ArticlePDF Available

Abstract and Figures

Recent advances in additive manufacturing have enabled fabrication of phononic crystals and metamaterials which exhibit spectral gaps, or stopbands, in which the propagation of elastic waves is prohibited by Bragg scattering or local resonance effects. Due to the high level of design freedom available to additive manufacturing, the propagation properties of the elastic waves in metamaterials are tunable through design of the periodic cell. In this paper, we outline a new design approach for metamaterials incorporating internal resonators, and provide numerical and experimental evidence that the stopband exists over the irreducible Brillouin zone of the unit cell of the metamaterial (i.e. is a three-dimensional stopband). The targeted stopband covers a much lower frequency range than what can be realised through Bragg scattering alone. Metamaterials have the ability to provide (a) lower frequency stopbands than Bragg-type phononic crystals within the same design volume, and/or (b) comparable stopband frequencies with reduced unit cell dimensions. We also demonstrate that the stopband frequency range of the metamaterial can be tuned through modification of the metamaterial design. Applications for such metamaterials include aerospace and transport components, as well as precision engineering components such as vibration-suppressing platforms, supports for rotary components, machine tool mounts and metrology frames.
Wave propagation properties of the internally resonating metamaterial: (a) Dispersion curves for the metamaterial with í µí±† í µí±‘ /í µí° ¶ and í µí±† í µí±™ /í µí° ¶ values of 0.033 and 0.1, respectively, with eigenmodes at selection of high symmetry points, and (b) start and end frequencies of the complete stopbands of metamaterials of different í µí±† í µí±‘ /í µí° ¶ values with the struts connected to resonators of large-size (green), mid-size (blue), and small-size (orange). The indicated percentages show the relative gap to mid-gap percentage. All frequencies (í µí±“) are normalised to the longitudinal wave speed in the medium í µí±£ and the unit cell size í µí° ¶. The dispersion curves of multiple metamaterials of different values of í µí±† í µí±‘ /í µí° ¶ and í µí±† í µí±™ /í µí° ¶ were predicted. The considered í µí±† í µí±‘ /í µí° ¶ values were 0.005, 0.01, 0.02, 0.025 and 0.033, and the considered í µí±† í µí±™ /í µí° ¶ values were 0.05 (large-size resonator), 0.1 (mid-size resonator) and 0.2 (small-size resonator). Figure 2b presents the stopbands for each of the considered metamaterials to show the impact of the design of the internal resonators on forming complete 3D stopbands. The relative gap to mid-gap percentages of selection of the presented stopbands (width of the stopband divided by its central frequency) are highlighted in Figure 2b. The large-size resonator showed the largest relative gap to mid-gap percentage of 68 %. The cut-on frequency of the top acoustic branches (i.e. the stopband end frequency) increased with the increase in the diameter of the struts, and with the increase in the size of the resonator. The stopbands of all the considered unit cell designs were below a normalised frequency of 0.1, as can be seen in Figure 2b. The stopbands of the large-size resonator had wider stopbands than that of the mid-size resonator. The average stopband width in the large-size resonator was calculated to be wider by 63 %, and 236 % than that of mid-size and small-size resonators, respectively. The mean frequency of the stopband showed a change of 2.4 % with the change in the resonator size. The results shown in Figure 2b can be used as a means of tuning the stopbands of the metamaterial for a specific application.
… 
Content may be subject to copyright.
1
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
Three-dimensional resonating
metamaterials for low-frequency
vibration attenuation
W. Elmadih1, D. Chronopoulos2, W. P. Syam1, I. Maskery3, H. Meng2 & R. K. Leach
1
Recent advances in additive manufacturing have enabled fabrication of phononic crystals and
metamaterials which exhibit spectral gaps, or stopbands, in which the propagation of elastic waves
is prohibited by Bragg scattering or local resonance eects. Due to the high level of design freedom
available to additive manufacturing, the propagation properties of the elastic waves in metamaterials
are tunable through design of the periodic cell. In this paper, we outline a new design approach for
metamaterials incorporating internal resonators, and provide numerical and experimental evidence
that the stopband exists over the irreducible Brillouin zone of the unit cell of the metamaterial (i.e. is a
three-dimensional stopband). The targeted stopband covers a much lower frequency range than what
can be realised through Bragg scattering alone. Metamaterials have the ability to provide (a) lower
frequency stopbands than Bragg-type phononic crystals within the same design volume, and/or (b)
comparable stopband frequencies with reduced unit cell dimensions. We also demonstrate that the
stopband frequency range of the metamaterial can be tuned through modication of the metamaterial
design. Applications for such metamaterials include aerospace and transport components, as well
as precision engineering components such as vibration-suppressing platforms, supports for rotary
components, machine tool mounts and metrology frames.
Phononic crystals (PCs) are engineered materials designed to control elastic wave propagation. PCs generally rely
on high impedance mismatches within their structural periodicity to form Bragg-type stopbands that exist due
to the destructive interference between transmitted and reected waves. e presence of destructive interference
prevents specic wave types from propagating. Kushwaha et al.1 presented the rst comprehensive calculation of
acoustic bands in a structure of periodic solids embedded in an elastic background. James et al.2 used a periodic
array of polymer plates submerged in water and provided experimental realisation of one-dimensional (1D)
and two-dimensional (2D) PCs. Montero de Espinosa et al.3 used aluminium alloy plates with cylindrical holes
lled with mercury to generate 2D ultrasonic stopbands. Tanaka et al.4 studied the homogeneity of PCs in the
perpendicular direction to the direction of propogation, and classied PCs into bulk PCs and slab PCs. Research
on the design, manufacturing and testing of PCs has mainly focused on 1D and 2D PCs515, although recently,
the research has been extended to include 3D PCs1624. Lucklum et al.25 discussed the manufacturing challenges
of 3D PCs and showed that additive manufacturing (AM) has the fabrication capabilities required for the real-
isation of geometrically complex 3D PCs2629. ere are a wide variety of AM technologies that may be used to
manufacture PC materials, such as laser powder bed fusion (LPBF), photo-polymerization, stereolithography
and inkjet printing3033. Although diering in the manufacturing resolution (the thickness of the build layer),
materials, design constraints and cost, these AM technologies create 3D parts from a CAD model. e creation of
the 3D parts is usually carried out layer by layer, and the thickness of the deposited layers, as well as the eects of
post-processing, determine the geometrical quality of the created 3D parts34,35.
