You are on page 1of 29

Acoustic Metamaterials for Superior

Deep-Subwavelength Acoustic Imaging


Joe Peyser
December 2018

6CCP3132 Third Year Literature Review, Supervisor: Wayne Dickson

Contents
1 Introduction 2
1.1 Metamaterials Background . . . . . . . . . . . . . . . . . . . . 2
1.2 From Optical to Acoustic Metamaterials . . . . . . . . . . . . 5
1.3 Subwavelength Imaging . . . . . . . . . . . . . . . . . . . . . . 6

2 Acoustic Metamaterials 10
2.1 The Ongoing Study of Acoustic Metamaterials . . . . . . . . . 10
2.2 Fundamentals of Acoustic Metamaterials . . . . . . . . . . . . 11
2.2.1 Negative Effective Mass Density . . . . . . . . . . . . . 11
2.2.2 Negative Effective Bulk Modulus . . . . . . . . . . . . 13
2.2.3 Double Negativity in Acoustic Metamaterials . . . . . 14
2.2.4 Strong Anisotropy in Acoustic Metamaterials . . . . . 16

3 Deep-Subwavelength Acoustic Imaging 17


3.1 Constructing from Acoustic Metamaterials . . . . . . . . . . . 17
3.2 Demonstrations of Effective Acoustic Superlensing and Hyper-
lensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.1 Acoustic Superlens . . . . . . . . . . . . . . . . . . . . 19
3.2.2 Acoustic Hyperlens . . . . . . . . . . . . . . . . . . . . 21

4 Applications of Acoustic Metamaterials 22

5 Conclusion 23

1
Abstract
The field of acoustic metamaterials began as an academic curiosity
stemming from advancements in optics and has now become an active
field with promising potential applications. This review will trace the
development of acoustic metamaterials from their roots in theoreti-
cal left-handed substances, leading to initial findings of negative mass
and bulk modulus in locally resonant structures, and congregating
in new techniques for resolving acoustic sources with subwavelength
spacing. The theories governing acoustic metamaterials and super-
lenses and hyperlenses will be expounded to aid the description of
deep-subwavelength acoustic imaging techiniques.

1 Introduction
The last two decades have seen a new development from the studies of op-
tics and acoustics in the form of a new class of materials now commonly
known as metamaterials. These artificial materials are characterised by their
subwavelength scale lattice structures that can act as effective mediums. If
designed with periodic constituent elements, the period will be too small to
see a diffractive effect from propagating waves at the wavelengths of inter-
est, therefor forming an effectively homogeneous medium. Materials of this
general classification can exhibit numerous novel properties that arise from
their effective homogenisation when interacting with propagating waves.
The first conceptions of metamaterials were designed to control and ma-
nipulate electromagnetic waves, however the known parallels between the
mathematics of optics and acoustic [1, 2] fueled research into similar compos-
ite materials engineered to interact with waves of sound. Early research into
acoustic metamaterials stemmed from work with phononic crystals [3], with
focus on applications in chemical analysis, wave guides, and high-frequency
signal processing [4], but more recent studies have led to new horizons in sub-
wavelength imaging; resolving objects beyond the diffraction limit. Acoustic
imaging of this sort could have applications in the fields of medical imaging,
material testing, and sonar detection. In this review, the concept of metama-
terials and acoustic metamaterials will be investigated and considered with
the prospects of constructing devices to be used for superior subwavelength
acoustic imaging.

1.1 Metamaterials Background


There have been numerous studies of composite materials throughout the
20th century, but when considering metamaterials the roots of the field are

2
typically traced back to Viktor Veselago. Veselago is credited to be the first to
speculate on the possibility of a substance having a negative refractive index
and the characteristics of electromagnetic waves propagating through them.
He postulated that a substance having simultaneously negative permeability,
µ, and permittivity, , will result in a negative refractive index, n [5], due to
the relation:

n = − µ, (1)
The negative index is necessary to energy conservation, as an light ray prop-
agating from a medium of positive index incident on a medium of negative
index would see its wave front reverse to advance in the opposite direction
to its energy flow (Fig. 1)
When expressing Maxwell’s equations with negative µ and , one will
see a left-handed triplet formed, as opposed to the typical right-handed set
of vectors observed in positive index materials. This prompted Veselago
to label substances exhibiting double negativity as left-handed substances.
Materials of this sort would exhibit various intriguing phenomena, including
the reversal of group velocity (Fig. 1), Snell’s law [5], Doppler shift [6], and
Valinov-Cerenkov radiation [7], as well as the amplification of evanescent
waves [8].

Figure 1: Negative refraction of a light ray between a medium of positive refractive index
and a medium for which n = −1. Left: Arrows show the flow of energy in the ray. Right: Arrows
show the direction of the wave vector and group velocity, which is reversed when entering the negative
medium. [9]

This double negativity is not present in natural substances, hence little


study of this theory was made for several decades, however it has been dis-
covered that artificial materials with this property can be manufactured if an

3
approximation is made. A periodic array of conducting elements can behave
as an effective medium with homogeneous properties when the size of each
element and the spacing between them is much smaller than the wavelength
of the interacting electromagnetic fields [10].
Taking this effective medium approximation into account, in 1996, John
Pendry simulated a material of negative effective permittivity for a certain
range of microwave frequencies, being composed of a periodic structure of
infinite 1µm thick wires, arranged in a simple cubic lattice with 5mm spac-
ing [11]. Pendry then extended this concept in 1999 to a material with
negative effective permeability, constructed from a periodic array of non-
magnetic conducting sheets [12]. In the same year David Smith combined
these previous conceptions to design a composite medium of periodically
spaced non-magnetic split ring resonators and continuous wires (Fig. 2), ex-
hibiting double negativity and a negative refractive index for the first time,
albeit only for a frequency region in the microwave regime [10].

Figure 2: A negative index metamaterial composed of split ring resonators and wires de-
posited on opposite sides of standard circuit boards. Structure height is 1cm. [9]

At this point it was realised that by exploiting the effective medium ap-
proximation, crafting composite materials with constituent spacing smaller
than the wavelength of the interacting field, one could extend beyond the
constraints of previously somewhat rigid natural material properties.

4
These artificial substances have gained the label metamaterials and have
led new research into photonic circuits [13], optical cloaking [14], and sub-
diffraction-limited imaging [8, 15], among other endeavours, but the concept
of metamaterials is not limited to manipulating interactions with electromag-
netic waves. Just as metrics that describe EM waves can be found analogous
to those of acoustic waves, permittivity and permeability have their acoustic
counterparts, which can also be manipulated through consideration of the
effective medium approximation [16].

