Appendix I

In the following, we derive analytically the nonlinear force response for a single beam.
Let us consider a beam with horizontal length l, thickness t, depth b, the second moment
17

of area of the cross section I and Young’s modulus E as shown in figure 2a. There is axial
compressive force p acting on the beam. Homogeneous differential beam equation [26]
can be written as
w0000 (x) + λ2 w00 (x) = 0, (11)
where w(x) is the displacement of the beam at the coordinate x from its central line
p
(i.e. shape of the deformed beam), λ2 = EI , while apostrophe denotes differentiation
0 dy(x)
operator y (x) = dx . Solution of (11) with the clamped-clamped boundary conditions
w(0) = w(l) = w0 (0) = w0 (l) = 0 results in the following eigenvalues and eigenvectors
written in compact form using the variable N = λl:
• for sin( N2 ) = 0 eigenvalues are Nj = (j + 1)π, j = 1, 3, 5, .., while eigenvectors are
wj (x) = C(1 − cos(Nj xl )).
• for tan( N2 ) = N2 eigenvalues are Nj = 1.43jπ, 1.23jπ, 1.16jπ, ..., j = 2, 4, 6, .., while
eigenvectors are wj (x) = C(1 − cos(Nj xl ) − 2 xl + N2j sin(Nj xl )).
where C is an arbitrary constant. We assume that the beams have as initial shape w0 (x)
which is the eigenvector of the first (j = 1) buckling mode w0 (x) = h2 (1 − cos(2π xl )). h is
the apex height of the middle point of the undeformed beam as illustrated in figure 2a.
Initial arc length of the beam s0 under the assumption of small deflection (curvature)
can be approximated as
ˆ lp ˆ l
1
s0 = 0 2
(1 + (w0 (x)) )dx ≈ (1 + (w00 (x))2 )dx. (12)
0 0 2
Arc length s of the deformed beam can be evaluated similarly, but w(x) should be used
instead of w0 (x). Let us define variations of three energy terms for the beam. The first
is the variation of the bending energy eb defined as
 EI ˆ l 
∂(eb ) = ∂ (w000 (x) − w00 (x))2 dx . (13)
2 0
The second is the variation of the compression energy ec

∂(ec ) = −p∂(s), (14)

and the third is the work done by an external force f

∂(ef ) = −f ∂(d), (15)

where d is the deflection of the center of the beam


l l
d = w0 −w , (16)
2 2
due to the externally applied force f assumed to be applied at the beam center. Axial
force p can be found through Hooke’s law as
 s
p = Ebt 1 − . (17)
s0
18

It is more convenient to perform further derivations using normalized variables.


Therefore, the following normalized parameters are introduced
x w(x) d sl s0 l h
X= , W (X) = , D = , S = 2 , S0 = 2 , Q =
l h h h h t (18)
f l3 eb l 3 ec l 3 ef l 3
F = , Eb = , Ec = , Ef = .
EIh EIh2 EIh2 EIh2
Odd normalized buckling modes, including their first and second order differentials can
be written as 
Nj = (j + 1)π 


Wj (X) = 1 − cos(Nj X) 

j = 1, 3, 5, ..., (19)
Wj∗ (X) = Nj sin(Nj X) 

Wj∗∗ (X) = Nj2 cos(Nj X)

while their even counterparts are



Nj = [1.43, 1.23, 1.16, ...]jπ 


2


Wj (X) = 1 − cos(Nj X) − 2X + sin(Nj X)


Nj j = 2, 4, 6, ..., (20)
Wj∗ (X) = Nj sin(Nj X) − 2 + 2 cos(Nj X)