Despite the benets of the recent ability to manufacture PCs with AM, their eectiveness at low-frequencies
is limited due to the dependency of the resulting stopbands on Bragg scattering. Bragg scattering occurs due
to destructive interference of the propagating waves with the in-phase reected waves, which occurs when the
wavelengths of the reected and propagating waves are similar. e reection occurs due to the dierence in the
1Manufacturing Metrology Team, Faculty of Engineering, University of Nottingham, Nottingham, NG8 1BB,
UK. 2Institute for Aerospace Technology & Composites Group, Faculty of Engineering, University of Nottingham,
Nottingham, NG8 1BB, UK. 3Centre for Additive Manufacturing, Faculty of Engineering, University of Nottingham,
Nottingham, NG8 1BB, UK. Correspondence and requests for materials should be addressed to W.E. (email: Waiel.
Elmadih@Nottingham.ac.uk)
Received: 7 April 2019
Accepted: 15 July 2019
Published: xx xx xxxx
OPEN
2
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
www.nature.com/scientificreports/
impedance (e.g. local density) of the PC. For the in-phase reection to occur, the Bragg law has to be satised36,
which is highly dependent on the cell size of the PC. Bragg scattering starts to occur when the wavelength is
approximately equal to twice the cell size of the PC36; around a normalised frequency (the quotient of cell size and
wavelength) of 0.5. us, there is a limiting dependency on the size of the unit cell of the PCs to form stopbands
by Bragg scattering. As a result of this dependency, unrealistic cell sizes need to be employed to satisfy the Bragg
law at low-frequencies.
It is possible to form stopbands below the lowest Bragg limit using metamaterials with periodically arranged
local resonators. e stopbands in these metamaterials are formed by absorbing wave energy around the resonant
frequency3744. e benets of resonator-based metamaterials include increased design freedom and the exibility
to obtain stopbands in structures of higher periodicity within a xed design volume compared to conventional
PCs. us, resonator-based metamaterials provide better-dened stopbands. Research on locally resonant meta-
materials includes the work of Liu et al.44, who rst developed a metamaterial using solid cores and silicone rub-
ber coatings. e periodically coated spheres of Liu et al. exhibited negative dynamic mass, as well as stopbands
at low frequencies. Numerous locally resonant metamaterials have been proposed. An example by Fang et al.45
showed arrays of Helmholtz resonators with negative dynamic bulk modulus. Qureshi et al.46 numerically inves-
tigated the existence of stopbands in cantilever-in-mass metamaterials. Lucklum et al.21 and D’Alessandro et al.47
independently veried the existence of stopbands in ball-rod metamaterials. Zhang et al.48 presented results of
a beam metamaterial with local resonance stopbands. Bilal et al.49 reported on the concept of combining local
resonance with Bragg scattering to form trampoline metamaterial with subwavelength stopbands. Matlack et al.50
developed a multimaterial structure that has wide stopbands using similar concept to that of Bilal et al.49. Most
of the above work, regarding both PCs and metamaterials, has employed analytical techniques to model and
optimise the suggested unit cells. Because analytical techniques can only model simple designs, the potential for
exploring the elastic capabilities of complex metamaterial designs has been limited.
We hereby report on 3D metamaterial comprising internal resonators, designed for targeting maximum elastic
wave attenuation below a normalised frequency of 0.1. is normalised frequency limit, chosen arbitrarily, is four
times lower than the lowest theoretical limit allowed for Bragg scattering stopbands. Due to its high normalised
stopband frequencies, a PC relies heavily on increasing the cell size to reduce the absolute stopband frequency.
e low normalised stopband frequencies of metamaterials allow for vibration attenuation at low absolute fre-
quencies using much more practical unit cell sizes (i.e. of more suitable dimensions for AM and applications). A
novel approach for tuning and designing the unit cell of the metamaterial is presented. e computation scheme
of the wave dispersion curves uses nite element (FE) modelling. In comparison to nite dierence time domain
(FDTD) modelling which suers from stair-casing eects51, and plain wave expansion (PWE) modelling which is
limited to structures of low impedance mismatch52, FE modelling guarantees an accurate description of the wave
dynamics within the 3D metamaterial. LPBF is employed for fabrication of the metamaterial, which is experimen-
tally tested for verication of the numerical predictions. e fundamental unit cell of the metamaterial is shown
in Figure1, and is periodically tessellated in 3D to allow a local resonance eect. e 3D wave propagation and
the complete stopbands of the metamaterial are presented in Figure2. e experimental response of the manufac-
tured metamaterial is shown in Figure3. Details of the computation, manufacturing and experimental methods
are provided in the subsequent sections.
Results and Discussion
e unit cell of the metamaterial featured in this work is shown in Figure1. e design is a cubic unit cell with
face-centered struts (FCC), and reinforcement struts in the x-, y- and z- directions (FCCxyz). FCC lattices gen-
erally have good compressive strength53, in comparison to body-centred cubic lattices (BCC). us, the FCCxyz
lattice is used as the host for the internal resonance mechanism of the metamaterial. e internal resonance
mechanism consists of six struts; each connects one side of a cubic mass to the inner walls of the FCCxyz unit cell.
Increasing the strut diameter Sd would increase the stiness of the resonator, while increasing the strut length Sl
would alter its volume fraction, which will have an impact on the stopband frequencies and the total mass.
Figure 1. e design of the resonating metamaterial: (a) Schema of the single unit cell of the metamaterial as
modelled in CAD, the labels show the strut diameter (Sd), strut length (Sl), and cell size (C), and photograph of
the 3 × 3 × 3 metamaterial as (b) digitally rendered, and (c) manufactured with LPBF.
3
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
www.nature.com/scientificreports/
Figure 2. Wave propagation properties of the internally resonating metamaterial: (a) Dispersion curves for
the metamaterial with Sd/C and Sl/C values of 0.033 and 0.1, respectively, with eigenmodes at selection of high
symmetry points, and (b) start and end frequencies of the complete stopbands of metamaterials of dierent
Sd/C values with the struts connected to resonators of large-size (green), mid-size (blue), and small-size
(orange). e indicated percentages show the relative gap to mid-gap percentage. All frequencies (f) are
normalised to the longitudinal wave speed in the medium v and the unit cell size C.
Figure 3. Experimental results acquired for the resonating metamaterial: (a) Transmissibility of the 3 × 3 × 3
metamaterial in the x- longitudinal direction (solid line), y- transverse direction (dotted line), and z- transverse
direction (dashed line) vis-à-vis the corresponding stopband as illustrated by the dispersion curves of the
innite metamaterial shown in (b), and (c) representative photograph of the experimental setup. e shaded
areas show the identied stopbands.