1.2 From Optical to Acoustic Metamaterials


The study of acoustic metamaterials is not a huge departure from that of
optical metamaterials. Much of the underlying physics of waves is shared
by electromagnetic and acoustic waves, which would lead one to believe it
possible to make a class of metamaterials designed for acoustics. Acoustic
waves are essentially mechanical waves, made up of periodic compressions and
extensions of the physical medium they are propagating through. Acoustic
metamaterials are effective in manipulating acoustic waves governed by New-
ton’s law of motion, the fluid continuity equation, and the thermodynamic
equation of state [17].
The effects of a medium on an acoustic wave are dictated by the mass
density, ρ, and the relative change in the volume of the medium as a reaction
to the compressive and extensive stress of the wave, referred to as the bulk
modulus κ [17]. Typically acoustic metamaterials will take advantage of
the dynamic capabilities of these properties, referring to the response of the
material to harmonically oscillating excitation rather than static forces [16],
on which this paper will elaborate.
The dependence of these metrics can be described by the continuity equa-
tion expressed as:

∇·v − P = 0, (2)
κ
as well as by Newton’s 2nd law:

∇P − iωρv = 0, (3)
where P is the pressure field and v is the velocity field.The density, ρ, and
bulk modulus, κ, are position dependent in general [18]. Considered with har-
monic field dependence e−iωt , the wave equation in a homogeneous medium
is given by:

5
ρ ∂ 2P
∇2 P − = 0. (4)
κ ∂t2
Now solving for a plane wave with wave vector k gives a refractive index of:

|k|c ρ 1
n2 = = = 2, (5)
ω κ c
q
where c is the speed of sound, given by c = κ/ρ [17]. It is clear to see
from equation (5) that in order for an acoustic wave to propagate through a
medium, ρ and κ will need to be simultaneously positive or simultaneously
negative, the latter resulting in a negative refractive index. One can draw
parallels when comparing to the case of electromagnetic waves for which,
n2EM = µ, and see that the two constitutive parameters can be mapped as
 → ρ and µ → κ−1 [17]. This analogy is simply owing to the wave equations
for acoustic and EM waves having the same mathematical form [19].
When constructing acoustic materials, mass density and bulk modulus
can take on frequency dependent effective values. Composite materials in-
corporating constituent elements with local resonances can exhibit negative
effective values for ρ and κ, and certain materials can give rise to strong
anisotropy due to the effective medium approximation [16].

1.3 Subwavelength Imaging


Pendry was first to utilise the concept of the negative refractive index to
design a perfect lens or superlens [8, 20]. The principle of a superlens is to
image subwavelength features of objects, or in other words, resolve objects
beyond the diffraction limit. Conventional optics have always been limited
to a resolution of about half a wavelength of light, and although various
methods of microscopy and spectroscopy have attempted to break past this
barrier [21], the constraining factor in these efforts is the decay of evanescent
waves in the far-field. Evanescent waves contain fine structure details of
objects, but have wave vectors that are too large to propagate into the far-
field. As a result, these waves decay exponentially with distance and the
phase correction of an ordinary lens will not restore this amplitude [8].

6
Figure 3: A doubly negative slab with refractive index n = −1. Left: Diverging waves are refracted
by the slab where they are brought to a first focal point and are refracted again when exiting the slab
to reach a second focal point. Right: The same slab here shown amplifying evanescent waves in the near
field. Here the red line represents the amplitude of the evanescent waves, decaying exponentially with
distance. [17]

Pendry discovered that using a medium of negative refractive index, in


this case a 40nm thick slab of silver, as a lens will amplify the evanescent
waves, essentially cancelling their decay (Fig. 3) [8]. For a plane wave in the
z direction with electric field E = E0 exp(−iωt + ikz z) the wave vector can
be given as:
q q
kz = ω 2 c−2 − kx2 − ky2 = i kx2 + ky2 − ω 2 c−2 , (6)
However in a material of negative index, the direction of the wave vector is
reversed (Fig. 1), giving an opposite sign to the wave vector:
q q
kz0 = − ω 2 c−2 − kx2 − ky2 = −i kx2 + ky2 − ω 2 c−2 . (7)
Putting this reversed wave vector back into the plane wave equation will
give out the factor exp(kz z), which is responsible for amplifying the non-
propagating evanescent field [8]. Energy is still conserved in this process as
the evanescent fields carry no energy [22].
This lens is capable of subwavelength imaging [15], leading to the conclu-
sion that both propagating and evanescent waves contribute to the resolution
of an image [8].
Pendry’s superlens was capable of subwavelength imaging and has been
further proven with identical and modified experimental models [15, 23, 24],
however they are generally limited by material losses [25] and noise from
surface plasmons [26]. Furthermore, the amplification of evanescent modes
allows their detection, but these waves still cannot be processed by conven-
tional optics in this form. The aim of a hyperlens is to tackle this issue

7
by transferring information carried by evanescent fields into the spectrum of
propagating waves, it would then be possible to detect these waves as well as
process them in the far field with more conventional imaging methods [22].

Figure 4: Magnifying optical hyperlens. a. Schematic of the optical hyperlens and conventional
lens layout for far-field imaging. b. Hyperlens imaging of line pair object with line-width 35 nm and
spacing of 150 nm. Left: Scanning electron microscope image of the line pair object. Centre: Magnified
hyperlens image with clearly resolved spacing. Right: Control image of same object without hyperlens
showing no spacing between lines. c. Averaged cross-section for image of line pair object in (b), red line
is with hyperlens, green line is without. d. Letters ’O’ and ’N’ with line-width 40 nm, imaged with
subwavelength resolution using hyperlens. [27]

A hyperlens is characterised by the hyperbolic functional form of the


dispersion relation produced by the strongly anisotropic metamaterials that
compose them (Fig. 5). Anisotropy is referring to the refractive index of a
material having different values for different directions, for example if nx 6=ny .
This implies different values for permittivity, x and y (or density ρx and
ρy ), are possible [28].
In an isotropic or double positive anisotropic medium, only the portion
of ky wave vector components inside the limits of the circle or ellipse (Fig.
5) will propagate in the material, losing the larger wave vector components

8
Figure 5: Possible equi-frequency contours in a 2D material. Red line denotes an isotropic medium
with circular dispersion. Blue line denotes an anisotropic medium with positive y and x with elliptical
dispersion. Black line denotes an anisotropic medium with negative y and positive x with hyperbolic
dispersion. |y | > |x | in both anisotropic cases. [16]

that carry the subwavelength details to evanescent decay [22, 16]. A larger
ratio between y and x results in stronger anisotropy and widens the ellipse
and flattens the equi-frequency contour to support higher ky values, therefor
increasing the resolution of the lens over the isotropic case. A medium of
anisotropic permittivity with opposite signs will support arbitrarily large
values for the wave vector, which in principle allows for infinite imaging
resolution.
Anisotropic metamaterials with opposite signs for permittivity in different
directions can be fabricated with use of the effectively homogenised mediums
4 [27], however the resolution is not infinite in real systems. The effective
medium approximation fails at the length scale of its constituent structural
components, and there is always a loss in material that accompanies these
negative parameters, limiting the largest supported ky .
Typical designs for hyperlenses will exploit radial geometry to magnify
sub-diffraction-limited objects [29]. The experimental example in Fig. 4a by
Liu et al. uses a curved periodic stack of 35 nm thick layers of Ag and Al2 O3
to act as a magnifying optical hyperlens [27]. This formed an anisotropic
metamaterial for which the radial and tangential permittivities had oppo-
site signs. Upon illumination of subwavelength scale objects (Fig. 4b,d),
the scattered evanescent fields enter the anisotropic medium and propagate
along the radial direction. Due to the conservation of angular momentum,
the tangential components of the wave vectors are compressed as the waves
travel outwards. The rays are also spatially dispersed by the geometry of the
hyperlens, forming a magnified image on the outer boundary of the hyper-
lens [22]. As can be seen in Fig. 4c, the low trough between the two peaks
indicates that the two subwavelength-spaced objects can be distinguished in
the magnified image.
Through the analogies that can be drawn between EM and acoustic waves,
the fundamental aspects of optical subwavelength imaging can be applied
to construct effective models for the acoustic regime [29]. To achieve this,