Wj∗∗ (X) = Nj2 cos(Nj X) − 2Nj sin(Nj X)

where star denotes differentiation operator with respect to the dimensionless variable X:
y ∗ (X) = dy(X)
dX
. Square brackets for Nj in equation (20) indicate that a specific term of
the sequence should be selected (e.g. for j = 2: Nj = 1.43jπ or for j = 6: Nj = 1.16jπ).
The buckling modes in equations (19)-(20) satisfy the following identities
ˆ 1
1
(Wj∗ (X))2 dX = Nj2 , (21)
0 2
ˆ 1
1
(Wj∗∗ (X))2 dX = Nj4 . (22)
0 2
All buckling modes Wj (X) are orthogonal to each other, similarly their differentials
Wj∗ (X) and Wj∗∗ (X) are, correspondingly, orthogonal to each other [27–29]. Normalized
initial shape of the beam is denoted as W0 (X) = 12 W1 (X). Arbitrary beam shape can
be represented through superposition of the buckling modes as

X
W (X) = Aj Wj (X), (23)
j=1

where Aj are the weight coefficients of the corresponding buckling modes. Using
orthogonality of buckling modes, equations (12), (18), (21) and the identity w0 (x) =
h
l
W ∗ (X), the following normalized parameters can be obtained
ˆ ∞ ∞
l2 1  1 h2 X ∗ 2
 l2 1X 2 2
S= 2 1+ 2( Aj Wj (X)) dX = 2 + AN , (24)
h 0 2 l j=1 h 4 j=1 j j
19
ˆ 1
l2  1 h2 ∗ 2
 l2 N12
S0 = 2 1+ (W 0 (X)) dX = + . (25)
h 0 2 l2 h2 16

Figure 9: The first ten buckling modes of the beam.

Our derivations are based on [24] in which there is a typo in equation (22) and the
consequent derivations contain this error. Therefore, we present here the derivations
without the error. At X = 0.5, all even modes and half of the odd buckling modes are
equal to zero (see figure 9). Using equations (16) and (18) we obtain
1 1 ∞
X
D = W0 −W =1−2 Aj . (26)
2 2 j=1,5,9,...

Please note that when a beam is in its initial prefabricated buckled shape (solid line
in figure 1a) d = 0 (D = 0). When the beam snaps to its alternative equilibrium
buckled state (dashed line in figure 1a) d = 2h (D = 2). Another useful expression can
be obtained by using equations (17), (18), (24), (25) and the definition of the second
3
moment of area I = bt12

N2 pl2 1 N2 1 X
1 2 2 1

= = − A N , (27)
12Q2 EI 12Q2 16 4 j=1 j j α

1 h 2
where α = 1 + 16 l2
N12 . Again using orthogonality of buckling modes and equations
(13), (18) and (22), we arrive at the following normalized bending energy variation
expressions
ˆ 1 ∞
1  N4 1
2 1X 2 4
∂Eb = ∂[ W0∗∗ (X) ∗∗
− W (X) dX] = ∂[ 1 ( − A1 )2 + A N ]. (28)
2 0 4 2 4 j=2 j j

Utilizing equations (14), (17), (18), (24), (25) and (27), we obtain the normalized
compression energy variation

N2 X 2 2
∂Ec = − ∂[ A N ]. (29)
4 j=1 j j

Work done by external force can be found from equations (15), (18) and (26)

X
 
∂Ef = −F ∂D = 2F ∂ Aj . (30)
j=1,5,9,...
20

The variation of total energy ∂Et = ∂Eb + ∂Ec + ∂Ef should satisfy ∂Et ≥ 0. If
equations (28) through (30) are substituted into the total energy, we obtain the following
expression in terms of the buckling mode coefficients
N2 N12  X  Nj2 
1 2 2 2 2
(N1 − N )A1 − + 2F ∂A1 + (Nj − N )Aj + 2F ∂Aj
2 4 j=5,9,13,...
2
(31)
X Nj2 2 2 2
(Nj − N )∂Aj ≥ 0
j=2,3,4,6,7,8,...
4

