+ All Categories
Home > Documents > on June 30, 2017 Non-reciprocalwave rspa ...4 rspa.royalsocietypublishing.org Proc.R.Soc.A...

on June 30, 2017 Non-reciprocalwave rspa ...4 rspa.royalsocietypublishing.org Proc.R.Soc.A...

Date post: 27-Jan-2021
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
22
rspa.royalsocietypublishing.org Research Cite this article: Nassar H, Chen H, Norris AN, Haberman MR, Huang GL. 2017 Non-reciprocal wave propagation in modulated elastic metamaterials. Proc. R. Soc. A 473: 20170188. http://dx.doi.org/10.1098/rspa.2017.0188 Received: 16 March 2017 Accepted: 1 June 2017 Subject Areas: mechanical engineering Keywords: non-reciprocity, pump wave, one-way transition, unidirectional bandgap Author for correspondence: G. L. Huang e-mail: [email protected] Non-reciprocal wave propagation in modulated elastic metamaterials H. Nassar 1 , H. Chen 1 , A. N. Norris 2 , M. R. Haberman 3 and G. L. Huang 1 1 Department of Mechanical and Aerospace Engineering, University of Missouri, Columbia, MO 65211, USA 2 Department of Mechanical and Aerospace Engineering, Rutgers University, Piscataway, NJ 08854-8058, USA 3 Department of Mechanical Engineering and Applied Research Laboratories, The University of Texas at Austin, Austin, TX 78712, USA HN, 0000-0001-7235-4498; GLH, 0000-0002-6238-6127 Time-reversal symmetry for elastic wave propagation breaks down in a resonant mass-in-mass lattice whose inner-stiffness is weakly modulated in space and in time in a wave-like fashion. Specifically, one-way wave transmission, conversion and amplification as well as unidirectional wave blocking are demonstrated analytically through an asymptotic analysis based on coupled mode theory and numerically thanks to a series of simulations in harmonic and transient regimes. High-amplitude modulations are then explored in the homogenization limit where a non-standard effective mass operator is recovered and shown to take negative values over unusually large frequency bands. These modulated metamaterials, which exhibit either non-reciprocal behaviours or non-standard effective mass operators, offer promise for applications in the field of elastic wave control in general and in one-way conversion/amplification in particular. 1. Introduction Resonant spring-mass lattices have been studied for at least a century now since they were first introduced by Lord Kelvin in the late 1800s in an attempt to explain dispersion for optical waves [1]. It is not until recently, however, that they have been implemented as subwavelength compact mechanical filters by Liu et al. [2] in a mass-in-mass design using 2017 The Author(s) Published by the Royal Society. All rights reserved. on June 30, 2017 http://rspa.royalsocietypublishing.org/ Downloaded from
Transcript
  • rspa.royalsocietypublishing.org

    ResearchCite this article: Nassar H, Chen H, Norris AN,Haberman MR, Huang GL. 2017 Non-reciprocalwave propagation in modulated elasticmetamaterials. Proc. R. Soc. A 473: 20170188.http://dx.doi.org/10.1098/rspa.2017.0188

    Received: 16 March 2017Accepted: 1 June 2017

    Subject Areas:mechanical engineering

    Keywords:non-reciprocity, pump wave, one-waytransition, unidirectional bandgap

    Author for correspondence:G. L. Huange-mail: [email protected]

    Non-reciprocal wavepropagation in modulatedelastic metamaterialsH. Nassar1, H. Chen1, A. N. Norris2, M. R. Haberman3

    and G. L. Huang1

    1Department of Mechanical and Aerospace Engineering, Universityof Missouri, Columbia, MO 65211, USA2Department of Mechanical and Aerospace Engineering, RutgersUniversity, Piscataway, NJ 08854-8058, USA3Department of Mechanical Engineering and Applied ResearchLaboratories, The University of Texas at Austin, Austin, TX 78712, USA

    HN, 0000-0001-7235-4498; GLH, 0000-0002-6238-6127

    Time-reversal symmetry for elastic wave propagationbreaks down in a resonant mass-in-mass lattice whoseinner-stiffness is weakly modulated in space and intime in a wave-like fashion. Specifically, one-waywave transmission, conversion and amplificationas well as unidirectional wave blocking aredemonstrated analytically through an asymptoticanalysis based on coupled mode theory andnumerically thanks to a series of simulations inharmonic and transient regimes. High-amplitudemodulations are then explored in the homogenizationlimit where a non-standard effective mass operatoris recovered and shown to take negative valuesover unusually large frequency bands. Thesemodulated metamaterials, which exhibit eithernon-reciprocal behaviours or non-standard effectivemass operators, offer promise for applications in thefield of elastic wave control in general and in one-wayconversion/amplification in particular.

    1. IntroductionResonant spring-mass lattices have been studied for atleast a century now since they were first introducedby Lord Kelvin in the late 1800s in an attemptto explain dispersion for optical waves [1]. It isnot until recently, however, that they have beenimplemented as subwavelength compact mechanicalfilters by Liu et al. [2] in a mass-in-mass design using

    2017 The Author(s) Published by the Royal Society. All rights reserved.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://crossmark.crossref.org/dialog/?doi=10.1098/rspa.2017.0188&domain=pdf&date_stamp=2017-06-28mailto:[email protected]://orcid.org/0000-0001-7235-4498http://orcid.org/0000-0002-6238-6127http://rspa.royalsocietypublishing.org/

  • 2

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    silicon-rubber-coated lead spheres stacked in an epoxy matrix. The renewed recognition of localresonances as a powerful mechanism for the generation of bandgaps has effectively opened up anew major thrust in acoustic and elastic metamaterials research (e.g. [3–8]).

    In particular, metamaterials with broken time-reversal symmetry and reciprocity constitutea growing area of interest as they have proved vital for achieving non-reciprocal wave control.In a medium where time-reversal symmetry is not observed, identical wave forms incident inopposite directions can be transformed differently. This phenomenon offers new possibilitiesin selective wave control as it increases the set of variables based on which waves can besegregated: their frequency, polarization, direction and, thanks to non-reciprocity, the sign of theirgroup velocity. Examples of applications include the realization of directional mode converters[9–11], selective acoustic circulators [12], directional wave reflectors [13,14] and directional waveaccelerators [15].

    A lossy material, where energy is irreversibly dissipated, is an example of a system withouttime-reversal symmetry. More interestingly, media including strongly nonlinear components orfeaturing specific chiral microstructures can have a diode effect on acoustic wave propagationeither in the bulk of the medium or along its edges [16–19]. These configurations can be qualifiedas intrinsic since they do not feature any external stimuli. By contrast, in extrinsic configurations,the breaking of time-reversal symmetry is the consequence of an externally applied loading eitherin the form of a biasing distribution of angular momenta [12,20–23] or in the form of a ‘pumpwave’. In the latter case, a pump wave is the name given to a wave-like space–time modulationof the constitutive properties of a host medium. By changing, say, the elastic stiffness of themedium in space and in time, a bias breaking time-reversal symmetry can be introduced. Thescattering of travelling elastic waves by a pump wave has been examined in connection withparametric amplification in the previous century by many authors [24–29] and subsequently inthe context of ‘dynamic materials’ by Lurie [30] and co-workers (see also [31] and referencestherein). Finally, evidence for one-way mode transitions and one-directional bandgaps in weaklymodulated waveguides were advanced by Yu & Fan [9] and Zanjani et al. [10,11] on one hand andby Swinteck et al. [13] and Trainiti & Ruzzene [14] on the other hand. Here, we build on their workand show how weakly modulated resonant metamaterials offer both possibilities simultaneously.