9
various methods that allow metamaterials to acquire effective values for mass
density, ρ, and bulk modulus, κ, are employed.

2 Acoustic Metamaterials
2.1 The Ongoing Study of Acoustic Metamaterials
Despite the early postulations of Veselago, research into metamaterials has
only been prevalent in the last twenty years with much of the public interest
in the optical branch. Alongside these developments, the field of acoustic
metamaterials has been steadily advancing, now reaching new horizons in
subwavelength imaging and acoustic cloaking [30, 31, 32, 33, 34, 35, 36, 37,
38, 39, 40].

Figure 6: Early realisation of a locally resonant acoustic metamaterial. Left: Cross-section of a


sample constituent cell, consisting of a small metallic sphere coated in a thin layer of silicone rubber with
diameter 1.55 cm. Right: Unit cells as shown on left glued together to form a cubic lattice structure. This
metamaterial exhibits a local resonance induced anomalous mass effect when interacting with acoustic
waves of wavelengths one and two orders of magnitude larger than the diameter of each unit cell. [41, 17]

Predating the term ”metamaterials”, theoretical proposals and experi-


mental realisations for phononic crystals were made in the 1990s [42, 43, 44].
These are periodically arranged lattice structures with lattice constants in
the order of the relevant elastic/acoustic wavelength, inspired by the quan-
tum mechanical band theory for solids. Phononic crystals are effective for
wave manipulation and are capable of exhibiting properties typically associ-
ated with metamaterials, such as a negative acoustic index [3], Due to sample
size constraints, the earliest phononic crystals were mostly only realised in
ultrasonic regimes [45], but the concept has been expanded to apply to lower
frequency acoustic waves (Fig. 6) [41].

10
More recently, similarly to efforts made to build materials with negative
permittivity and permeability [9], methods for designing metamaterials that
can exhibit negative effective values for mass density and bulk modulus have
been rapidly advancing [17, 16]. Prospects of simultaneously negative mass
density and bulk modulus in a metamaterial to achieve a negative acous-
tic index have driven the conception of various theoretical and experimen-
tal models. Ding et al. considered an array of rubber-coated gold spheres
overlapped with a shifted array of bubble-contained water spheres [46]. A
one-dimensional Helmholtz resonator-shunted waveguide has been shown to
display a frequency dependent modulus at ultrasonic driving forces [47], and
modifications to include rod-spring resonators theoretically allow this meta-
material to exhibit double negativity [48, 49]. A successful experimental
realisation of a negative index acoustic metamaterial was first achieved by
Lee et al. in 2010, with a 1D periodic array of interspaced elastic membranes
and side-holes at subwavelength scales [50]. This is merely a combination of
a membrane structure to induce a frequency dispersion in mass [51], and a
waveguide shunted with side-holes to simultaneously induce a similar disper-
sion for bulk modulus [52].
Since this first realisation, a multitude of approaches for retrieving these
effective constituent parameters have arisen, including layered lattices and
other methods of incorporating elastic membranes [53, 54], but most of these
techniques are still dependent of retrieving dynamic values of mass and mod-
ulus [55] through local resonances in effectively homogeneous mediums [17].

2.2 Fundamentals of Acoustic Metamaterials


Constructing acoustic metamaterials is reliant on taking advantage of nat-
urally anisotropic materials and locally resonant constituent units with size
and spacing smaller than the wavelength of acoustic waves that are being
manipulated. This subwavelength spacing is the principality of all metama-
terials and is what permits their novel properties. For acoustic metamate-
rials, the aspects that most studies aim for are strong anisotropy as well as
negative values for mass density, ρ, and bulk modulus, κ, which are typically
not found in natural materials. Acquiring these properties in new materials
requires some understanding of dynamic systems

2.2.1 Negative Effective Mass Density


Of course, there is no actual negative mass in real substances, it is achieved as
a result of inaccurate modelling of acoustic metamaterials [56]. Composite
materials that allow relative motions between the constituent components

11
can display inertial responses that differ from that of rigid bodies, for example
a bucket becoming more difficult to carry when water is sloshing around inside
of it [17].

Figure 7: Mass-spring model to describe frequency dependent effective mass. Left: A single
mass-in-mass spring system with external driving force. Right: The effective mass model of the system
on the left if observed as a single mass system. [56]

Figure 8: Mass-spring model to describe frequency dependent effective mass for propagating
waves. Infinitely long structure of mass-in-mass spring systems. [56]

Chan et al. [18] was able to introduce an effective negative mass by


employing a one-dimensional mass-spring system as a coupled oscillator. A
simple mass-in-mass unit is shown in Fig. 7. To understand the mathematics
of this effect, we consider system in Fig. 7 under an external harmonic
oscillation force, F (ω), where ω is the angular frequency. If there is no
friction between mass m1 and mass m2 , then the total force exerted on m1
is given by F (ω) + k2 (x2 − x1 ), where x1 and x2 are the displacements of m1
and m2 respectively [17]. Mass m2 undergoes harmonic oscillation with an
equation of motion given by m2 ẍ2 = −k2 (x2 − x1 ), where ẍ1,2 = −ω 2 x1,2 ,
and solving for x2 in terms of x1 , one obtains
 m2 ω0 2 
F (ω) = m1 + ẍ1 , (8)
ω0 2 − ω 2

12
q
where ω0 = k2 /m2 is the local resonant frequency of m2 [17]. If an observer
cannot see the inner structure of the system, as seen on the right in Fig. 7,
then the system’s apparent inertia acquires a frequency dispersion and an
effective mass, mef f given by

m2 ω0 2
mef f = m1 + . (9)
ω0 2 − ω 2
The expression above indicates that the effective mass is highly dependent
on the local resonant frequency ω0 and as the forcing frequency ω approaches
ω0 , the effective mass becomes negative, experiencing stronger negative effect
when closer to the local resonance [56]. It is important to note that below the
resonant frequency, the mass is displaced in the same direction as the driving
force, but pushing past the resonance causes this relation to be reversed [9].
This negative effective mass is just a result of using the equations of motion of
a single mass to represent a two mass system. This example can be extended
to wave transmission by considering an infinitely long one-dimensional lattice
system, an example of which is depicted in figure 8. However to achieve
double negativity, the modulus, k1 in this case will need to be negative.