4
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
www.nature.com/scientificreports/
Modelling of the elastic wave propagation in the metamaterials was carried out in 3D using the scheme
described in the Methods Section. e modelling used sucient tetrahedral elements, such that the frequency
of the rst vibration mode converged with the FE mesh density (approximately 6000 nodes per unit cell). e
elements of the converged mesh used three degrees of freedom (DOF) per node with adaptive mesh size to suf-
ciently model narrow regions in the metamaterials54. To mathematically model the elastic wave propagation,
the contours of the irreducible Brillouin zone (IBZ) of the unit cells of the metamaterials were scanned. Several
characteristic points exist within the contours of the IBZ including Γ(0,0,0), X(π/C,0,0), M(π/C,π/C,0), and
R(π/C,π/C,π/C), where C is the unit cell size (also referred to as α or L in other literature50,55,56). e scan of the
IBZ was carried out using a total of 360 combinations of wavenumbers (90 combinations for each wave propaga-
tion direction). e corresponding dispersion properties along the path Γ–X–R–M–Γ of the IBZ were predicted
and the complete stopbands were identied. e dispersion curves of a metamaterial unit cell with Sd/C and Sl/C
values of 0.033 and 0.1, respectively, are presented in Figure2a. It was observed that the metamaterial exhibits a
stopband below a normalised frequency of 0.1. e stopband spans a normalised frequency range of 0.028, start-
ing from 0.039 to 0.067, and is formed by an internal resonance that cuts the rst three acoustic wavebands (wave-
bands cutting-on at zero frequency) and splits them into two branches (i.e. top and bottom acoustic branches).
e dispersion curves of multiple metamaterials of dierent values of Sd/C and Sl/C were predicted. e con-
sidered Sd/C values were 0.005, 0.01, 0.02, 0.025 and 0.033, and the considered Sl/C values were 0.05 (large-size
resonator), 0.1 (mid-size resonator) and 0.2 (small-size resonator). Figure2b presents the stopbands for each of
the considered metamaterials to show the impact of the design of the internal resonators on forming complete
3D stopbands. e relative gap to mid-gap percentages of selection of the presented stopbands (width of the stop-
band divided by its central frequency) are highlighted in Figure2b. e large-size resonator showed the largest
relative gap to mid-gap percentage of 68%. e cut-on frequency of the top acoustic branches (i.e. the stopband
end frequency) increased with the increase in the diameter of the struts, and with the increase in the size of the
resonator. e stopbands of all the considered unit cell designs were below a normalised frequency of 0.1, as can
be seen in Figure2b. e stopbands of the large-size resonator had wider stopbands than that of the mid-size res-
onator. e average stopband width in the large-size resonator was calculated to be wider by 63%, and 236% than
that of mid-size and small-size resonators, respectively. e mean frequency of the stopband showed a change of
2.4% with the change in the resonator size. e results shown in Figure2b can be used as a means of tuning the
stopbands of the metamaterial for a specic application.
For verication of the complete stopband in the proposed metamaterial, LPBF was used to manufacture a 3D
structure of nite periodicity. Details about the LPBF process can be found in the Methods Section. e geomet-
rical dimensions and periodicity of the metamaterial were selected to be suitable for the LPBF process. e manu-
factured metamaterial, presented in Figure1c, had a unit cell size of 30 mm and a 3D periodicity of three. e Sd/C
and Sl/C values were selected to provide the lowest stopband start frequency, when referenced to the stopband
start frequencies presented in Figure2b while considering the lowest manufacturable feature size with LPBF57
(See Methods Section); this meant that the Sd/C and Sl/C values had to be 0.033 and 0.1, respectively. e 3D
transmissibility of the metamaterial was obtained experimentally and is presented in Figure3a. e longitudinal
transmissibility had a value of 0 dB near the normalised frequency of zero, which indicates complete transmission
of the excitation waves. At the vibration resonances, the longitudinal transmissibility was greater than 0 dB and
reached 28 dB, which indicates high amplication of the excitation waves. Within the stopband, the longitudinal
transmissibility reached 77 dB. e eect of lattice periodicity on the transmissibility within the stopband can
be seen elsewhere12,58. For this investigation, considering the manufacturable feature size of LPBF (See Methods
Section), we have chosen 3 × 3 × 3 as a reasonable example. e results showed that the metamaterial in this work
has double the transmissibility reduction experimentally reported by Croënne et al.12 for their 3D PC which had
double the spatial periodicity used in this work.
e 3D elastic wave propagation in the internally resonating metamaterials was modelled using a hybrid
scheme. e scheme uses the FE method combined with innite periodicity assumptions. It was shown that the
metamaterials exhibit complete stopbands far below the lowest frequency limit of Bragg-type stopbands, which
exist in traditional PCs. A metamaterial of nite periodicity was manufactured using LPBF. An experimental
setup was assembled, comprising a broadband vibration shaker, a laser vibrometer, and dedicated signal gen-
eration and acquisition units. e experimental setup was used to test the 3D vibration transmissibility of the
manufactured metamaterial. It was shown that the metamaterial could attenuate the vibration waves within the
stopband range. e experimental results showed that, within the stopband, the longitudinal transmissibility
of vibration waves in the metamaterial reached 77 dB. Tuning of the stopband can be achieved by adjusting
the size of the resonator and the diameter of the struts to suit the requirements of various applications. For this
particular metamaterial, the stopband was from 1.63 kHz to 2.8 kHz with a unit cell size of 30 mm. Unit cells of
suitable dimensions for AM and applications, and higher periodicity within a certain design volume, in compar-
ison to PCs, can be employed to obtain low absolute frequency stopbands; resulting in higher attenuation within
the stopbands.
Methods
Modelling of elastic wave propagation using a hybrid wave and nite element scheme. e
proposed scheme for computing the dispersion curves used a combination of FE modelling and periodic struc-
ture theory. e metamaterials were modelled using FE modelling which allows for accurate representation of the
geometrically complex metamaterials. e complete mass and stiness matrices of the designs, K and, M respec-
tively, were extracted. e Bloch theorem59, which governs the periodic displacement and forcing conditions was
employed. e periodic structure theory assumed an innite 3D spatial periodicity of the unit cell60,61. Figure4 is
a schema of the segmentation of the unit cell of the metamaterial into sets of DOF, which were used for modelling
the periodicity of the unit cell.