    The present paper aims to characterize wave propagation in a resonant metamaterial where theinner stiffness is weakly modulated in a wave-like fashion. It is demonstrated both analyticallyand numerically that the dispersive two-branch nature of the band diagram of the metamaterialcombined with the loss of time-reversal symmetry offers a rich wave control potential includingselective one-way conversion/transmission and reflection. Note that as suggested by Casadeiet al. [32] and Chen et al. [33,34], modulating the inner stiffness of a metamaterial is technologicallyfeasible using programmable piezoelectric components. Other techniques for guiding pumpwaves can make use of shock waves in soft materials [35], of magnetic fields soliciting magneto-rheological elastomers [36] or of the photo-elastic effect [13,37].

    Furthermore, beyond symmetry breaking, we explore a set of strong space–time modulationsthat turn out to have a significant impact on the behaviour of the effective mass operator inthe homogenization limit. As a matter of fact, in addition to the main bandgap of the inner-resonant metamaterial, a set of secondary bandgaps appear at low frequencies. Most importantly,modulation-induced bandgaps are proved to be able to merge together into broader gaps incontrast to what is systematically observed at low frequencies in non-modulated metamaterials(e.g. [4]).

    The rest of the paper goes as follows. Section 2 is a brief reformulation of known resultson non-modulated mass-in-mass metamaterials. In §3, we derive the first-order asymptoticsfor the dispersion curve of a weakly modulated metamaterial using coupled mode theory. Itis shown that a pump wave couples various free waves propagating through the metamaterialas soon as they satisfy a phase matching condition. Depending on the sign of the involvedgroup velocities, the coupling will be termed ‘veering’ (co-directional coupling) leading toone-way mode conversion or will be termed ‘locking’ (contra-directional coupling) leading toone-directional bandgaps. A quantitative and qualitative analysis of these phenomena based on

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 3

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    M

    K

    k

    un

    vn

    m

    Figure 1. A discrete spring mass-in-mass metamaterial.

    our analytical findings and a series of numerical experiments is presented. Section 4 is a limit casestudy describing the behaviour of the metamaterial under strong extremely short-wavelengthmodulations. Finally, concluding remarks are presented in §5.

    2. Wave propagation in non-modulated elastic metamaterialsWe start by characterizing free and forced wave propagation in non-modulated inner-resonantmetamaterials. The results will serve as a theoretical basis for the next section where we deal withmodulated metamaterials.

    (a) Discrete and continuous modelling of a metamaterialA well-known model for inner-resonant metamaterials is the discrete spring mass-in-mass latticedepicted in figure 1 (see [1], for a brief historical account). The outer rigid shells have mass Mand are connected by linear springs of constant K. The inner resonators have mass m and areconnected to outer masses through springs of constant k. The motion equations read

    M∂2t un = −k(un − vn) + K(un+1 + un−1 − 2un)and m∂2t vn = −k(vn − un),

    ⎫⎬⎭ (2.1)

    where un ≡ un(t) and vn ≡ vn(t) denote the displacement of the nth outer and inner masses,respectively, and ∂2t indicates differentiating with respect to time t, twice. When the frequencyω of the propagating waves is significantly smaller than the phononic cut-off frequency 2

    √K/M,

    a continuous counterpart to the above discrete motion equations exists and can be written as

    M∂2t u = −k(u − v) + E∂2x uand m∂2t v = −k(v − u),

    ⎫⎬⎭ (2.2)

    where ∂x is the space derivative. The displacements u and v are now functions of the continuousspace variable x and of time t. The coefficient E = K�2 can be identified as the elastic stiffness ofthe master structure, � being the spacing between the centres of two consecutive outer masses.

    Unlike earlier studies such as the one by Brillouin [1], recent papers investigate the propertiesof the considered metamaterial through the mean of an effective mass density operator firstintroduced by Auriault & Bonnet [38] in the more general context of stiff-matrix soft-inclusionselastodynamics. Here, since k will be subject to space–time modulations, the effective massdensity cannot be as easily defined and the present multi-displacement approach is preferred(e.g. [39]).

    It is also helpful to combine the motion equations into a single governing equation usingmatrices. Thus, let

    M =[

    M 00 m

    ], E =

    [E 00 0

    ], k =

    [k −k

    −k k

    ]and ψ =

    [uv

    ]. (2.3)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 4

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    Given these matrix notations, the governing equations can be grouped into

    M∂2t ψ = −kψ + E∂2xψ . (2.4)

    (b) Free and forced wave propagationPlane harmonic waves

    ψ(x, t) =Ψ ei(qx−ωt) (2.5)

    of wavenumber q, frequency ω and amplitude Ψ can travel through a homogeneous metamaterialupon satisfying

    (−ω2M + k + q2E)Ψ = 0. (2.6)

    This eigenvalue problem admits non-trivial solutions when the dispersion relation det (−ω2M +k + q2E) = 0 is satisfied, an equivalent form of which is

    D(ω, q) ≡ Eq2 − ω2(

    M + m Ω2

    Ω2 − ω2)

    = 0, (2.7)

    with Ω = √k/m being the resonance frequency of the inner masses.A solution (q,ω) to equation (2.7) is hereafter referred to as an eigenmode. The corresponding

    amplitude is denoted Ψq,ω and is called an eigenvector. Also, we will call eigenstate thecombination of an eigenmode and an eigenvector, as well as the corresponding free wave.The equation of motion (2.2)2 for the resonators implies that the eigenvectors can be defined as

    Ψq,ω =Ψω =[

    UV

    ]=

    [k − ω2m

    k

    ]. (2.8)

    In conclusion, a free wave propagating through a homogeneous metamaterial is a superpositionof plane harmonic waves Ψω ei(qx−ωt), where (q,ω) is an eigenmode solution to equation (2.7).

    Now, it is assumed that the metamaterial is subjected to a harmonic plane wave body forceof wavenumber q, frequency ω and amplitude F. The motion equation becomes

    M∂2t ψ = −kψ + E∂2xψ + F ei(qx−ωt). (2.9)When (q,ω) is not an eigenmode, F will always drive ψ into bounded oscillations of the samefrequency ω and wavenumber q and of amplitude Ψ = (−ω2M + k + q2E)−1F. By contrast, when(q,ω) is an eigenmode, F will cause ψ to have oscillations of increasing and unbounded amplitudeunless F was orthogonal to Ψω. In general,

    (−ω2M + k + q2E)Ψ = F (2.10)necessarily implies the secular condition

    Ψω · F =Ψω · (−ω2M + k + q2E)Ψ = 0, (2.11)thanks to the symmetry of M, E and k. This orthogonality property will provide essentialinformation on wave dispersion in modulated metamaterials.

    3. Wave propagation in modulated elastic metamaterialsWave-like modulations of the metamaterial are first introduced. A perturbative approach basedon the amplitude of the modulation then offers an efficient way to understand, to leadingorder, how an incident wave is transformed, converted and transmitted by the modulation. Theconstructed wave solutions demonstrate non-reciprocal directional behaviour by which wavestravelling in one direction are either blocked or converted in a different manner than wavestravelling in the opposite direction.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 5

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    k~

    (x, t) = k + 2dk cos (qmx – wmt)

    wm/qmk~

    (t)

    k~

    (t + Dt)

    Figure 2. An illustration of a modulated metamaterial. The sinusoidal modulation travels at a uniform speedωm/qm. (Onlineversion in colour.)