2.2.2 Negative Effective Bulk Modulus


To describe conditions for negative effective bulk modulus, we can refer to an
experimental model demonstrated by Nicholas Fang et al. [47]. In this exper-
iment, numerous Helmholtz resonators were arranged in a one-dimensional
lattice seen in figure 9a, and submerged in water where they could be submit-
ted to ultrasonic pulses from a source on one end of the channel. A Helmholtz
resonator consists of a cavity of known volume with rigid walls and a small
hole in one side. Pressure variation in the channel causes the fluid in the
hole to oscillate in and out, producing adiabatic compression and rarefaction
of the liquid in the cavity [47]. These resonators are considered analogous to
electrical inductor-capacitor circuits [19], with the closed cavities acting as
capacitors and the necks of the holes acting as inductors.
The structures and their periodicity are designed to be considerable smaller
than the corresponding wavelength at the ultrasonic frequency range, mean-
ing this composition can be homogenised as an effective medium [16]. The ex-
perimental results seen in figure 9b, show that this composition of Helmholtz
resonators in a periodic array gives rise to a frequency dependent effective
modulus κef f , over a narrow range of frequencies, which is expressed as

Gω02
 
κ−1
ef f (ω) = κ−1
0 1− 2 , (10)
ω − ω02 + iΓω

13
(a)
(b)
Figure 9: One-dimensional ultrasonic metamaterial a. Illustration of a chain of Helmholtz resonators
forming a one-dimensional acoustic metamaterial to demonstrate negative effective bulk modulus. The
resonators are rigid walled cavities with holes opening to the closed channel of the chain. b. Experimental
results of calculated effective bulk modulus against frequency for model depicted in (a). The blue line
represents the real component of κef f and the red line represents the imaginary part of κef f . Both
imaginary and real parts display negative modulus in narrow band-widths. The real part also displays a
positve resonance at a higher frequency band. [47]

where G is a geometrical factor, ω is the pulse frequency, ω0 is the resonant


frequency, i denotes the imaginary component, and Γ is the dissipation loss
in the resonating elements [47]. This form is analogous to the expression
for effective permeability, µef f , derived by Pendry et al. [12], which further
supports the analogy made between the principles of optical and acoustic
metamaterials.
This negative modulus arises from the displacement in centre of mass
inside each cavity of the resonators. At frequencies near the resonance, the
induced displacement becomes very large, representative of the energy ac-
cumulated over many pulse cycles. Once a significant amount of energy is
stored in the resonator relative to the driving field, the instantaneous dis-
placement of the centre of mass in the unit cell switches from in-phase to
out-of-phase with the driving field. It is this phase change that produces the
negative response from the material, and is also responsible for a measured
negative time delay observed in the ultrasonic pulses at the output of the
material [47].
It is important to note that since this displacement of the centre of mass
is instigated by compression-extensional motion, the centre of mass is sta-
tionary, therefor an effective mass dispersion in frequency is not observed.

2.2.3 Double Negativity in Acoustic Metamaterials


The effect of simultaneously negative mass density and bulk modulus in a
metamaterial is a negative acoustic index. There are numerous methods of

14
achieving double negativity in acoustic metamaterials (such as those previ-
ously discussed) that typically take advantage of locally resonant structures
to produce frequency dispersions for ρef f and κef f .

Figure 10: Soft 3D acoustic metamaterials composed of ultra-slow Mie resonators. a. Pho-
tograph of the experimental setup. Two large broadband ultrasonic transducers (emitter/receiver) with
acoustic metamaterial sample of silicone rubber microbeads in water-based gel placed between them.
Propagation distance on z axis can be varied precisely by the motorised linear stage the device is set up
on. b. Optical microscopy image of macroporous silicone rubber microbeads embedded in a water-based
gel matrix. c. Scanning electron microscope image of the core of a single silicone rubber microbead. [57]

A straightforward three-dimensional example involves a concentrated sus-


pension of macroporous silicone rubber microbeads to act as Mie resonators
in a soft 3D acoustic metamaterial, demonstrated by Brunet et al. [57]. In
this example, silicone rubber microbeads are randomly dispersed in a water-
based gel to form a 3D matrix that behaves acoustically like water (Fig.
10b). The porosity of the beads allows for periodicity between rubber sili-
cone components to be in the scale of micrometers (Fig. 10c), and the high
ratio between the speed of acoustic waves through the gel and the microbeads
permits the effective parameters, ρef f and κef f .
The microbeads act as Mie resonators, with strong monopolar resonance
that gives rise to a frequency distribution for ρef f , and dipolar resonance
giving rise to a frequency distribution for κef f [57]. Ultrasonic waves prop-

15
agating through this effective medium (Fig. 10a) then experience
q double
negativity, and as a result, a negative acoustic index (n = ρ/κ) in a broad
band of ultrasonic frequencies. This particular metafluid also has applica-
tions in transformation acoustics due to it being capable of a zero-valued
acoustic index [39, 40].

2.2.4 Strong Anisotropy in Acoustic Metamaterials


Negative index acoustic metamaterials gained a large amount of interest,
much like their optical counterparts, negative effective parameters are not
necessary for metamaterials to achieve new novel properties that can be
of great practical use. With respect to the homogenisation of an effective
medium, metamaterials with strong anisotropy not typically seen in natural
materials can be realised.

Figure 11: Theoretical example of an anisotropic metamaterial. This metamaterial is composed


of alternating layers of metal (white) and air (colour). The x-direction is parallel to the layers of metal
and the y-direction is perpendicular. The air regions beyond the metal layers are the considered the ends
of the wave guides. [16]

Anisotropy indicates that the refractive index of a material is directionally


dependent, meaning acoustic waves propagating through anisotropic mate-
rials will display the previously explained equi-frequency contours in Fig.
5 (section 1.3). Metamaterials with this property are often constructed in
similar ways to the theoretical example shown in Fig. 11, in this case using
alternating layers of metal and air with subwavelength spacing between the
layers. Local resonance is not required with this design, meaning effective
values for ρ and κ can be determined analytically as:

16
1 V 1−V
= + , (11)
ρx ρ1 ρ2

ρy = V ρ1 + (1 − V )ρ2 , (12)

1 V 1−V
= + , (13)
κ κ1 κ2
where V is the fraction of the volume filled by each layer of metal, ρ1 and
κ1 are the density and bulk modulus of the metal, ρ2 and κ2 are the density
and bulk modulus of the filling fluid between the metal layers, and ρx and ρy
are the directional density components of the effective medium [16].
It is easy to see from equations (11) and (12) that when there is a large
difference in densities of the metal and the filling fluid (for example brass
and air), a strong anisotropy is observed, resulting in a much larger effective
density in the y-direction and a flatter elliptical equi-frequency contour (Fig.
5). If one of the directional components for effective density is negative, then
the corresponding equi-frequency contour will become hyperbolic (Fig. 5),
however producing this sort of dispersion would require the use of locally
resonant structures similar to those prior discussed [16, 32].
As can be seen by this example, achieving strong anisotropy in acoustic
metamaterials does not involve heavily complicated structures and simply
exploits the naturally high ratio in density between certain metals and air.
Metamaterials of this classification are widely used for deep-subwavelength
acoustic imaging, as the flat hyperbolic and elliptical dispersions can support
arbitrarily large wave vectors that hold the sub-diffraction-limit details that
are undetectable through conventional imaging techniques [16, 17].