5
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
www.nature.com/scientificreports/
e nodal displacement matrices q were arranged in the following sequence to allow for the 3D spatial peri-
odicity of the unit cell
=
qqqqqqqqq qqqqqqqqqqq[],(1)
IN FSBTLRFB FT SB ST FL FR SL SR BL BR TL TR T
where the subscripts IN, L, R, T, B, F, and S indicate the DOF of the nodes existing at the inside, le, right, top,
bottom, front, and back of the unit cell as illustrated in Figure4. A transformation matrix R was considered to
project the nodal displacement matrices as follows
=
qRq,(2)
where
=
=
−−
−−
−−
R
I
I
II
II
II
I
I
II
I
I
II
I
I
I
q
q
q
q
q
q
q
q
e
e
e
e
e
ee
e
e
ee
e
e
ee
000 000
0000 00
0000 00
00 00 00
00 00 00
00 0000
00 0000
0000 00
0000 00
0000 00
0000 00
0000 00
0000 00
0000 00
0000 00
0000 00
0000 00
0000 00
0000 00
,and ,
(3)
ik
ik
ik
ik
ik
ik ik
ik
ik
ik ik
ik
ik
ik ik
IN
F
B
L
FB
FL
BL
y
z
x
z
y
yz
x
y
xy
x
z
xz
where k is the wavenumber for the waves propagating in x-, y- and z- directions within the considered regions of
the IBZ. Subsequently, the projected stiness and mass matrices of the reduced sets of DOF,
K
and
M
, were com-
puted as
′′==KRKR MRMR,,and (4)
Assuming no external excitation under Bloch-Floquet59 boundary conditions, the following eigenvalue prob-
lem was derived in the wave domain
ϕω−=
KM()0, (5)
2
where ω is the angular frequency and ϕ is the eigenvector. Eq.5 provided the wave propagation characteristics of
the metamaterials in 3D space. By substituting a set of presumed wavenumbers in a given direction, the derived
eigenvectors ϕ provided the deformation of the unit cell under the passage of each wave type at an angular fre-
quency ω. To obtain normalised frequencies, the frequency eigenvalues of Eq.5 were normalised to the unit cell
size C and the speed of longitudinal waves in the lattice material v, which was calculated as the square root of the
quotient of the elastic modulus and material density. A complete description of each passing wave, including x-,
y- and z-directional wavenumbers and wave shapes, at a certain frequency range is acquired with modulo 2π.
When modelling the dispersion curves of the metamaterial used in this work, suitable 3D translation of all solid
features and voids within the unit cell is obtained when the design is approximated as a simple cube, thus, allow-
ing for the use of the IBZ of simple cubic lattice for modelling the dispersion curves. Such approximation is
known to provide accurate dispersion relations as can be seen elsewhere6264. e computation did not include
damping, though it should be noted that structural damping can be directly introduced to Eq.4 by including an
imaginary part of the
K
matrix65. Alternatively, if full viscous damping properties are to be considered, then
dedicated eigenvalue problem solvers can be employed59.
Additive manufacturing technology employed. Internally resonating metamaterial samples were fab-
ricated on a laser powder bed fusion (LPBF) system using PA12 polymer material. e material properties for
PA12 can be found in Table1. e LPBF system used a 21 W laser of scan speed and hatch spacing of 2500 mms1
and 0.25 mm, respectively. e nominal spot size of the laser was 0.3 mm and the layer thickness was 0.1 mm.
PA12 powder was used to ll the powder bed of dimensions 1320 mm × 1067 mm × 2204 mm at a temperature of
173 °C. Geometrical features of sizes below 0.8 mm are usually manufactured with considerable losses in mechan-
ical properties, due to the existence of unsolidied powder within the manufactured features57. To ensure that all
geometrical features were manufactured in agreement with the specied design, the size of the narrowest meta-
material feature was designed to be 1 mm57.
6
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
www.nature.com/scientificreports/
Experimental measurements on vibration transmissibility. e metamaterial sample was suspended
using piano strings to approximate free-free boundary conditions. e approach taken to suspend the metama-
terial, similar to the approach taken by Zhang et al.48 and Chen et al.66, supports the metamaterial uniformly. An
alternative approach, which can also be used for approximation of free-free boundary conditions, can be found
in the work of D’Alessandro et al.47. e metamaterial was adhesively axed from one side to a connector which
was, in turn, bolted to an acceleration sensor. e acceleration sensor was linked to the armature of the shaker
(Modal Shop Shaker 2060E)67 through a stinger. e stinger is a 1.5 mm rod which connects to the acceleration
sensor, and decouples cross-axis force inputs, thus, minimising errors during measurements68. As part of the
experimental setup, the beam of a laser vibrometer was projected perpendicularly to the opposite surface of the
metamaterial to take longitudinal acceleration measurements. e transverse acceleration measurements were
taken by projecting the beam of the laser vibrometer perpendicularly to the side surfaces of the metamaterial. e
laser vibrometer was set to measure the structural response in the longitudinal and transverse directions from
a normalised frequency of 0 to 0.15. e acceleration data within the tested frequency range were also obtained
through the acceleration sensor. e combination of the measurements of both the laser vibrometer and the
acceleration sensor provided the transmissibility of the specimen. Figure3c is a representative photograph of the
experimental setup. All measurements were taken with a normalised frequency resolution of less than 3.7 × 105
and were complexly averaged, considering both the phase and the magnitude of the measurements, over 100
spectral sweeps.
Material property Val u e
Young’s modulus 1.5 × 103 MPa
Density 950 kgm3
Table 1. Material properties of PA1269.
Figure 4. Selection of the segmentation of the unit cell of the metamaterial into DOF as used for modelling the
periodicity of the unit cell. e magenta points represent the (a) front nodes, (b) le nodes, (c) top nodes, (d)
top-le nodes, (e) top-front nodes, and (f) front-le nodes.
7
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
www.nature.com/scientificreports/
References
1. ushwaha, M. S., Halevi, P., Dobrzynsi, L. & Djafari-ouhani, B. Acoustic band structure of periodic elastic composites. Phys. ev.
Lett. 71, 2022–2025 (1993).
2. James, ., Woodley, S. M., Dyer, C. M. & Humphrey, V. F. Sonic bands, bandgaps, and defect states in layered structures—eory
and experiment. J. Acoust. Soc. Am. 97, 2041–2047 (1995).
3. de Espinosa, F. ., Jiménez, E. & Torres, M. Ultrasonic band gap in a periodic two-dimensional composite. Phys. ev. Lett. 80,
1208–1211 (1998).
4. Miyashita, T. Sonic crystals and sonic wave-guides. Meas. Sci. Technol. 16, 47–63 (2005).
5. Tanaa, Y., Tomoyasu, Y. & Tamura, S. Band structure of acoustic waves in phononic lattices: Two-dimensional composites with
large acoustic mismatch. Phys. ev. B 62, 7387–7392 (2000).
6. Chen, Y., Qian, F., Zuo, L., Scarpa, F. & Wang, L. Broadband and multiband vibration mitigation in lattice metamaterials with
sinusoidally-shaped ligaments. Extrem. Mech. Lett. 17, 24–32 (2017).
7. Bilal, O. . & Hussein, M. I. Ultrawide phononic band gap for combined in-plane and out-of-plane waves. Phys. ev. E 84, 65701
(2011).
8. Oudich, M., Assouar, M. B. & Hou, Z. Propagation of acoustic waves and waveguiding in a two-dimensional locally resonant
phononic crystal plate. Appl. Phys. Lett. 97, 193503 (2010).
9. Vasseur, J. O. et al. Experimental and theoretical evidence for the existence of absolute acoustic aand gaps in two-dimensional solid
phononic crystals. Phys. ev. Lett. 86, 3012–3015 (2001).