    (a) Wave-like modulationsUsing programmable piezoelectric active components for instance [32–34], it is possible to controlthe value of the coupling stiffness k and to modulate it in space and in time in a wave-like periodicfashion (figure 2). Thus, let

    k̃ = k̃(x, t) = k̃(qmx − ωmt) (3.1)be the modulated stiffness and be a 2π -periodic function of the phase qmx − ωmt. In what follows,we assume that the difference between the modulated stiffness k̃ and its average value k remainssmall. Furthermore, the profile of k̃ is taken to be sinusoidal. In summary, the modulation takesthe form of a harmonic plane wave of wavenumber qm and frequency ωm and has the expression

    k̃(x, t) = k + 2δk(x, t) = k + 2δk cos(qmx − ωmt), (3.2)where δk is small compared with k.

    This choice is motivated by two facts. First, studying low-amplitude modulations allowsto understand the continuous transition between a non-modulated (δk = 0) and a modulated(0 �= δk � k) metamaterial, and consequently how non-reciprocity gradually starts manifesting.Second, weakly modulated media will prove to exhibit a directional wave-conversion potentialthat has no counterpart in strongly modulated media thus extending the results obtained by Yu &Fan [9] and by Zanjani et al. [10,11] to inner-resonant metamaterials.

    Calling ψ̃ the displacement field in the modulated metamaterial, the motion equation becomes

    M∂2t ψ̃ = −k̃ψ̃ + E∂2x ψ̃ , (3.3)with

    k̃(x, t) = k + 2δk(x, t) = k + 2δk cos(qmx − ωmt). (3.4)As all coefficients are either constant or periodic, ψ̃ decomposes into a superposition of Floquet–Bloch waves

    ψ̃(x, t) = Ψ̃ (qmx − ωmt) ei( ˜qx−ω̃t) (3.5)of wavenumber q̃ and frequency ω̃, where Ψ̃ is 2π -periodic just like the modulation.

    (b) Perturbation methodTaking advantage of the smallness of δk, we look for solutions in the perturbed form

    Ψ̃ =Ψ + δΨ + · · · , q̃ = q + δq + · · · and ω̃=ω + δω + · · · (3.6)where δΨ , δq and δω are first-order corrections to Ψ , q and ω, respectively, and are all of the sameorder of magnitude as the modulation amplitude δk. Our purpose is to calculate the amplitudeΨ̃ to leading order, i.e. Ψ , and to determine how the first-order corrections (δq, δω) modify thedispersion curve.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 6

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    Let ξ be the non-dimensional phase variable

    ξ = qmx − ωmt. (3.7)Injecting (3.6) and (3.5) into (3.3) yields to leading order

    M(−iω − ωm∂ξ )2Ψ = −kΨ + E(iq + qm∂ξ )2Ψ . (3.8)Recalling that Ψ is a 2π -periodic function of ξ , it can be expanded into a Fourier series

    Ψ (ξ ) =∑j∈Z

    Ψ j eijξ , (3.9)

    where the constant amplitudes Ψ j are solutions to the algebraic eigenvalue problem

    [−(ω + jωm)2M + k + (q + jqm)2E]Ψ j = 0. (3.10)Now, for Ψ to exist and be non-zero, one of the Fourier coefficients Ψ j must exist and also be non-zero. Accordingly, there exists at least one integer j such that (q + jqm,ω + jωm) is an eigenmodesolution to equation (2.7). Call J the set of all such indices. By a harmless phase change andwithout loss of generality, it is possible to ensure that 0 ∈ J. In other words, (q,ω) is an eigenmodeand we can write

    Ψ (ξ ) =Ψ 0 +∑

    j∈J−{0}Ψ j eijξ (3.11)

    where Ψ j are eigenvectors associated with the eigenmodes (q + jqm,ω + jωm), for j ∈ J. Finally,with the notations of the previous section, we recover

    Ψ (ξ ) = V0Ψω +∑

    j∈J−{0}VjΨωj e

    ijξ , (3.12)

    with ωj =ω + jωm. For later use, we similarly define qj = q + jqm.Physically speaking, Ψω is interpreted as an incident wave and the Ψωj are seen as scattered,

    or reflected waves generated by the modulation. Two eigenstates (q,ω) and (q′,ω′) are thereforecoupled, i.e. can be identified as a pair of incident and scattered waves, whenever there existsan integer j such that the phase matching condition

    q′ − q = jqm and ω′ − ω= jωm (3.13)is satisfied. Subsequent calculations will in fact show that j must be equal to ±1 for the couplingto take place. In general, when the pump wave has its first N Fourier components non-zero, j cantake integer values between −N and N (see equation (3.29)).

    A graphical illustration can help identifying pairs of coupled modes (figure 3). Starting withthe dispersion curve of the non-modulated metamaterial C = C0, we plot the translated curvesCj = C − (jqm, jωm), j = ±1. Points of no-intersection of C correspond to cases where (q,ω) is theonly eigenmode among the (qj,ωj) so that

    Ψ (ξ ) = V0Ψω (3.14)is composed of a single uncoupled mode: the incident wave. Here, the metamaterial appearstransparent and no scattered waves are generated. Points of intersection C ∩ Cj on the other handcorrespond to cases where both (q,ω) and (qj,ωj) are eigenmodes. Hence,

    Ψ (ξ ) = V0Ψω + VjΨωj eijξ (3.15)is the sum of a pair of coupled modes, one incident and one scattered. On the example of figure 3,there appears to be three such intersection pairs labelled as A, B and C: when one ‘leg’ of a givenpair is incident, the other ‘leg’ is scattered and vice versa. Last, we exclude rare cases where morethan two curves intersect at a single point.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 7

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    3.0

    2.5

    2.0

    1.5

    1.0

    0.5

    0–3 –2 –1 0

    q/Q1 2 2

    w/W

    (A)

    (B)

    (C)

    0 1 –1

    Figure 3. Coupling and non-reciprocity: the pairs of coupled modes A, B and C are pinned. Notice that they are non-symmetrically distributed, a symptom of the breaking of time-reversal symmetry. The normalizing wavenumber is Q=√k/E = √k/K/�. For this illustration, we usedm/M= 1, qm/Q= 1.8 andωm/Ω = 0.4. (Online version in colour.)

    Note that non-reciprocity is already at play here since, generally speaking, when a pairof eigenstates (q,ω) and (qj,ωj) are coupled, their counterparts propagating in the oppositedirections, i.e. (−q,ω) and (−qj,ωj) are not coupled. In other words, if the eigenstate (q,ω) isscattered if incident in one direction, it will not be scattered when incident in the oppositedirection. Graphically, this corresponds to the loss of parity in the plot of figure 3.