3 Deep-Subwavelength Acoustic Imaging


3.1 Constructing from Acoustic Metamaterials
Advancements in the field of metamaterials have led to breakthroughs in sub-
wavelength imaging for both the optics and acoustics. Many of the principles
associated with optical subwavelength imaging introduced in section 1.3 are
valid for the acoustic regime, and indeed have motivated designs for acoustic
superlenses and hyperlenses that have seen experimental realisation [17].
The constraint that this imaging faces is the decay of evanescent waves in
the far-field, and the aim of a superlens or hyperlens is to capture the infor-
mation carried by these components to form images of higher resolution than

17
previously possible through conventional methods. Similarly to the optical
case, doubly negative as well as singly negative acoustic metamaterials are
capable of amplifying evanescent waves. A two-dimensional array of elastic
membranes with negative effective density has been shown experimentally
to amplify the evanescent components of acoustic waves, demonstrating that
singly negative metamaterials are capable of super-resolution imaging [58,
59]. The experimental model by Park et al. can effectively resolve an image
of two 600 - 900 Hz sources spaced λ/17 apart, however this type of superlens
relies on surface-plasmon-like resonance to act as an interface between the
positive surrounding medium and the and the singly negative metamaterial,
the evanescent modes will decay exponentially with distance from the inter-
face, meaning the lens is only effective if its thickness is subwavelength [17].
This limitation restricts the practicality of similar singly negative acoustic
superlens designs, as they will not be effective in the higher frequency ultra-
sound regime.

Figure 12: Suplerlensing slab composed of a membrane-based negative density acoustic meta-
material a. 166 plastic squares with circular windows to hold thin membranes, forming a metamaterial
slab with lattice constant 28 mm. b. The metamaterial slab is placed between two rigid plastic plates
closed by anechoic walls to absorb reflections. The object and image lines are both 20 mm from each
interface between the metamaterial and surrounding air.

The approach of incorporating double negative acoustic metamaterials to


enhance evanescent waves has also been experimentally realised. Kaina et al.

18
[36] achieved this with an intriguing model composed of a compact hexago-
nal array of opened aluminium drinks cans. The array acted as a flat acous-
tic superlens, with the constituent units displaying Helmholtz resonator-like
propertied, generating monopolar modes and negative values for bulk mod-
ulus [47]. The walls of the cans are thin enough to act as thin membranes
[53], and as a result, the acoustic index of the effective medium can take
on negative values at locally resonant frequencies, in this case around 400
Hz, which allows imaging of sources spaced down to λ/7 apart. The waves
propagating through this metamaterial have two focal points, with a path in
the same form as in Fig. 3.
Although utilising negative acoustic index metamaterials alone is a valid
approach for subwavelength imaging, the constituent unit cells of these ma-
terials have complicated structures that require advanced fabrication tech-
niques to produce on scales comparable to the wavelengths associated with
higher frequency acoustic waves, such as ultrasonic waves which hold the
highest practical value [59]. An alternative to this method of constructing a
superlens is to exploit the anisotropy that can arise in an effective medium
[16]. With this approach, the periodic components of the metamaterial can
remain simple enough to manufacture on a scale in the order of a few millime-
ters rather than centimeters [35]. Strongly anisotropic materials also support
higher k values due to their elliptical and hyperbolic frequency dispersions
(Fig. 5), opening up the possibilities for potential acoustic hyperlens designs
[16, 22].
With locally resonant structures and anisotropic materials as building
blocks, numerous methods for resolving and magnifying subwavelength detail
with up to λ/50 resolution have been theorised and experimentally demon-
strated.

3.2 Demonstrations of Effective Acoustic Superlensing


and Hyperlensing
3.2.1 Acoustic Superlens
An acoustic superlens is designed to enhance the evanescent field compo-
nents of an acoustic wave in order to allow for their detection, and does not
necessarily magnify the wave in question. By this definition, any material
that amplifies evanescent wave is classified as a superlens. This would place
the two examples discussed in section 3.1 by Park and Kaina [59, 36] in this
distinction, as both are capable of effectively amplifying evanescent modes of
sound waves up to around 1000 Hz, however the negative parameters associ-
ated with the metamaterials that compose them are responsible for losses in

19
material [26, 16].
Without the reliance on negative effective parameters, Jia et al. [35] con-
structed a 2D acoustic superlens by employing a metamaterial consisting of
periodically arrayed layers of air and Perspex with the same structural ar-
rangement as the theoretical example in Fig. 11. Due to Perspex having
about 1000 times the density of air, the effective medium displays strong
anisotropic density, with effective density in the x-direction becoming almost
negligible. The reflection of acoustic waves between the Perspex layers also
gives rise to Fabry-Perot resonances, for which large evanescent wave com-
ponents couple strongly with [17, 34]. This allows the superlens to support
large wave vectors at a resonant frequency of 1960 Hz, and resolve an image
with λ/14 half-power beam width [35].

(a) (b)
Figure 13: Holey-structured metamaterial for deep-subwavelength acoustic imaging. a. Ex-
perimental model composed of an array of square brass alloy tubes with 0.79 mm width and 1.58 mm
spacing, fitted in parallel into a 4 inch wide square aluminium tube. b. Left: Imaging object letter ’E’
with line-width 3.18 mm (λ/50) perforated in an ultra-thin brass plate. Right: Measure image of letter
’E’ from 2.18 kHz frequency source placed behind the image, obtained at distance 1.58 mm from the out-
put plane. Red dashed line represents cross-sectional field distribution. White indicates areas of higher
pressure intensity [34].

Zhu et al. [34] expanded the concept for Jia’s superlens to three dimen-
sions by crafting a metamaterial composed of an array of square brass alloy
tubes (Fig. 13a). The use of brass over Perspex results in even stronger
anisotropy, as the density of brass is about 7000 time that of air [16]. In
addition, this particular model can exhibit effectively infinite anisotropy, due
to the cross-section of the metamaterial having the same refractive index as
air when approximated as an effective homogeneous medium. This strong
anisotropy in conjunction with Fabry-Perot resonances allows for resolution
of acoustic images with beamwidths in the order of λ/50 (Fig. 13b) [34].
Although this design for an acoustic superlens is practical for imaging sub-
wavelength details, there is still significant loss induced by the Fabry-Perot
resonances [31]. To overcome these losses, hyperlensing methods convert the

20
evanescent field components into propagating waves to offset their decay and
process the subwavelength details of an image in the far-field [22].