10. Pennec, Y. et al. Acoustic channel drop tunneling in a phononic crystal. Appl. Phys. Lett. 87, 261912 (2005).
11. uzzene, M. & Scarpa, F. Directional and band-gap behavior of periodic auxetic lattices. Phys. status solidi 242, 665–680 (2005).
12. Croënne, C., Lee, E. J. S., Hu, H. & Page, J. H. Band gaps in phononic crystals: Generation mechanisms and interaction eects. AIP
Adv. 1,41401 (2011).
13. Nassar, H., Chen, H., Norris, A. N., Haberman, M. . & Huang, G. L. Non-reciprocal wave propagation in modulated elastic
metamaterials. Proc. . Soc. A Math. Phys. Eng. Sci. 473, 20170188 (2017).
14. Phani, A. S. In Dynamics of lattice materials (eds Phani, A. S. & Hussein, M. I.) 53–59 (John Wiley and Sons, 2017).
15. ruisová, A. et al. Ultrasonic bandgaps in 3D-printed periodic ceramic microlattices. Ultrasonics 82, 91–100 (2018).
16. Wormser, M., Warmuth, F. & örner, C. Evolution of full phononic band gaps in periodic cellular structures. Appl. Phys. A 123, 661
(2017).
17. Chen, Y., Yao, H. & Wang, L. Acoustic band gaps of three-dimensional periodic polymer cellular solids with cubic symmetry. J. Appl.
Phys. 114 (2013).
18. Abueidda, D. W., Jasiu, I. & Sobh, N. A. Acoustic band gaps and elastic stiness of PMMA cellular solids based on triply periodic
minimal surfaces. Mater. Des. 145, 20–27 (2018).
19. Bücmann, T. et al. Tailored 3d mechanical metamaterials made by dip-in direct-laser-writing optical lithography. Adv. Mater. 24,
2710–2714 (2012).
20. Bilal, O. ., Ballagi, D. & Daraio, C. Architected lattices for simultaneous broadband attenuation of airborne sound and mechanical
vibrations in all directions. Phys. ev. Appl. 10, 54060 (2018).
21. Luclum, F. & Velleoop, M. J. Bandgap engineering of three-dimensional phononic crystals in a simple cubic lattice. Appl. Phys.
Lett. 113, 201902 (2018).
22. Tanier, S. & Yilmaz, C. Design, analysis and experimental investigation of three-dimensional structures with inertial amplication
induced vibration stop bands. Int. J. Solids Struct. 72, 88–97 (2015).
23. Zhou, X.-Z., Wang, Y.-S. & Zhang, C. Eects of material parameters on elastic band gaps of two-dimensional solid phononic crystals.
J. Appl. Phys. 106, 14903 (2009).
24. Warmuth, F., Wormser, M. & örner, C. Single phase 3D phononic band gap material. Sci. ep.7, 3843 (2017).
25. Luclum, F. & Velleoop, M. J. Design and fabrication challenges for millimeter-scale three-dimensional phononic crystals. Crystals
7, 348 (2017).
26. Zheng, X. et al. Multiscale metallic metamaterials. Nat. Mater. 15, 1100 (2016).
27. Wang, Q. et al. Lightweight mechanical metamaterials with tunable negative thermal expansion. Phys. ev. Lett. 117, 175901 (2016).
28. Li, X. & Gao, H. Mechanical metamaterials: smaller and stronger. Nat. Mater. 15, 373 (2016).
29. ompson, M. . et al. Design for additive manufacturing: Trends, opportunities, considerations, and constraints. CIP Ann. 65,
737–760 (2016).
30. Conner, B. P. et al. Maing sense of 3-D printing: Creating a map of additive manufacturing products and services. Addit . Manuf.
1–4, 64–76 (2014).
31. Vaezi, M., Seitz, H. & Yang, S. A review on 3D micro-additive manufacturing technologies. Int. J. Adv. Manuf. Technol. 67, 1721–1754
(2013).
32. Singh, S., amarishna, S. & Singh, . Material issues in additive manufacturing: A review. J. Manuf. Process. 25, 185–200 (2017).
33. Guo, N. & Leu, M. C. Additive manufacturing: Technology, applications and research needs. Front. Mech. Eng. 8, 215–243 (2013).
34. Islam, M. N., Boswell, B. & Pramani, A. An investigation of dimensional accuracy of parts produced by three-dimensional printing.
In the World Congress on Engineering 2013,522–525 (2013).
35. Lee, P., Chung, H., Lee, S. W., Yoo, J. & o, J. eview: Dimensional accuracy in additive manufacturing processes. In ASME.
International Manufacturing Science and Engineering Conference. 1, V001T04A045 (2014).
36. Maldovan, M. Phonon wave interference and thermal bandgap materials. Nat. Mater. 14, 667 (2015).
37. aghavan, L. & Phani, A. S. Local resonance bandgaps in periodic media: Theory and experiment. J. Acoust. Soc. Am. 134,
1950–1959 (2013).
38. Nouh, M., Aldraihem, O. & Baz, A. Wave propagation in metamaterial plates with periodic local resonances. J. Sound Vib. 341,
53–73 (2015).
39. Wang, P., Casadei, F., ang, S. H. & Bertoldi, . Locally resonant band gaps in periodic beam lattices by tuning connectivity. Phys.
e v. B 91, 20103 (2015).
40. Nouh, M. A., Aldraihem, O. J. & Baz, A. Periodic metamaterial plates with smart tunable local resonators. J. Intell. Mater. Syst. Struct.
27, 1829–1845 (2015).
41. Bacigalupo, A. & Gambarotta, L. Simplied modelling of chiral lattice materials with local resonators. Int. J. Solids Struct. 83,
126–141 (2016).
42. Sharma, B. & Sun, C. T. Local resonance and Bragg bandgaps in sandwich beams containing periodically inserted resonators.
J. Sound Vib. 364, 133–146 (2016).
43. Yilmaz, C., Hulbert, G. M. & iuchi, N. Phononic band gaps induced by inertial amplication in periodic media. Phys. ev. B 76,
54309 (2007).
44. Liu et al. Locally resonant sonic materials. Science 289, 1734–1736 (2000).
45. Fang, N. et al. Ultrasonic metamaterials with negative modulus. Nat. Mater. 5, 452–456 (2006).
46. Qureshi, A., Li, B. & Tan, . T. Numerical investigation of band gaps in 3D printed cantilever-in-mass metamaterials. Sci. ep. 6,
28314 (2016).
47. D’Alessandro, L., Belloni, E., Ardito, ., Corigliano, A. & Braghin, F. Modeling and experimental verication of an ultra-wide
bandgap in 3D phononic crystal. Appl. Phys. Lett. 109, 221907 (2016).