    (c) Uncoupled modesLet us assume first that (q,ω) is a non-intersection point. Thus, Ψ takes the form (3.14), and thecoefficient V0 can be chosen arbitrarily by linearity. Regarding the determination of δq and δω,combining again (3.3), (3.5) and (3.6) and keeping first-order terms entail

    [M(iω + ωm∂ξ )2 − E(iq + qm∂ξ )2 + k]δΨ = (2ωδωM − 2δk − 2qδqE)Ψ . (3.16)Then, δΨ propagates through the non-modulated medium under the influence of a modulation-induced body force

    F = (2ωδωM − 2qδqE)Ψ − 2δkΨ . (3.17)

    Averaging with respect to ξ over a unit cell [0, 2π ] simplifies the above equation into

    (−ω2M + q2E + k) 〈δΨ 〉 = (2ωδωM − 2qδqE)Ψ , (3.18)where we have used the fact that δΨ is periodic and that k is the average of k̃ so that 〈δk〉 = 0.Therein, 〈〉 ≡ (1/2π ) ∫2π0 is the averaging operator. As (q,ω) is an eigenmode, the orthogonalitycondition

    Ψω · (2ωδωM − 2qδqE)Ψ = 0 (3.19)

    holds. As a consequence, by using (3.14), we obtain

    Ψω · (2ωδωM − 2qδqE)Ψω = 0. (3.20)This condition is equivalent to a variation of the dispersion relation (2.7),

    ∂D∂ωδω + ∂D

    ∂qδq = 0 ⇔ δω= cδq, (3.21)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 8

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    where c is the group velocity,

    c(ω, q) = − ∂D∂q

    /∂D∂ω

    = qE(Ω2 − ω2)2

    ω(M(Ω2 − ω2)2 + mΩ4) . (3.22)

    In other words, the frequency and wavenumber remain on the non-modulated dispersion curveC0 of the non-modulated medium for any permissible change δω, δq. The group speed relation alsoimplies that δω vanishes whenever δq does. This means that an incident wave of a given frequencyω propagates with the same wavenumber q̃ = q, to first order, in both the modulated and the non-modulated metamaterials. Accordingly, to first order, introducing a wave-like modulation doesnot alter the dispersion curve in the vicinity of non-coupled modes.

    (d) Coupled modesWhen (q,ω) coincides with an eigenmode that is an intersection point (q0,ω0), there exists a non-zero integer j such that (qj,ωj) is an eigenmode as well. We say that the modes (q0,ω0) and (qj,ωj)are coupled. The amplitude Ψ now takes the form (3.15) and the first-order motion equation reads

    [M(iω0 + ωm∂ξ )2 − E(iq0 + qm∂ξ )2 + k]δΨ (ξ )= [−2(iω0 + ωm∂ξ )iδωM − 2δk(ξ ) + 2(iq0 + qm∂ξ )iδqE]VjΨωj eijξ

    + (2ω0δωM − 2δk(ξ ) − 2q0δqE)V0Ψω0 . (3.23)The modulation-induced body force has again to satisfy some orthogonality conditions. To

    obtain them, consider the average of the above equation over a unit cell:

    (−ω20M + q20E + k) 〈δΨ (ξ )〉 = −2δjkVjΨωj + (2ω0δωM − 2q0δqE)V0Ψω0 , (3.24)with a superimposed bar indicating complex conjugation and

    δjk =〈δk e−ijξ

    〉(3.25)

    being the jth Fourier coefficient of the modulation k. Therefore, 〈δΨ 〉 is excited by a resonatingbody force at the eigenmode (q0,ω0) entailing

    − 2VjΨω0 · δjkΨωj + V0Ψω0 · (2ω0δωM − 2q0δqE)Ψω0 = 0. (3.26)

    In the same manner, multiplying equation (3.23) by e−ijξ and averaging yields a secondrelationship between V0 and Vj. The two equations combined imply that the coefficients V0 andVj form a solution to the eigenvalue problem⎡

    ⎣Ψω0 · (2ω0δωM − 2q0δqE)Ψω0 −2Ψω0 · δjkΨωj−2Ψωj · δjkΨω0 Ψωj · (2ωjδωM − 2qjδqE)Ψωj

    ⎤⎦ [V0

    Vj

    ]=

    [00

    ], (3.27)

    where the involved eigenvectors are explicitly given by equation (2.8).Forming the inner products and using the result for the uncoupled mode yields⎡

    ⎢⎢⎢⎢⎣

    (δω

    c0− δq

    )q0E(k − mω20)2 −δjkm2ω20ω2j

    −δjkm2ω20ω2j(δω

    cj− δq

    )qjE(k − mω2j )2

    ⎤⎥⎥⎥⎥⎦

    [V0

    Vj

    ]=

    [00

    ], (3.28)

    where c0 = c(ω0, q0) and cj = c(ωj, qj) are the group speeds. Finally, the first-order corrections(δq, δω) to the dispersion curve can be determined by setting the determinant to zero, yielding

    (δq − δω

    c0

    ) (δq − δω

    cj

    )= 1

    q0qj

    ( |δjk|m2ω20ω2jE(k − mω20)(k − mω2j )

    )2. (3.29)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 9

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    0.4

    0.2

    0

    –0.2

    –0.4

    0.4

    0.4

    0.2

    0.2

    0

    0

    –0.2

    –0.2

    –0.4

    –0.4

    w–w

    0,j

    W

    w–w

    0,j

    W

    q–q0,jQ

    0.40.20–0.2–0.4q–q0,j

    Q

    ~1,0

    0,–1

    (a)

    (b)(c)

    Figure 4. First-order corrections to the dispersion curve of a modulated metamaterial in the vicinity of the pairs of coupledmodes a, b and c of figure 3. (Online version in colour.)

    The resulting dispersion curve can be plotted in two steps. First, use equation (2.7) to plotthe dispersion curve everywhere except in the vicinity of the pairs of coupled modes (figure 3).Second, for each pair of coupled modes {(q0,ω0), (qj,ωj)}, obtain δω as a function of δq usingequation (3.29),

    δω= δq2

    (c0 + cj) ±

    √√√√√( δq2

    (c0 − cj))2

    + c0cjq0qj

    ( |δjk|m2ω20ω2jE(k − mω20)(k − mω2j )

    )2, (3.30)

    and plot both ω0,j + δω as a function of q0,j + δq for δω�∣∣ω0,j∣∣ and δq � ∣∣q0,j∣∣ (figure 4). The

    resulting dispersion curve is depicted in figure 5.It is worth mentioning that the foregoing analysis is valid regardless of the specific profile of

    δk(ξ ), the coupling between modes (q0,ω0) and (qj,ωj) being as strong as the jth Fourier coefficientδjk is large. In the illustrations, however, δk(ξ ) is sinusoidal and j takes only the values ±1 asmentioned earlier.

    In the following, we closely study the behaviour of the modulated metamaterial in the vicinityof a pair of coupled modes in two cases depending on whether the coupled modes have co-directional or contra-directional group velocities. In the vicinity of pair B where both modes havepositive group velocities, a phenomenon called ‘locking’ occurs, whereas in the vicinity of pairsA and C, ‘veering’ occurs [40]. In the present context, veering and locking correspond to whetherthe modulated metamaterial behave as a directional non-reciprocal wave converter/transmitter(pair B) or converter/reflector (pairs A and C).

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 10

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    2.5

    0

    1

    ~

    2.0

    1.5

    1.0w

    /W

    q/Q

    0.5

    0–3 –2 –1 0 1 2 3

    –1

    Figure 5. The dispersion curve of a modulatedmetamaterial obtained by combining figures 3 and 4. (Online version in colour.)

    It is of interest to underline the fact that veering and associated conversion/reflectionphenomena, at the origin of directional bandgaps, can be observed in non-dispersive mediasuch as the standard one-dimensional elastic medium [13,14]. Nonetheless, locking and relatedconversion/transmission phenomena necessitate the presence of a dispersive host medium suchas the metamaterial considered in this paper.