3.2.2 Acoustic Hyperlens


A hyperlens not only provides significant magnification of an image, but
converts evanescent waves into propagating waves [22]. An acoustic hyperlens
uses the same principles as optical hyperlenses to support high wave vector, k,
components, as it is possible for acoustic waves to display the same dispersion
patterns as EM waves (Fig. 5) [60].
There have been theoretical and experimental realisations of different
acoustic hyperlens designs. Ao and Chan [32] demonstrated that a two-
dimensional acoustic metamaterial exhibiting strong anisotropy, with a lay-
ered structure similar to the hyperlens design seen in Fig. 4, is effective at
resolving and magnifying two acoustic sources spaced about λ/4 apart.

Figure 14: Experimental magnifying imaging of a dual-source sub-diffraction-limited object


at 6.6 kHz using an acoustic hyperlens. a. Experimental realisation of the acoustic hyperlens,
made of 36 brass fins embedded on a brass substrate. The fins span 180◦ in the angular direction, running
radially from a radius of 2.7 cm to 21.8 cm. b. 3D simulation of the pressure field overlayed with the
periodic elastic fin structure. c. In red is the magnified image measured at the outer edge of the lens,
showing two distinct peaks and one distinct trough. In blue is the control experiment measured under the
same conditions without the hyperlens, showing a single broad peak.

21
An acoustic metamaterial that displays a hyperbolic frequency dispersion
for the 1 - 2.5 kHz range has been demonstrated, using a metamaterial con-
sisting of multiple arrays of clamped thin plates that act similarly to thin
membranes to generate negative values for density [61]. This realisation is
capable of subwavelength imaging, but only supports partial focusing.
A theoretical proposal for a planar acoustic hyperlens by Wu and Chen
[30] suggests a design that also exploits strong anisotropy to support high
k-modes, but then uses transformation acoustics to add magnification effects
to a planar channeling lens.
Another experimental realisation by Li et al. [31] has demonstrated acous-
tic super-resolution with 8 times magnification. This hyperlens is composed
of brass fins arranged in a fan-like structure seen in Fig. 14a, which exploits
the subwavelength spacing of the fins and the huge difference between the
densities of brass and air to achieve a significantly flat equi-frequency con-
tour. Due to the radial and axial densities of the metamaterial having the
same sign, this contour is elliptical in shape, but its high eccentricity allows
access to large k components [17].
The magnification of the image is owning to the large ratio between the
inner and outer radii, which allows the compression of a significant portion
of evanescent components into the band of propagating waves [31]. The
subwavelength objects then appear to be larger than the diffraction limit
for the associated frequency range when observed on the outer edge of the
hyperlens (Fig. 14b).
This particular hyperlens design can distinguish between acoustic sources
with spacing down to λ/7, achieving resolution not possible with the hyper-
lens removed (Fig. 14c) without the dependence on locally resonant struc-
tures [31].
Extending upon this design would be a three-dimensional structure, form-
ing a brass hemisphere with perforated holes and similar periodicity [31].
This 3D design can be generalised to use the same principles of the 2D hy-
perlens, as the velocity fields of acoustic waves are longitudinally polarised
[16].

4 Applications of Acoustic Metamaterials


Acoustic metamaterials with specific properties are generally constructed for
certain applications, with the desired use of the material usually dictating
it’s design and composition. Phononic crystals that display novel properties
that have led to uses in the development of high frequency signal processing
devices for wireless communications [4]. They have also seen applications in

22
chemical analysis, through detecting small changes in the acoustic response
of different chemicals [45].
Acoustic metamaterials that can acquire frequency dependent constituent
parameters and negative or near-zero refractive indices have been found use-
ful in transformation acoustics, leading to experimental demonstrations of
acoustic cloaking [37, 38, 39, 40] as well as methods for slowing acoustic
waves [62].
The techniques used for deep-subwavelength imaging discussed in this re-
view are part of a still developing field, so practical applications are not yet
being employed, however the advancements in the field of acoustic imaging
is significant. Industrial applications include the improvement of any sys-
tem that relies on the detection of acoustic waves, such as ultrasonic medical
imaging, underwater sonar mapping, and non-destructive materials testing
[16, 17, 31]. Currently the only way to obtain higher resolution with these
practices is to increase the amplitude of the acoustic waves in use, increasing
the energy of the pulses and potentially harming patients, or damaging ma-
terial samples. Effective superlensing or hyperlensing of these pulses could
improve the resolution of images obtained from patients or materials without
the risk of damage.

5 Conclusion
In only twenty years, the field of metamaterials has evolved from a mere
thought experiment to become highly regarded and driven by experimental
research. The acoustic branch has broken away from its direct association
with optical metamaterials to become an advanced field with its own indus-
trial applications.
Methods of subwavelength acoustic imaging presented in this review can
be considered as proofs of concept for an acoustic superlens or hyperlens, with
further research leading to a translation of a device from the experimental
stages to an industrial setting. Improving the designs to incorporate broader
ranges of operating frequencies, as well as diminishing the cumbersome size
of the metamaterials that compose them will be what aids this effort.
Advancements in the field are typically restrained by limitations of the
fabrication methods currently employed for manufacturing acoustic metama-
terials, leaving room for intuitive new developments for these methods that
could propel the field towards the desired functionalities for this classification
of metamaterials. The improvements to ultrasonic medical imaging and non-
destructive materials testing would be hugely significant and are the main
driving force in the development of deep-subwavelength acoustic imaging.