Condition in equation (31) should be satisfied for positive and negative variations of
Aj . Therefore, the expressions inside the parantheses in the first and second terms of
equation (31) must be equal to zero. Therefore, the following equations are obtained
for some coefficients of the buckling modes
1  4F N12 
A1 = 2 − (32a)
N − N12 N12 2
4F
Aj = 2 2 , j = 5, 9, 13, ... (32b)
Nj (N − Nj2 )
There is no expression for other coefficients (i.e. j = 2, 3, 4, 6, 7, 8, ...) for the general
case. However, we can make the following mathematical arguments. From equation
(32) it is clear that these coefficients are functions of only two variables Aj = Aj (N, F ),
j = 1, 5, 9, .... While the last summation term in equation (31) shows that other
coefficients Aj , j = 2, 3, 4, 6, 7, 8, ... do not depend on F . Also from equation (26):
D = D(Aj ), j = 1, 5, 9, .... Therefore, only coefficients in equation (32) describe the
force F versus displacement D equation, while other buckling modes are not involved
in this relationship.

8. References

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


honeycomb materials,” Journal of Materials Research, vol. 33, no. 3, p. 290–299, 2018.
[2] D. M. Correa, T. Klatt, S. Cortes, M. Haberman, D. Kovar, and C. Seepersad, “Negative stiffness
honeycombs for recoverable shock isolation,” Rapid Prototyping Journal, vol. 21, no. 2, pp.
193–200, 2015.
[3] B. M. Goldsberry and M. R. Haberman, “Negative stiffness honeycombs as tunable elastic
metamaterials,” Journal of Applied Physics, vol. 123, no. 9, p. 091711, 2018.
[4] D. Restrepo, N. D. Mankame, and P. D. Zavattieri, “Phase transforming cellular materials,”
Extreme Mechanics Letters, vol. 4, pp. 52 – 60, 2015.
[5] P. Cazottes, A. Fernandes, J. Pouget, and M. Hafez, “Bistable buckled beam: Modeling of
actuating force and experimental validations,” ASME. J. Mech. Des., vol. 131, no. 10, pp.
1–10, 2009.
[6] D. M. Correa, C. C. Seepersad, and M. R. Haberman, “Mechanical design of negative stiffness
honeycomb materials,” Integrating Materials and Manufacturing Innovation, vol. 4, no. 1, p. 10,
Jul 2015.
[7] B. A. Fulcher, D. W. Shahan, M. R. Haberman, C. C. Seepersad, and P. S. Wilson, “Analytical and
experimental investigation of buckled beams as negative stiffness elements for passive vibration
and shock isolation systems,” Journal of Vibration and Acoustics, vol. 136, pp. 1–12, 2014.
21