    (e) Directional wave conversion and transmissionSuppose an incident wave has frequency close to the intersection frequency ω0, specifically ω0 +δω. Equation (3.29) yields the first-order correction to the wavenumbers q0 and qj of the incidentand scattered waves. Both are shifted by the two values δq+ and δq−,

    δq± = δω2

    (1c0

    + 1cj

    √√√√√(δω

    2

    (1c0

    − 1cj

    ))2+ 1

    q0qj

    ( |δjk|m2ω20ω2jE(k − mω20)(k − mω2j )

    )2. (3.31)

    The associated coupling strengths C = Vj/V0 can be expressed as

    C± = −(δq± − δωc0

    )q0E(k − mω20)2

    ¯δjkm2ω20ω2j=

    ((1c0

    − 1cj

    )δω ∓ (δq+ − δq−)

    )q0E(k − mω20)2

    2δjkm2ω20ω2j

    .

    Above and in the remainder of this subsection, it is assumed that q0 and qj have the same signmaking δq and C real (e.g. pair B in figure 5).

    Having two possible values for δq and C, the constructed solution ψ is a superposition of twowaves

    ψ(x, t) = {V0+eiδqax(Ψω ei(q0x−ω0t) + (Ca + Cb)Ψωj ei(qjx−ωjt))+ V0−e−iδqax(Ψω ei(q0x−ω0t) − (Ca − Cb)Ψωj ei(qjx−ωjt))} ei(δqbx−δωt). (3.32)

    Here, V0± correspond to the roots δq±, respectively, and for simplicity we have set

    δq± = ±δqa + δqb and C± = ±Ca + Cb, (3.33)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 11

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    where the specific values of δqa, δqb and Ca, Cb follow from equations (3.31) and (3.32),respectively.

    To interpret these results, let us assume that at some position x, say x = 0, the only eigenmodepresent is the one in the vicinity of (q0,ω0). This is the case in particular when this eigenmodecorresponds to an incident wave that just entered the modulated metamaterial. Therefore, wehave V0± = (Ca ∓ Cb)V0 so that

    ψ(x, t) = 2V0{[(Ca cos(δqax) − iCb sin(δqax))Ψω ei(q0x−ω0t)

    + i(C2a − C2b) sin(δqax)Ψωj ei(qjx−ωjt)]} ei(δqbx−δωt). (3.34)

    As x increases from zero, the amplitude of the initial mode (q0,ω0) decays until it reaches itsminimum amplitude at x = d ≡ π/(2δqa). This is a consequence of the fact that |Ca| ≥ |Cb| withequality if and only if δjk = 0, which corresponds to the trivial case of no modulation. At the sametime, the amplitude of mode (qj,ωj) grows from zero, reaching its maximum after propagatingthrough a distance d. The described modal conversion becomes total when Cb = 0, which occurs ifeither δω= 0 or c0 = cj. The latter possibility is precluded by assumption, and in fact it is clear fromfigure 3 that the group speeds, i.e. slopes of the curves, are markedly different at the intersectionpoints.

    Total mode conversion, therefore, occurs if δω= 0, with

    δq± = ±∣∣δjk∣∣ m2ω20ω2j

    E∣∣k − ω20m∣∣

    ∣∣∣k − ω2j m∣∣∣ √q0qj and C± = ∓

    ∣∣k − ω20m∣∣∣∣∣k − ω2j m∣∣∣√

    q0qj

    . (3.35)

    In this special case, the incident wave has the precise frequency of mode (q0,ω0) and, afterpropagating through a distance d = π/(2δq+), it will have vanished completely leaving the placefor mode (qj,ωj) that will be the only one present at x = d. Therefore, a metamaterial modulatedthrough a distance d converts an incident wave V0ψω0 e

    i(q0x−ω0t) into a wave CV0ψωj ei(qjx−ωjt)

    changing thus its wavenumber and frequency according to

    q0 → q0 + jqm and ω0 →ω0 + jωm (3.36)

    and amplifying or weakening its amplitude by a factor C. By contrast, a wave V0ψω0 ei(−q0x−ω0t)

    incident in the opposite direction is simply transmitted unaltered as (−q0,ω0) would be anuncoupled mode.

    In conclusion, by tuning qm and ωm and using a modulated metamaterial, it is possibleto realize a selective and directional wave converter/amplifier. To qualitatively illustrate thetuning potential of the wave mode conversion phenomenon, figure 6 shows how the normalizedconversion distance Qd and the inverse amplification factor 1/ |C| vary in the mode space (levelsets). For a given input mode (q0,ω0) (marked with a black cross), the possible output modes(qj,ωj) are distributed along the dispersion curve of the non-modulated medium (yellow solidlines). To each choice of output mode, there correspond a value of d and a value of C which canbe read from the diagram. In the figure, the material constants of the metamaterial are such thatM/m = 1 and δjk/k = 0.1 and the input mode is (0.81Q, 0.525Ω). It appears then that the closer theoutput frequency is to the resonance frequency, the shorter d is and the higher C is. Therefore, forsensing applications, it seems appropriate to choose qm and ωm such that the output frequency ωjfalls in the vicinity of the resonance frequency Ω . In this manner, a weak signal of frequency ω0overwhelmed by ambient noise can be amplified and converted into an almost resonant signalof frequency Ω . Note that conversion is possible for both acoustic and optical modes, and it isfurther possible to switch between them by appropriately choosing qm and ωm.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 12

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    3.0 100

    90

    80

    70

    60

    50

    40

    30

    20

    10

    0

    5.0

    4.5

    4.0

    3.5

    3.0

    2.5

    2.0

    1.5

    1.0

    0.5

    0

    2.5

    2.0

    1.5wj/

    W

    wj/

    W

    qj /Q qj /Q

    1.0

    0.5

    0 0.5 1.0 1.5 2.0 2.5 3.0 0.5 1.0 1.5 2.0 2.5 3.0

    3.0

    2.5

    2.0

    1.5

    1.0

    0.5

    0

    (b)(a)

    Figure 6. Variations (a) of the normalized conversion length Qd and (b) of the inverted amplification factor 1/|C| as functionsof the output mode (qj ,ωj) for a given input mode (q0,ω0). The crosses mark the input mode. (Online version in colour.)

    (f) Directional wave reflectionIn the alternative scenario where q0 and qj have opposite signs (e.g. pairs A and C in figure 5),expression (3.31) implies that the wavenumber changes δq± in the vicinity of δω= 0 are

    δq± = δω2

    (1c0

    + 1cj

    )

    ± i

    √√√√√⎛⎝ 1

    |q0qj|

    ( |δjk|m2ω20ω2jE(k − mω20)(k − mω2j )

    )2−

    (δω

    2

    (1c0

    − 1cj

    ))⎞⎠2

    = δqb ± iδqc.

    These are complex valued in the neighbourhood of δω= 0, specifically for the range

    |δω|<2|δjk|m2ω20ω2j

    |q0qj|1/2E(k − mω20)(k − mω2j )

    ∣∣∣∣∣ 1c0 −1cj

    ∣∣∣∣∣−1

    . (3.37)

    Therefore, the field Ψ̃ is composed of two damped waves of frequencies ω0 and ωj and of complexwavenumbers q̃0 = q0 + δq and q̃j = qj + δq, respectively, where the complex first-order correctionδq is given above and is due to the wave-like modulation.

    The solution for the mode (q0,ω0) incident from x< 0 is, for x> 0,

    ψ(x, t) = V0e−δqcx(Ψω ei(q0x−ω0t) + C+Ψωj ei(qjx−ωjt)) ei(δqbx−δωt). (3.38)Consequently, within this frequency bandwidth centered on ω0, an incident wave of frequencyω0 + δω in the direction q0 only penetrates into the modulated metamaterial for a distance of theorder of 1/

    ∣∣δqc∣∣ and is then reflected into a wave of frequency ωj going in the direction qj oppositeto q0. The wave-like modulation thus opens locally a couple of directional bandgaps around thecoupled eigenmodes (q0,ω0) and (qj,ωj) (see subplots (A) and (C) of figure 5).