8
SCIENTIFIC REPORTS | (2019) 9:11503 | https://doi.org/10.1038/s41598-019-47644-0
www.nature.com/scientificreports
www.nature.com/scientificreports/
48. Zhang, H., Xiao, Y., Wen, J., Yu, D. & Wen, X. Flexural wave band gaps in metamaterial beams with membrane-type resonators:
eory and experiment. J. Phys. D. Appl. Phys. 48, 435305 (2015).
49. Bilal, O. . & Hussein, M. I. Trampoline metamaterial: Local resonance enhancement by springboards. Appl. Phys. Lett. 103, 111901
(2013).
50. Matlac, . H., Bauhofer, A., rödel, S., Palermo, A. & Daraio, C. Composite 3D-printed meta-structures for low frequency and
broadband vibration absorption. Proc. Natl. Acad. Sci. 113, 8386–8390 (2015).
51. Marwaha, A., Marwaha, S. & Hudiara, I. S. Analysis of Curved Boundaries by FDTD and FE Methods. IETE J. es. 47, 301–310
(2001).
52. Qian, D. & Shi, Z. Using PWE/FE method to calculate the band structures of the semi-innite beam-lie PCs: Periodic in z-direction
and nite in x–y plane. Phys. Lett. A 381, 1516–1524 (2017).
53. Leary, M. et al. Selective laser melting (SLM) of AlSi12Mg lattice structures. Mater. Des. 98, 344–357 (2016).
54. SAS IP Inc. Mesh Generation. (2019). Available at, https://www.sharcnet.ca/Soware/Ansys/17.0/en-us/help/wb_msh/msh_tut_
asf_meshgeneration.html. (Accessed: 10th January 2019).
55. Phani, A. S., Woodhouse, J. & Flec, N. A. Wave propagation in two-dimensional periodic lattices. J. Acoust. Soc. Am. 119,
1995–2005 (2006).
56. Hsu, F. C. et al. Acoustic band gaps in phononic crystal strip waveguides. Appl. Phys. Lett. 96, 3–6 (2010).
57. Tasch, D., Mad, A., Stadlbauer, . & Schagerl, M. icness dependency of mechanical properties of laser-sintered polyamide
lightweight structures. Addit. Man uf. 23, 25–33 (2018).
58. Ampatzidis, T., Leach, . ., Tuc, C. J. & Chronopoulos, D. Band gap behaviour of optimal one-dimensional composite structures
with an additive manufactured stiener. Compos. Part B Eng. 153, 26–35 (2018).
59. Collet, M., Ouisse, M., uzzene, M. & Ichchou, M. N. Floquet–Bloch decomposition for the computation of dispersion of two-
dimensional periodic, damped mechanical systems. Int. J. Solids Struct. 48, 2837–2848 (2011).
60. Mead, D. M. Wave propagation in continuous periodic structures: research contributions from Southampton, 1964–1995. J. Sound
Vib. 190, 495–524 (1996).
61. Cotoni, V., Langley, . S. & Shorter, P. J. A statistical energy analysis subsystem formulation using nite element and periodic
structure theory. J. Sound Vib. 318, 1077–1108 (2008).
62. D’Alessandro, L. et al. Modelling and experimental verication of a single phase three-dimensional lightweight locally resonant
elastic metamaterial with complete low frequency bandgap. In 2017 11th International Congress on Engineered Materials Platforms
for Novel Wave Phenomena (Metamaterials) 70–72 (2017).
63. D’Alessandro, L., Zega, V., Ardito, . & Corigliano, A. 3D auxetic single material periodic structure with ultra-wide tunable bandgap.
Sci. ep. 8, 2262 (2018).
64. Wang, Y.-F. & Wang, Y.-S. Complete bandgap in three-dimensional holey phononic crystals with resonators. J. Vib. Acoust. 135,
41009 (2013).
65. Adhiari, S. Damping modelling using generalized proportional damping. J. Sound Vib. 293, 156–170 (2006).
66. Chen, S.-B., Wen, J.-H., Wang, G., Han, X.-Y. & Wen, X.-S. Locally resonant gaps of phononic beams induced by periodic arrays of
resonant shunts. Chinese Phys. Lett. 28, 94301 (2011).
67. e Modal Shop. 60 lbf Modal Shaer. (2010). Available at, http://www.modalshop.com/lelibrary/60lbf-Modal-Shaer-Datasheet-
(DS-0076).pdf. (Accessed: 19th February 2018).
68. e Modal Shop. Modal Exciter 60 lbf: Model 2060E. (2019). Available at, http://www.modalshop.com/excitation/60-lbf-Modal-
Exciter?ID=250. (Accessed: 10th March 2019).
69. Materialise. P. A. 12 (SLS): Datasheet. (2018). Available at, http://www.materialise.com/en/manufacturing/materials/pa-12-sls.
(Accessed: 31st January 2018).
Acknowledgements
This work was supported by the Engineering and Physical Sciences Research Council [grant number EP/
M008983/1].
Author Contributions
W.E. wrote the main body of the manuscript, performed the experimental tests and the numerical analysis
of the considered design. D.C. and W.S. contributed to the research idea and helped writing the introductory
section of the manuscript and revisited the results section. I.M. prepared the samples to be experimentally tested.
H.M. contributed to writing the introductory section. R.L. contributed to the research idea and supervised the
work conducted by his team members. All authors analysed the results together and provided feedback on the
manuscript.
Additional Information
Competing Interests: e authors declare no competing interests.
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2019
Preprint
Full-text available
Mechanical computing provides an information processing method adapting and interacting with the environment via living materials. As in electronic computing, power supply in mechanical computing is still the challenge. Designing self-powered logic gates can expand application scenarios of mechanical computing for environmental interaction. Here we formulate a framework of self-powered phononic logic gates as the basis for mechanical computing of the integrated acoustic circuit. Via tuning non-reciprocal bands, resonant band and obstacle band of a topologically imbalanced graded phononic crystal that breaks the spatial inversion symmetry, complete seven Boolean logic gates are realized on one metamaterial. The input of the logic gate, Lamb wave, is converted to the electric signal as the self-powered output by combination of the superior evanescent effect of the defect mode and the positive piezoelectric effect. An exemplify real-time heart rate monitoring powered by the graded phononic crystal is demonstrated for high-density energy conversion. The self-powered non-reciprocal phononic logic gates can be implemented on any length scale and broad external conditions.
Article
Full-text available
This paper formalizes smooth curve coloring (i.e., curve identification) in the presence of curve intersections as an optimization problem, and investigates theoretically properties of its optimal solution. Moreover, it presents a novel automatic technique for solving such a problem. Formally, the proposed algorithm aims at minimizing the summation of the total variations over a given interval of the first derivatives of all the labeled curves, written as functions of a scalar parameter. The algorithm is based on a first-order finite difference approximation of the curves and a sequence of prediction/correction steps. At each step, the predicted points are attributed to the subsequently observed points of the curves by solving an Euclidean bipartite matching subproblem. A comparison with a more computationally expensive dynamic programming technique is presented. The proposed algorithm is applied with success to elastic periodic metamaterials for the realization of high-performance mechanical metafilters. Its output is shown to be in excellent agreement with desirable smoothness and periodicity properties of the metafilter dispersion curves. Possible developments, including those based on machine-learning techniques, are pointed out.