[8] L. Kashdan, C. C. Seepersad, M. Haberman, and P. S. Wilson, “Design, fabrication, and evaluation
of negative stiffness elements using sls,” Rapid Prototyping Journal, vol. 18, no. 3, pp. 194–200,
2012.
[9] T. Frenzel, C. Findeisen, M. Kadic, P. Gumbsch, and M. Wegener, “Tailored buckling microlattices
as reusable light-weight shock absorbers,” Advanced Materials, vol. 28, no. 28, pp. 5865–5870,
2016.
[10] M. E. Pontecorvo, S. Barbarino, G. J. Murray, and F. S. Gandhi, “Bistable arches for morphing
applications,” Journal of Intelligent Material Systems and Structures, vol. 24, no. 3, pp. 274–286,
2013.
[11] O. A. Ganilova and J. J. Low, “Application of smart honeycomb structures for automotive passive
safety,” Proceedings of the Institution of Mechanical Engineers, Part D: Journal of Automobile
Engineering, vol. 232, no. 6, pp. 797–811, 2018.
[12] X. Yu, J. Zhou, H. Liang, Z. Jiang, and L. Wu, “Mechanical metamaterials associated with stiffness,
rigidity and compressibility: A brief review,” Progress in Materials Science, vol. 94, pp. 114 –
173, 2018.
[13] C. Ren, D. Yang, and H. Qin, “Mechanical performance of multidirectional buckling-based negative
stiffness metamaterials: An analytical and numerical study,” Materials (Basel), vol. 11, no. 7,
pp. 1–19, June 2018.
[14] A. Rafsanjani, A. Akbarzadeh, and D. Pasini, “Snapping mechanical metamaterials under tension,”
Advanced Materials, vol. 27, no. 39, pp. 5931–5935, 2015.
[15] J. Liu, H. Qin, and Y. Liu, “Dynamic behaviors of phase transforming cellular structures,”
Composite Structures, vol. 184, pp. 536 – 544, 2018.
[16] T. A. M. Hewage, K. L. Alderson, A. Alderson, and F. Scarpa, “Double-negative mechanical
metamaterials displaying simultaneous negative stiffness and negative poisson’s ratio properties,”
Advanced Materials, vol. 28, no. 46, pp. 10 323–10 332, 2016.
[17] K. Virk, A. Monti, T. Trehard, M. Marsh, K. Hazra, K. Boba, C. D. L. Remillat, F. Scarpa, and
I. R. Farrow, “SILICOMB PEEK kirigami cellular structures: mechanical response and energy
dissipation through zero and negative stiffness,” Smart Materials and Structures, vol. 22, no. 8,
p. 084014, jul 2013.
[18] G. Puglisi and L. Truskinovsky, “Mechanics of a discrete chain with bi-stable elements,” Journal
of the Mechanics and Physics of Solids, vol. 48, no. 1, pp. 1 – 27, 2000.
[19] N. Nadkarni, C. Daraio, and D. M. Kochmann, “Dynamics of periodic mechanical structures
containing bistable elastic elements: From elastic to solitary wave propagation,” Phys. Rev. E,
vol. 90, p. 023204, Aug 2014.
[20] S. Shan, S. H. Kang, J. R. Raney, P. Wang, L. Fang, F. Candido, J. A. Lewis, and K. Bertoldi,
“Multistable architected materials for trapping elastic strain energy,” Advanced Materials,
vol. 27, no. 29, pp. 4296–4301, 2015.
[21] B. Haghpanah, L. Salari-Sharif, P. Pourrajab, J. Hopkins, and L. Valdevit, “Multistable shape-
reconfigurable architected materials,” Advanced Materials, vol. 28, no. 36, pp. 7915–7920, 2016.
[22] T. Klatt, M. Haberman, and C. Seepersad, “Selective laser sintering of negative stiffness
mesostructures for recoverable, nearly-ideal shock isolation,” 1 2013, pp. 1010–1022, 24th
International Solid Freeform Fabrication Symposium - An Additive Manufacturing Conference,
SFF 2013 ; Conference date: 12-08-2013 Through 14-08-2013.
[23] K. Che, C. Yuan, J. Wu, J. Qi HH, and J. Meaud, “Three-dimensional-printed multistable
mechanical metamaterials with a deterministic deformation sequence,” ASME. Journal of
Applied Mechanics, vol. 84, no. 1, pp. 1–10, September 2016.
[24] J. Qiu, J. H. Lang, and A. H. Slocum, “A curved-beam bistable mechanism,” Journal of
Microelectromechanical Systems, vol. 13, no. 2, pp. 137–146, April 2004.
[25] A. Amado-Becker, J. Ramos-Grez, M. J. Yañez, Y. Vargas, and L. Gaete, “Elastic tensor stiffness
coefficients for sls nylon 12 under different degrees of densification as measured by ultrasonic
technique,” Rapid Prototyping Journal, vol. 14, no. 5, pp. 260 – 270, 2008.
22

[26] S. P. Timoshenko and J. M. Gere, Theory of Elastic Stability, 2nd ed. McGraw-Hill, 1985.
[27] P. Hassanpour, E. Esmailzadeh, W. Cleghorn, and J. Mills, “Generalized orthogonality condition
for beams with intermediate lumped masses subjected to axial force,” Journal of Vibration and
Control, vol. 16, no. 5, pp. 665–683, 2010.
[28] K. D. Hjelmstad, Fundamentals of Structural Mechanics, 2nd ed. Springer, 2005.
[29] W.-F. Chen and T. Atsuta, Theory of Beam Columns: Volume 1. J Ross Publishing, 2007.

View publication stats

You might also like