    (g) Numerical simulations of non-reciprocal conversion and reflectionIt is of interest to illustrate through direct numerical simulations the non-reciprocal behaviourof the modulated metamaterial characterized analytically so far. Specifically, we compute thetransient response of the mass-in-mass lattice subject to various excitations and analyse itsdisplacement field. Computations are performed thanks to a home-made finite difference code.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 13

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    2.5

    2.0

    1.5

    1.0w

    /W

    0.5

    0

    q/Q–3 –2 –1 0 1 2 3

    Figure 7. The dispersion curve of amodulatedmetamaterial obtained numerically (level sets) compared to the one determinedanalytically (dashed lines). (Online version in colour.)

    non-modulated modulatedwm/qm

    non-modulated

    Figure 8. Set-up for the numerical testing of non-reciprocity: a modulated portion of themetamaterial serves as a transmitterbetween two non-modulated portions.

    The constitutive and modulation parameters that we used are such that m/M = 1, qm/Q =1.8, ωm/Ω = 0.4 and δjk/k = 0.1. These values are enough to characterize the non-dimensionalresponse of the modulated metamaterial and can be completed by any particular set of parametersspecifying m, K, k and � as long as scale separation is enforced through Ω � 2√K/M.

    First, we considered a modulated metamaterial lattice containing 3000 mass-in-mass unitcells placed in a whole circle so as to simulate periodic boundary conditions. The excitationwas a broad-band tone burst displacement imposed at one of the outer masses. The imposeddisplacement has the form A0(H(t) − H(t − N/fc))(1 − cos(ωct/N)) sin(ωct), where A0 is anamplitude, fc =ωc/2π is the central frequency, H is the Heaviside function and N is the numberof cycles. The ensuing displacements at all positions and times u(x, t) are Fourier-analysed inspace and time in order to obtain Uq,ω whose amplitude is plotted as level sets over the (q,ω)-plane and giving rise to the dispersion curve (figure 7). Numerical results are shown to agreewith our analytical predictions. Note that we had to use different central frequencies fc in order tocover the whole frequency range of interest with an acceptable accuracy, the plot of figure 7 beingthe superposition of all of these intermediary results.

    Then, to test for non-reciprocal wave reflection and conversion at pairs A and B, respectively,a modulated portion of the metamaterial (1000 unit cells) is placed between two non-modulatedones (figure 8). The excitation is once again an imposed tone-burst-shaped displacement but witha much narrower spectrum (i.e. high number of cycles N) and is applied once at the left end andonce at the right end as illustrated. Figures 9 and 10 compare plots of incident, transmitted andreflected waves and, respectively, demonstrate numerically the phenomenon of non-reciprocalreflection/conversion predicted for pair A and of non-reciprocal transmission/conversion forpair B.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 14

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    incident

    transmitted

    incident

    transmitted

    reflected

    time (ms) time (ms)

    incident reflected

    1.0

    –1.0

    0.8

    –0.8

    0.6

    –0.6

    0.4

    –0.4

    0.2

    –0.20 0

    u(t)

    (arb

    .uni

    ts)

    Uw

    (arb

    .uni

    ts)

    1.0

    –1.0

    0.8

    –0.8

    0.6

    –0.6

    0.4

    –0.4

    0.2

    –0.2

    1.00.90.80.70.60.50.40.30.20.1

    0 0.5 1.0 1.5 2.0 2.5 3.0 0.5 1.0 1.5 2.0 2.5 3.0

    1.00.90.80.70.60.50.40.30.20.1

    0w/W w/W

    (b)(a)

    (c) (d )

    0 5 10 2015 25 30 0 5 10 2015 25 30

    Figure 9. Plots of transient responses u(t) demonstrating non-reciprocal reflection/conversion of waves in a modulatedmetamaterial. On (a), a left-going incident wave of central frequencyω≈ 1.8Ω is transmitted unaltered if not for dispersiveeffects. On (b), the same wave incident to the right is almost entirely reflected, slightly amplified and its frequency shifted toω≈ 1.5Ω . In the latter case, frequency spectra of incident and reflected waves are shown on (c) and (d), respectively. (Onlineversion in colour.)

    4. Short-wavelength modulationsIn this section, we investigate a family of short-wavelength modulations for which qm is muchlarger than the wavenumber of the propagating wave but remains small compared with thecharacteristic wavenumber 2π/� so that (2.2) remains valid. The hypothesis on the amplitudeof the modulation is relaxed on the other hand. The motion of the resonators will be describedby a Hill’s equation and will give rise to Bragg scattering phenomena but in time rather than inspace. Even though non-reciprocity disappears in this short-wavelength regime, the modulatedmetamaterial remains a playground for phenomena that have no counterpart in non-modulatedmetamaterials as it features a non-standard effective mass density.

    (a) HomogenizationOur starting point is the equations of motion (2.2) with k = k̃(x, t) modulated. Following themethods of homogenization theory, the variations of u and v over space and time are split intotwo slow variations over x and t and one fast over ξ = qmx − ωmt as in

    u(x, t) = u(x, qmx − ωmt, t) and v(x, t) = v(x, qmx − ωmt, t), (4.1)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 15

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    incident

    incident

    incident

    transmitted

    transmittedtransmitted

    1.0

    –1.0

    0.8

    –0.8

    0.6

    –0.6

    0.4

    –0.4

    0.2

    –0.20

    1.0

    –1.0

    0.8

    –0.8

    0.6

    –0.6

    0.4

    –0.4

    0.2

    –0.20

    u(t)

    (arb

    .uni

    ts)

    1.00.90.80.70.60.50.40.30.20.1

    1.00.90.80.70.60.50.40.30.20.1

    0 0.5 1.0 1.5 2.0 2.5 3.0w/W

    0 0.5 1.0 1.5 2.0 2.5 3.0w/W

    (b)(a)

    (c) (d )

    0 5 10 2015 25 30 35 40 0 5 10 2015 25 30 35 40time (ms) time (ms)

    Uw

    (arb

    .uni

    ts)

    Figure 10. Plots of transient responses u(t) demonPlots of transient responses u(t) demonstrating non-reciprocalreflection/conversion ofwaves in amodulatedmetamaterial. On (a), a left-going incidentwave of central frequencyω≈ 1.8Ωis transmitted unaltered if not for dispersive effects. On (b), the same wave incident to the right is almost entirely reflected,slightly amplified and its frequency shifted toω≈ 1.5Ω . In the latter case, frequency spectra of incident and reflected wavesare shown on (c) and (d), respectively. (Online version in colour.)

    where u(x, ξ , t) and v(x, ξ , t) are solutions to the two-scale motion equations

    M(∂t − ωm∂ξ )2u = −k̃(u − v) + E(∂x + �−1qm∂ξ )2uand m(∂t − ωm∂ξ )2v= −k̃(v − u).

    ⎫⎬⎭ (4.2)

    Above, we have introduced a scaling parameter � taken to be infinitely small to account for thelimit qm → ∞.