Article
In this study, a Roman-bridge-inspired metamaterial for vibration isolation is proposed that simultaneously realizes a geometrical nonlinear structure and band gap effect, through the arrangement of periodic materials in a semicircular arched structure. Finite element analysis and band diagram analysis are performed to verify the effect of the periodic arch-structured metamaterial by calculating the transmissibility and band gap frequency range of cuboid, arched and periodic arched models. Based on the calculation results, the relationship between the geometrical parameters of the periodic arch structure and band gap frequency are analyzed, and an empirical correlation is formulated for the generation of a training database for a machine learning optimization model. The geometrical parameters of the metamaterial were optimized using the machine learning model based on an autoencoder architecture with supervised learning. An optimized metamaterial with a band gap area at three target frequencies is generated by the trained neural network, and the effectiveness of the optimized structure is validated by both numerical analysis and a frequency response test.
Article
Flexible piezoresistive strain sensors have promising applications in wearables and soft robotics. For sensing dynamic strains, such as a runner’s gait or a slipping object held by a robotic gripper, the sensor must be capable of vibration strain sensing for a range of amplitudes and frequencies. This article presents the characterization and design optimization of a flexible sensor using Schwarz P-type triply periodic minimal surface (TPMS) structures for vibration strain sensing. Sensors are fabricated using a 3D-printing process, coating a silicone rubber (SR) matrix with graphene nanoplatelets (GNP). Sensors are characterized under uniaxial compressive strain amplitudes from 0-10% and frequencies of 10-110 Hz. Frequency and time-domain analyses demonstrate sensor performance and explain the unique deformation mechanisms of the TPMS structure. Low sensor delays of less than 6.3 ms demonstrate its capability for high-frequency sensing. The frequency independence of the sensor is demonstrated, as the mean frequency dependence error is only ±3.89%. In addition, the TPMS sensor was parametrized, and a size optimization was performed on it using a multi-objective adaptive firefly algorithm (MOAFA). The MOAFA converged within 1618 unique function evaluations, a reduction of 85.5% compared to the entire set of feasible solutions. The optimized sensor was fabricated and characterized, displaying a reduced mean frequency dependence error of only ±2.18%.
Article
Full-text available
In solid state physics, a bandgap (BG) refers to a range of energies where no electronic states can exist. This concept was extended to classical waves, spawning the entire fields of photonic and phononic crystals where BGs are frequency (or wavelength) intervals where wave propagation is prohibited. For elastic waves, BGs are found in periodically alternating mechanical properties (i.e., stiffness and density). This gives birth to phononic crystals and later elastic metamaterials that have enabled unprecedented functionalities for a wide range of applications. Planar metamaterials are built for vibration shielding, while a myriad of works focus on integrating phononic crystals in microsystems for filtering, waveguiding, and dynamical strain energy confinement in optomechanical systems. Furthermore, the past decade has witnessed the rise of topological insulators, which leads to the creation of elastodynamic analogs of topological insulators for robust manipulation of mechanical waves. Meanwhile, additive manufacturing has enabled the realization of 3D architected elastic metamaterials, which extends their functionalities. This review aims to comprehensively delineate the rich physical background and the state‐of‐the art in elastic metamaterials and phononic crystals that possess engineered BGs for different functionalities and applications, and to provide a roadmap for future directions of these manmade materials. The rich theoretical framework of bandgaps is delineated in structure‐borne‐sound metamaterials and phononic crystals, while highlighting elastic bandgap engineering for contemporary applications such as surface acoustic wave devices, optomechanics, seismic wave control, and high Q‐factor sensors. This review also provides a roadmap for future directions for these artificial functional materials.
Article
Full-text available
Additive manufacturing has enabled the construction of increasingly complex mechanical structures. However, the variability of mechanical properties may be higher than that of conventionally manufactured structures. Typically, the computational cost of the numerical modeling of such structures considerably increases when variability is considered. In deterministic analyses of periodic structures, the dispersion diagrams obtained for the first Brillouin zone (FBZ) can be used to predict attenuation bands for any direction of propagation. This can be further simplified considering only the contour of the irreducible Brillouin zone (IBZ) if the unit cell presents symmetries. The objective of the current study is to present evidence that, when considering mechanical and geometric variability, the stochastic dispersion diagram along the contour of the IBZ statistically coincides with the FBZ and can be used to predict attenuation bands of finite metastructures assuming the symmetry of variability. A computationally efficient methodology is proposed, which uses the imaginary part of the wavenumbers and the Monte Carlo (MC) method for identifying stochas-tic attenuation bands that are robust against different types of variability. It is also shown that the proposed methodology can be combined with a highly * Corresponding author precise Bayesian estimator to infer the spatially distributed variability. The stochastic results obtained by scanning only the IBZ contour of the proposed two-dimensional unit cell statistically coincides with results obtained scanning the FBZ under symmetry of the spatial variability, as it occurs in deterministic cases considering deterministic symmetry. This is not a direct result, because each individual sample is asymmetric. Therefore, under symmetry of variability, the stochastic results computed for supercells and for finite metastructures consisting of a finite number of cells also coincide with the results for the IBZ contour of the unit cell.
Article
Engineering the architecture of materials is a new and very promising approach to obtain vibration isolation properties. The biggest challenge for lattice structures exhibiting vibration isolation properties is the trade-off between compactness and wide and low-frequency bandgaps, i.e., frequency ranges where the propagation of elastic or acoustic waves is prohibited. Here, we, both numerically and experimentally, propose and demonstrate a new design concept for compact metamaterials exhibiting extraordinary properties in terms of wide and low frequency bandgap and structural characteristics. With its 4 cm side length unit cell, its bandgap opening frequency of 1478 Hz, its band-stop filter behavior in the range 1.48–15.24 kHz, and its structural characteristics, the proposed 1×1×3 metastructure represents great progress in the field of vibration isolation and a very promising solution for hand-held vibration probes applications that were unattainable so far through conventional materials.
Article
Full-text available
Phononic crystals and acoustic metamaterials are architected lattices designed to control the propagation of acoustic or elastic waves. In these materials, the dispersion properties and the energy transfer are controlled by selecting the geometry of the lattices and their constitutive material properties. Most designs, however, only affect one mode of energy propagation, transmitted either as acoustic airborne sound or as elastic structural vibrations. Here, we present a design methodology to attenuate both acoustic and elastic waves simultaneously in all polarizations. We experimentally realize a three-dimensional load-bearing architected lattice, composed of a single material, that responds in a broadband frequency range in all directions and polarizations for airborne sound and elastic vibrations simultaneously.