23
References
[1] P. M. Morse. Vibration and Sound. 1st ed. McGraw-Hill Book Com-
pany, Inc., 1936.
[2] P. M. Morse, K. U. Ingard, and R. T. Beyer. Theoretical Acoustics.
1st ed. Princeton University Press, 1986.
[3] Xiangdong Zhang and Zhengyou Liu. “Negative refraction of acous-
tic waves in two-dimensional phononic crystals”. In: Applied Physics
Letters 85.2 (2004), pp. 341–343. doi: 10.1063/1.1772854.
[4] Abdelkrim Khelif and Ali Adibi. Phononic Crystals: Fundamentals and
Applications. 1st ed. Springer, 2016, pp. 139, 240, 241.
[5] Viktor G Veselago. “The Electrodynamics of Substances with Simulta-
neously Negative Values of  and µ”. In: Soviet Physics Uspekhi 10.4
(1968), p. 509.
[6] Evan J. Reed, Marin Solja, and John D. Joannopoulos. “Reversed
Doppler Effect in Photonic Crystals”. In: Phys. Rev. Lett. 91 (13 Sept.
2003), p. 133901. doi: 10.1103/PhysRevLett.91.133901.
[7] Chiyan Luo et al. “Cerenkov Radiation in Photonic Crystals”. In: Sci-
ence 299.5605 (2003), pp. 368–371. doi: 10.1126/science.1079549.
[8] J. B. Pendry. “Negative Refraction Makes a Perfect Lens”. In: Phys.
Rev. Lett. 85 (18 Oct. 2000), pp. 3966–3969. doi: 10.1103/PhysRevLett.
85.3966.
[9] D. R. Smith, J. B. Pendry, and M. C. K. Wiltshire. “Metamaterials
and Negative Refractive Index”. In: Science 305.5685 (2004), pp. 788–
792. issn: 0036-8075. doi: 10.1126/science.1096796.
[10] D. R. Smith et al. “Composite Medium with Simultaneously Negative
Permeability and Permittivity”. In: Phys. Rev. Lett. 84 (18 May 2000),
pp. 4184–4187. doi: 10.1103/PhysRevLett.84.4184.
[11] J. B. Pendry et al. “Extremely Low Frequency Plasmons in Metallic
Mesostructures”. In: Phys. Rev. Lett. 76 (25 June 1996), pp. 4773–4776.
doi: 10.1103/PhysRevLett.76.4773.
[12] J. B. Pendry et al. “Magnetism from conductors and enhanced non-
linear phenomena”. In: IEEE Transactions on Microwave Theory and
Techniques 47.11 (Nov. 1999), pp. 2075–2084. issn: 0018-9480. doi:
10.1109/22.798002.

24
[13] Atsushi Ishikawa et al. “Deep Subwavelength Terahertz Waveguides
Using Gap Magnetic Plasmon”. In: Phys. Rev. Lett. 102 (4 Jan. 2009),
p. 043904. doi: 10.1103/PhysRevLett.102.043904.
[14] D. Schurig et al. “Metamaterial Electromagnetic Cloak at Microwave
Frequencies”. In: Science 314.5801 (2006), pp. 977–980. issn: 0036-
8075. doi: 10.1126/science.1133628.
[15] Nicholas Fang et al. “Sub-Diffraction-Limited Optical Imaging with
a Silver Superlens”. In: Science 308.5721 (2005), pp. 534–537. issn:
0036-8075. doi: 10.1126/science.1108759.
[16] Lee Ren Fok. “Anisotropic and Negative Acoustic Index Metamateri-
als”. In: UC Berkeley Electronic Theses and Dissertations (2010). url:
https://escholarship.org/uc/item/21h3j5xr.
[17] Guancong Ma and Ping Sheng. “Acoustic metamaterials: From local
resonances to broad horizons”. In: Science Advances 2.2 (2016). doi:
10.1126/sciadv.1501595.
[18] C. T. Chan, Jensen Li, and K. H. Fung. “On extending the con-
cept of double negativity to acoustic waves”. In: Journal of Zhejiang
University-SCIENCE A 7.1 (Jan. 2006), pp. 24–28. issn: 1862-1775.
doi: 10.1631/jzus.2006.A0024.
[19] L.E. Kinsler et al. Fundamentals of Acoustics. 4th ed. John Wiley &
Sons, Dec. 1999, pp. 38, 19.
[20] J B Pendry and S A Ramakrishna. “Focusing light using negative
refraction”. In: Journal of Physics: Condensed Matter 15.37 (2003),
p. 6345. url: http://stacks.iop.org/0953-8984/15/i=37/a=004.
[21] Eric Betzig and Jay K. Trautman. “Near-Field Optics: Microscopy,
Spectroscopy, and Surface Modification Beyond the Diffraction Limit”.
In: Science 257.5067 (1992), pp. 189–195. issn: 0036-8075. doi: 10.
1126/science.257.5067.189.
[22] Zubin Jacob, Leonid V. Alekseyev, and Evgenii Narimanov. “Optical
Hyperlens: Far-field imaging beyond the diffraction limit”. In: Opt.
Express 14.18 (Sept. 2006), pp. 8247–8256. doi: 10 . 1364 / OE . 14 .
008247.
[23] Igor I. Smolyaninov, Yu-Ju Hung, and Christopher C. Davis. “Magni-
fying Superlens in the Visible Frequency Range”. In: Science 315.5819
(2007), pp. 1699–1701. issn: 0036-8075. doi: 10.1126/science.1138746.

25
[24] Anthony Grbic and George V. Eleftheriades. “Overcoming the Diffrac-
tion Limit with a Planar Left-Handed Transmission-Line Lens”. In:
Phys. Rev. Lett. 92 (11 Mar. 2004), p. 117403. doi: 10.1103/PhysRevLett.
92.117403.
[25] S Anantha Ramakrishna. “Physics of negative refractive index mate-
rials”. In: Reports on Progress in Physics 68.2 (2005), p. 449. url:
http://stacks.iop.org/0034-4885/68/i=2/a=R06.
[26] David R. Smith et al. “Limitations on subdiffraction imaging with
a negative refractive index slab”. In: Applied Physics Letters 82.10
(2003), pp. 1506–1508. doi: 10.1063/1.1554779.
[27] Zhaowei Liu et al. “Far-Field Optical Hyperlens Magnifying Sub-Diffraction-
Limited Objects”. In: Science 315.5819 (2007), pp. 1686–1686. issn:
0036-8075. doi: 10.1126/science.1137368.
[28] Clifford Truesdell, Walter Noll, and Stuart S. Antman. The Non-Linear
Field Theories of Mechanics. 3rd ed. Springer, 2003.
[29] Dylan Lu and Zhaowei Liu. “Hyperlenses and metalenses for far-field
super-resolution imaging”. In: Nature Communications 3 (Nov. 2012).
Review Article, p. 1205. doi: 10.1038/ncomms2176.
[30] Liang-Yu Wu and Lien-Wen Chen. “Acoustic planar hyperlens via
transformation acoustics”. In: Journal of Physics D: Applied Physics
44.12 (2011), p. 125402.
[31] Jensen Li et al. “Experimental demonstration of an acoustic magnifying
hyperlens”. In: Nature Materials 8 (Oct. 2009), p. 931. doi: 10.1038/
nmat2561.
[32] Xianyu Ao and C. T. Chan. “Far-field image magnification for acoustic
waves using anisotropic acoustic metamaterials”. In: Phys. Rev. E 77
(2 Feb. 2008), p. 025601. doi: 10.1103/PhysRevE.77.025601.
[33] Sébastien Guenneau et al. “Acoustic metamaterials for sound focusing
and confinement”. In: New Journal of Physics 9.11 (2007), p. 399. url:
http://stacks.iop.org/1367-2630/9/i=11/a=399.
[34] J. Zhu et al. “A holey-structured metamaterial for acoustic deep-subwavelength
imaging”. In: Nature Physics 7 (Nov. 2010), p. 52. doi: 10 . 1038 /
nphys1804.
[35] Han Jia et al. “Subwavelength imaging by a simple planar acoustic
superlens”. In: Applied Physics Letters 97.17 (2010), p. 173507. doi:
10.1063/1.3507893.