    It is possible to assume that the unknown fields u and v have sinusoidal plane harmonicvariations with respect to x and t of wavenumber q and frequency ω, that is that they areFloquet–Bloch waves:

    u(x, ξ , t) = U(ξ ) ei(qx−ωt) and v(x, ξ , t) = V(ξ ) ei(qx−ωt). (4.3)Thus, in physical terms, our working hypothesis �� 1 is equivalent to q � qm. The periodicamplitudes U and V are next expanded into the asymptotic series

    U = U0 + �U1 + �2U2 + · · · and V = V0 + �V1 + �2V2 + · · · (4.4)where Un and Vn are corrections of order n. Note that we are only interested in obtaining theequations governing the leading-order amplitudes U0 and V0 as they describe the behaviour of ametamaterial subjected to a short-wavelength modulation.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 16

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    Inserting the above expansions into the equations of motion (4.2), a hierarchy of motionequations is obtained. To the lowest order (in �−2), one has

    0 = E∂2ξU0 (4.5)

    implying that U0 is a uniform amplitude, whereas V0 remains unknown so far. The first-orderequations (in �−1) turn out to play no role and will not be detailed. The second-order equations(in �0) are

    −Mω2U0 = −k̃(U0 − V0) − Eq2U0 + ∂2ξU2

    and m(iω + ωm∂ξ )2V0 = −k̃(V0 − U0).

    ⎫⎬⎭ (4.6)

    It can be seen that the first of these equations involves the second-order correction U2, whereasthe second one is a leading-order equation governing U0 and V0. By averaging the first equationand keeping the second one as it is, we obtain a couple of motion equations governing U0 and V0,

    −Mω2U0 = −〈k(U0 − V0)〉 − Eq2U0

    and m(iω + ωm∂ξ )2V0 = −k̃(V0 − U0).

    ⎫⎬⎭ (4.7)

    Recall that U0 was proven uniform but V0 is generally ξ -dependent.

    (b) A series representation of the effective massGiven that ωm∂ξ = −∂t, the resonators satisfy the forced Hill’s equation

    m(−iω + ∂t)2V0 = −k̃(V0 − U0), (4.8)

    where k̃U0 is seen as a periodic forcing term. We focus our attention on characterizing, if any, theresonance frequenciesΩeff of the resonators. This problem is similar to determining the dispersioncurve resulting from Bragg scattering in a one-dimensional phononic crystal and has no analyticalsolutions, in general. Nonetheless, some interesting qualitative properties can still be inferredthanks to a series expansion of the effective mass operator.

    Before doing so, a re-writing of Hill’s equation will prove useful. So let T = 2π/ωm be themodulation period and τ = t/T a non-dimensional time variable. The coupling k̃ is therefore a1-periodic function in τ and the resonators equation can be recast into

    m(−iν + ∂τ )2V0 = −T2k̃(V0 − U0), (4.9)

    where ν = Tω is a non-dimensional frequency.At a given ν, let φj(ν), j ∈ N, be the eigenstates of equation (4.9). Thus, to each φj(ν) there

    corresponds an eigenvalue T2j (ν) such that

    m(−iν + ∂τ )2φj(ν) = −T2j (ν)k̃φj(ν). (4.10)

    In this manner, using standard modal analysis techniques, the periodic amplitude V0 can beexpanded into the series

    V0 =∑j∈N

    〈φ̄j(ν)k̃V

    0〉φj(ν) = −T2

    ∑j∈N

    〈k̃φ̄j(ν)

    〉φj(ν)

    T2j (ν) − T2U0, (4.11)

    where a bar indicates complex conjugation. Therefore, the coupling force has the expression

    〈k̃(U0 − V0)

    〉=

    ∑j∈N

    T2j (ν)

    T2j (ν) − T2∣∣∣〈k̃φj(ν)〉∣∣∣2 U0. (4.12)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 17

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    Tj

    T1 (Tw)

    T0 (Tw)Tw–TWeff

    TWeff

    2p

    q4q3q2q1

    Figure 11. The first two branches of a typical dispersion curve associated with Hill’s equation (4.9). (Online version in colour.)

    As a result, the effective mass density takes the form

    Meff(ν) = M + m∑j∈N

    T2

    T2 − T2j (ν)∣∣∣〈k̃φj(ν)〉 〈φj(ν)〉∣∣∣ , (4.13)

    or, as a function of the frequency ω and the modulation period T,

    Meff(ω, T) = M + m∑j∈N

    T2

    T2 − T2j (Tω)∣∣∣〈k̃φj(Tω)〉 〈φj(Tω)〉∣∣∣ . (4.14)

    (c) General properties of the effective mass densityEven though a closed-form expression is not available, the foregoing series expansion of theeffective mass density operator is quite informative. First of all, the resonance frequencies Ωeffare solutions ω to the equation

    T2j (Tω) − T2 = 0. (4.15)The first two solution branches T0 and T1 are sketched in figure 11. Depending on the value of T,the effective mass density will exhibit or lack resonances.

    (i) When the above equation has a solution Ωeff(T), it means that T falls into a1 ‘passingband’ of Hill’s equation. In this case, by Floquet–Bloch theory, all the frequencies±Ωeff(T) + (2π/T)Z are equally solutions, i.e. are equally resonance frequencies. Anexample is plotted in figure 11 where the line T = θ1 intersects the first ‘passing band’T0 at regularly spaced points with a step of 2π/T. The effective mass density will divergeat each of these intersections.

    (ii) When the above equation has no solutions, T falls within a ‘stopping band’ of Hill’sequation. In this case, no frequency ω will make the inner masses resonate. An exampleis given in figure 11 where the line T = θ3 does not intersect any of the branches Tj.

    In both cases, however, the existence of bandgaps in the dispersion diagram of the hostingmetamaterial is neither confirmed nor denied. In fact, the derived expression of the effectivemass density operator, resulting from time-domain scattering, is genuinely different fromthe one usually obtained in non-modulated metamaterials (compare for instance with theexpressions provided by Auriault & Bonnet [38] and by Milton et al. [4]). It turns out thatthe modulated metamaterial will sometimes exhibit non-resonating bandgaps, reminiscent ofdamped resonances, or, more surprisingly, will exhibit non-gapping resonances. More details aregiven subsequently.

    To start with, recall that the effective mass density operator will remain finite away fromresonance frequencies and will diverge near them to ±∞. The sign of the blow-up can be1Quotation marks are used here to underline that we are referring to the band structure of Hill’s equation describing theresonators and not the band structure of the host metamaterial.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 18

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    determined thanks to the following considerations. When T is in the nth passing band, T2n(Tω) −T2 changes sign only at resonance frequencies. For instance, in figure 11, for T = θ1, T20(Tω) − T2alternates signs at each intersection (it is positive over the blue filled regions and negative overthe red ones). Therefore, when ω goes from one resonance frequency to the next, the denominatorT20(Tω) − T2 will vanish from the same side (positive or negative) twice in a row and will makethe effective mass density diverge towards the same infinity twice in a row as well. This marksanother fundamental difference with the effective mass density of non-modulated metamaterialswhich, between two resonance frequencies, systematically goes from −∞ to +∞.

    More accurately, the above reasoning needs to be adjusted in the limiting case where T2n(Tω) −T2 vanishes but does not change sign. This happens when the level of T is tangent from above orfrom below to a branch Tj. For instance, in figure 11, T = θ2 is tangent from above to T0 so thatT20(Tω) − T2 remains non-positive. Conversely, T = θ4 is tangent from below to T1 and T21(Tω) −T2 stays non-negative. In these cases, the effective mass density diverges to the same infinity whenω approaches any resonance frequency by increasing or decreasing values.