Article
Full-text available
The design and the combination of innovative metamaterials are attracting increasing interest in the scientific community because of their unique properties that go beyond the ones of natural materials. In particular, auxetic materials and phononic crystals are widely studied for their negative Poisson's ratio and their bandgap opening properties, respectively. In this work, auxeticity and phononic crystals bandgap properties are properly combined to obtain a single phase periodic structure with a tridimensional wide tunable bandgap. When an external tensile load is applied to the structure, the auxetic unit cells change their configurations by exploiting the negative Poisson's ratio and this results in the tuning, either hardening or softening, of the frequencies of the modes limiting the 3D bandgap. Moreover, the expansion of the unit cell in all the directions, due to the auxeticity property, guarantees a fully 3D bandgap tunability of the proposed structure. Numerical simulations and analytical models are proposed to prove the claimed properties. The first experimental evidence of the tunability of a wide 3D bandgap is then shown thanks to the fabrication of a prototype by means of additive manufacturing.
Article
Full-text available
While phononic crystals can be theoretically modeled with a variety of analytical and numerical methods, the practical realization and comprehensive characterization of complex designs is often challenging. This is especially important for the nearly limitless possibilities of periodic, three-dimensional structures. In this contribution, we take a look at these design and fabrication challenges of different 3D phononic elements based on recent research using additive manufacturing. Different fabrication technologies introduce specific limitations in terms of, e.g., material choices, minimum feature size, aspect ratios, or support requirements that have to be taken into account during design and theoretical modeling. We discuss advantages and disadvantages of additive technologies suitable for millimeter and sub-millimeter feature sizes. Furthermore, we present comprehensive experimental characterization of finite, simple cubic lattices in terms of wave polarization and propagation direction to demonstrate the substantial differences between complete phononic band gap and application oriented directional band gaps of selected propagation modes.
Article
In this work, we present a comprehensive theoretical and experimental study of three-dimensional phononic crystals arranged in a simple cubic lattice. The band structure is analytically modeled as a 3D mass spring system and numerically calculated within the corresponding simple cubic Brillouin zone. We report on a design yielding a record bandgap of 166% relative width, validated by simulations and measurements of longitudinal and shear wave transmission in different spatial directions. In the additively fabricated samples, gap suppression reaches -80 dB relative to a solid reference. Comparison of different unit cell geometries showcases approaches to engineer gap width and suppression, as well as transmission bands outside the gap.
Article
Laser sintering (LS), as an additive manufacturing process for production of polymer structures, provides the possibility of directly manufacturing personalized, structural motorcycle components for motor sports. To create such lightweight structures, the wall thickness and position limits of the LS systems need to be investigated in detail. Appearing process-related flaws such as different amounts of crystallinity, surface roughness, and defects such as pores exhibit dimensions similar to the wall thickness. To study the process-related effects on the mechanical properties of 450 tensile test specimens in z-direction, the build areas of two LS systems were screened and a detailed wall thickness investigation was conducted. In addition, dynamic mechanical analysis, differential scanning calorimetry, and scanning electron microscopy for several wall thicknesses similar to the spot size were conducted. The investigations showed that the Young's moduli and ultimate tensile strengths of the produced specimens of the two commercial EOS systems, P396 and P770, are similar and evenly distributed. However, distinct differences were found in elongation at break. The scattering of mechanical properties is more in the specimens produced by P770 than in those produced by P396. The Poisson's ratio does not vary between thin- and thick-walled structures. Furthermore, structures with a thickness below 1 mm showed distinctive losses in stiffness, ultimate tensile strength, and elongation at break.
Article
In this work, the banded behaviour of composite one-dimensional structures with an additive manufactured stiffener is examined. A finite element method is used to calculate the stiffness, mass and damping matrices, and periodic structure theory is used to obtain the wave propagation of one-dimensional structures. A multi-disciplinary design optimisation scheme is developed to achieve optimal banded behaviour and structural characteristics of the structures under investigation. Having acquired the optimal solution of the case study, a representative specimen is manufactured using a carbon fibre cured plate and additive manufactured nylon-based material structure. Experimental measurements of the dynamic performance of the hybrid composite structure are conducted using a laser vibrometer and electrodynamic shaker setup to validate the finite element model.
Article
In this paper, the acoustic band structure, sound attenuation, and uniaxial elastic modulus of three cellular solids are studied computationally. The cellular solids are generated based on mathematical surfaces, called triply periodic minimal surfaces (TPMS), which include Schwarz Primitive, Schoen IWP, and Neovius surfaces. Finite element method is used to find the acoustic band gaps and sound attenuation of the TPMS structures. The numerical investigation revealed the existence of acoustic bandgaps at low frequencies and low relative densities compared to other cellular structures reported in the literature. The band gap analysis is numerically validated using structures with finite dimensions subjected to varying pressure with multiple frequencies. The influence of the porosity of TPMS on the width of the band gaps is also reported. In the considered porosity range, it is found that lower porosities result in wider acoustic band gaps. Furthermore, the uniaxial moduli of these TPMS are numerically determined using periodic boundary conditions. When the uniaxial modulus of the TPMS-structures is studied against their porosities, it is found that the response of the TPMS-structures lies between stretching- and bending-dominating.
Article
Engineering the architectures of materials is a new approach to obtain unusual properties and functionalities in solids. Phononic crystals with periodically architected microstructures and compositions exhibit omnidirectional phononic band gaps, offering a unique capability to steer mechanical wave propagation. The coupled architecture material design strategy, however, poses a significant challenge to design phononic crystals with broadband and multiband vibration control capabilities. Here we propose and demonstrate a new metamaterial design concept in which symmetry-broken ligaments with ordered topology are taken as the constitutive elements for regular lattice materials. Through integrative computational modeling, 3D printing, and vibration testing we demonstrate that the proposed lattice metamaterials can exhibit broad and multiple omnidirectional band gaps over a wide range of the geometry parameters that define the ligament. We show that the designed microstructure of the lattice metamaterial is robust and can be extended to baseline lattices with other topologies.
Article
The transmission of longitudinal ultrasonic waves through periodic ceramic microlattices fabricated by Robocasting was measured in the 2 12 MHz frequency range. It was observed that these structures (scaffolds of tetragonal and hexagonal spatial arrangements with periodicity at length-scales of 100 m) exhibit well-detectable acoustic band structures with bandgaps. The locations of these gaps were shown to be in close agreement with the predictions of numerical models, especially for the tetragonal scaffolds. For the hexagonal scaffolds, a mixing between longitudinal and shear polarizations of the propagation modes was observed in the model, which blurred the matching of the calculated band structures with the experimentally measured bandgaps.