26
[36] Nadège Kaina et al. “Negative refractive index and acoustic superlens
from multiple scattering in single negative metamaterials”. In: Nature
525 (Sept. 2015), p. 77. doi: 10.1038/nature14678.
[37] Lucian Zigoneanu, Bogdan-Ioan Popa, and Steven A. Cummer. “Three-
dimensional broadband omnidirectional acoustic ground cloak”. In: Na-
ture Materials 13 (Mar. 2014), p. 352. doi: 10.1038/nmat3901.
[38] Shu Zhang, Chunguang Xia, and Nicholas Fang. “Broadband Acoustic
Cloak for Ultrasound Waves”. In: Phys. Rev. Lett. 106 (2 Jan. 2011),
p. 024301. doi: 10.1103/PhysRevLett.106.024301.
[39] Huanyang Chen and C T Chan. “Acoustic cloaking and transformation
acoustics”. In: Journal of Physics D: Applied Physics 43.11 (2010),
p. 113001. url: http://stacks.iop.org/0022-3727/43/i=11/a=
113001.
[40] Huanyang Chen and C. T. Chan. “Acoustic cloaking in three dimen-
sions using acoustic metamaterials”. In: Applied Physics Letters 91.18
(2007), p. 183518. doi: 10.1063/1.2803315.
[41] Zhengyou Liu et al. “Locally Resonant Sonic Materials”. In: Science
289.5485 (2000), pp. 1734–1736. issn: 0036-8075. doi: 10.1126/science.
289.5485.1734.
[42] M. Sigalas and E.N. Economou. “Band structure of elastic waves in
two dimensional systems”. In: Solid State Communications 86.3 (1993),
pp. 141–143. issn: 0038-1098. doi: https://doi.org/10.1016/0038-
1098(93)90888-T.
[43] M. S. Kushwaha et al. “Acoustic band structure of periodic elastic
composites”. In: Phys. Rev. Lett. 71 (13 Sept. 1993), pp. 2022–2025.
doi: 10.1103/PhysRevLett.71.2022.
[44] F. R. Montero de Espinosa, E. Jiménez, and M. Torres. “Ultrasonic
Band Gap in a Periodic Two-Dimensional Composite”. In: Phys. Rev.
Lett. 80 (6 Feb. 1998), pp. 1208–1211. doi: 10.1103/PhysRevLett.
80.1208.
[45] Ming-Hui Lu, Liang Feng, and Yan-Feng Chen. “Phononic crystals and
acoustic metamaterials”. In: Materials Today 12.12 (2009), pp. 34–42.
issn: 1369-7021. doi: 10.1016/S1369-7021(09)70315-3.
[46] Yiqun Ding et al. “Metamaterial with Simultaneously Negative Bulk
Modulus and Mass Density”. In: Phys. Rev. Lett. 99 (9 Aug. 2007),
p. 093904. doi: 10.1103/PhysRevLett.99.093904.

27
[47] Nicholas Fang et al. “Ultrasonic metamaterials with negative modu-
lus”. In: Nature Materials 5 (Apr. 2006), p. 452.
[48] L. Fok and X. Zhang. “Negative acoustic index metamaterial”. In:
Phys. Rev. B 83 (21 June 2011), p. 214304. doi: 10.1103/PhysRevB.
83.214304.
[49] Y. Cheng, J. Y. Xu, and X. J. Liu. “One-dimensional structured ul-
trasonic metamaterials with simultaneously negative dynamic density
and modulus”. In: Phys. Rev. B 77 (4 Jan. 2008), p. 045134. doi:
10.1103/PhysRevB.77.045134.
[50] Sam Hyeon Lee et al. “Composite Acoustic Medium with Simultane-
ously Negative Density and Modulus”. In: Phys. Rev. Lett. 104 (5 Feb.
2010), p. 054301. doi: 10.1103/PhysRevLett.104.054301.
[51] Sam Hyeon Lee et al. “Acoustic metamaterial with negative density”.
In: Physics Letters A 373.48 (2009), pp. 4464–4469. issn: 0375-9601.
doi: 10.1016/j.physleta.2009.10.013.
[52] Sam Hyeon Lee et al. “Acoustic metamaterial with negative modulus”.
In: Journal of Physics: Condensed Matter 21.17 (2009), p. 175704. url:
http://stacks.iop.org/0953-8984/21/i=17/a=175704.
[53] Min Yang et al. “Coupled Membranes with Doubly Negative Mass
Density and Bulk Modulus”. In: Phys. Rev. Lett. 110 (13 Mar. 2013),
p. 134301. doi: 10.1103/PhysRevLett.110.134301.
[54] Min Yang et al. “Homogenization scheme for acoustic metamaterials”.
In: Phys. Rev. B 89 (6 Feb. 2014), p. 064309. doi: 10.1103/PhysRevB.
89.064309.
[55] Ping Sheng et al. “Dynamic mass density and acoustic metamaterials”.
In: Physica B: Condensed Matter 394.2 (2007), pp. 256–261. issn: 0921-
4526. doi: 10.1016/j.physb.2006.12.046.
[56] H.H. Huang, C.T. Sun, and G.L. Huang. “On the negative effective
mass density in acoustic metamaterials”. In: International Journal of
Engineering Science 47.4 (2009), pp. 610–617. issn: 0020-7225. doi:
10.1016/j.ijengsci.2008.12.007.
[57] Thomas Brunet et al. “Soft 3D acoustic metamaterial with negative
index”. In: Nature Materials 14 (Dec. 2014), p. 384. doi: 10.1038/
nmat4164.
[58] Choon Mahn Park et al. “Amplification of Acoustic Evanescent Waves
Using Metamaterial Slabs”. In: Phys. Rev. Lett. 107 (19 Nov. 2011),
p. 194301. doi: 10.1103/PhysRevLett.107.194301.

28
[59] Jong Jin Park et al. “Acoustic superlens using membrane-based meta-
materials”. In: Applied Physics Letters 106.5 (2015), p. 051901. doi:
10.1063/1.4907634.
[60] Johan Christensen and F. Javier Garcıa de Abajo. “Anisotropic Meta-
materials for Full Control of Acoustic Waves”. In: Phys. Rev. Lett. 108
(12 Mar. 2012), p. 124301. doi: 10.1103/PhysRevLett.108.124301.
[61] Chen Shen et al. “Broadband Acoustic Hyperbolic Metamaterial”. In:
Phys. Rev. Lett. 115 (25 Dec. 2015), p. 254301. doi: 10.1103/PhysRevLett.
115.254301.
[62] Xuefeng Zhu et al. “Implementation of dispersion-free slow acoustic
wave propagation and phase engineering with helical-structured meta-
materials”. In: Nature Communications 7 (May 2016), p. 11731. doi:
10.1038/ncomms11731.

29

You might also like