    In summary, the series of infinities visited by Meff, for increasing positive ω, alternatesperiodically following the pattern (+ − −+) or (− + +−) except at the very end of a passingband where the pattern changes into (+ + ++) and at the very beginning of a new passingband in which case the pattern is (− − −−). Hereafter, these patterns are referred to as signatures.Recall that the signature of a non-modulated metamaterial with multiple resonances is necessarily(+ − +−).

    Last, note that for infinitely small T unique values of the resonance frequency Ωeff(T) −→T→0

    Ωeff

    and mass Meff(ω, T) −→T→0

    Meff are observed, where

    Ωeff =√

    〈k〉m

    and Meff = M + mΩ2eff

    Ω2eff − ω2. (4.16)

    The example described next, where a closed-form analytical expression of Meff is available, willhelp appreciate the general properties presented in this subsection and illustrate new phenomenathat have no counterpart in non-modulated inner-resonant metamaterials.

    (d) Example: a high-amplitude modulationA fairly simple closed-form analytical solution for Hill’s equation can be obtained when thecoupling stiffness k̃ is a ‘Dirac comb’ as in

    Ω2(t) = k̃(t)m

    = α + Tβ∑j∈Z

    δ(t − jT), (4.17)

    where δ is the Dirac function. Though experimentally challenging, such modulation can beachieved in theory by letting k̃ take a first positive value, k = mα, over a period of time equal toT −�t and then another positive value, k + δk = m(α + Tβ/�t), over a period of �t, for as small�t as possible (figure 12). The considered modulation has therefore a very high amplitude.

    The analytical advantage of Dirac combs is that they have particularly simple Fouriercoefficients. For instance, those of Ω2 are Ω20 = α + β, Ω2η �=0 = β, where the Fourier variableη spans (2π/T)Z. Taking this and the fact that U0 is uniform into account, expansion ofequation (4.8) yields

    − (ω + η)2Vη = −αVη − β∑η′

    Vη′ + αUδn0 + βU, (4.18)

    where superscripts have been dropped to simplify notations. Summing the coefficients Vη overall η leads to ∑

    η′Vη′ =

    (1 + ω

    2

    α − ω2 + βg(ω, T)

    )U, (4.19)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 19

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    k + dk

    k~

    k

    T Dt � T t

    Figure 12. A high-amplitude modulation.

    8

    6

    4

    2

    0

    8

    6

    4

    2

    0

    8

    6

    4

    2

    0–5 0

    Meff/M5 –5 0

    Meff/M5 –5 0

    Meff/M5

    w/

    a÷–w

    / a÷–

    w/

    a÷–

    (e) ( f )

    (b)(a) (c)

    (d )

    (i)(g) (h)

    Figure 13. The profile of the normalized effective mass densityMeff/M as a function of the normalized frequencyω/√α for

    various values of T√α = 0.1; 1.25; 1.49; 1.73; 2.01; 2.96; 3.12; 3.71; 100 (a − i, in order). The fundamental bandgap is shaded.

    The properties of the corresponding metamaterial are such thatM= m andα= β . (Online version in colour.)

    where

    g(ω, T) =∑η

    α − ω2α − (ω + η)2 =

    (α − ω2)T√α sin(T√α)2α(cos(Tω) − cos(T√α)) . (4.20)

    This allows us to conclude that the coupling force is

    〈k(U − V)〉 = − mω2

    α − ω2(α − βω

    2

    α − ω2 + βg(ω, T)

    )U, (4.21)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 20

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    and that the effective mass can be expressed as

    Meff(ω, T) = M +mα

    α − ω2(α − βω

    2

    α − ω2 + βg(ω, T)

    ). (4.22)

    This closed-form expression of Meff was used to generate the plots of figure 13. Values ofT were chosen to illustrate different phenomena predicted by the general considerations ofthe previous subsection. On (a), the standard profile for a high-frequency short-wavelengthmodulation is plotted. On (b), multiple resonances with a signature (+ − −+) appear. Unlike thesignature (+ − +−) of non-modulated metamaterials, the signature (+ − −+) does not obstructbandgaps merging (c). Non-gaping resonances appear on (d) as T leaves the first ‘passing band’.Resonances disappear on (e) where T is now in the first ‘stopping band’. Plot ( f ) demonstrates thepossibility of non-resonating bandgaps. A bandgap which extends to both sides of a resonancefrequency is illustrated on (g). As T enters the second ‘passing band’, the signature is invertedinto (− + +−) and resonances become denser because the step 2π/T keeps decreasing as it is seenon (h). Last, on (i), for infinitely large T, the metamaterial is no longer modulated and a standardprofile is recovered. In comparison to figure 11, plots (a,b,c) correspond to values similar to T = θ1,plot (d) to T = θ2, plots (e,f ) to T = θ3, plot (g) to T = θ4 and plots (h, i) to higher values of T.

    The configurations sketched in figure 13c,g are of particular interest as they feature unusuallybroad bandgaps almost twice as wide as the bandgap that the non-modulated metamaterialexhibits on plot (i). Note finally that in some frequency domains, Hill’s equation is known tofeature instabilities that should be avoided for the above results to be fully useful [1].

    5. ConclusionQualitatively, within a weakly modulated medium, travelling and pump waves interact onlywhen a phase-matching condition is satisfied. If so, the incident wave will be scattered by thepump wave. When the incident and scattered waves propagate in the same direction, we speakof ‘veering’: the incident wave is converted into a transmitted wave with a shifted frequency.Otherwise, the incident wave is fully reflected, a phenomenon called ‘locking’. Note that veeringcan only occur in dispersive media, whereas locking occurs in both dispersive and non-dispersivemedia. The modulated metamaterial considered in this paper was shown to feature both effects.

    The asymmetric nature of the phase matching condition immediately implies that theinteraction between travelling and pump waves is directional. Thus, waves that are eitherconverted or reflected when incident in one direction are simply transmitted unaltered if incidentin the opposite direction. The amplification factor and characteristic length of these one-wayphenomena turned out to be tunable as well thanks to the dispersive nature of the hostmetamaterial.

    Furthermore, beyond non-reciprocal effects, strong space–time modulations have a significantimpact on the behaviour of the effective mass operator in the homogenization limit. As a matterof fact, in addition to the main bandgap of the inner-resonant structure, a set of low-frequencytime-domain-Bragg-scattering bandgaps was observed. Most importantly, in contrast with non-modulated metamaterials, these new modulation-induced bandgaps were shown to be able tomerge.

    Modulating the resonance frequency of an inner resonant metamaterial so as to breaktime-reversal symmetry and reciprocity or to generate new non-standard effective massoperators offers a promising potential in the field of wave control in general and in one-wayconversion/amplification in particular.

    Data accessibility. This paper does not have any experimental data.Authors’ contributions. H.N. and G.H. conceived the core concepts and mathematical model. H.N., A.N. and G.H.derived and interpreted the analytical results. H.C. and H.N. carried the numerical simulations. G.H., A.N.and M.H. supervised the modelling and simulation. All authors discussed the results, commented on themanuscript and approved it.

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://rspa.royalsocietypublishing.org/

  • 21

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    Competing interests. The authors have no competing interests.Funding. This work is supported by the Air Force Office of Scientific Research under grant no. AF 9550-15-1-0016 with Program Manager Dr Byung-Lip (Les) Lee and the NSF EFRI under award no. 1641078.

    References1. Brillouin L. 1953 Wave propagation in periodic structures: electric filters and crystal lattices. Dover

    Phoenix Editions. New York, NY: Dover Publications.2. Liu Z, Zhang X, Mao Y, Zhu YY, Yang Z, Chan CT, Sheng P. 2000 Locally resonant sonic

    materials. Science 289, 1734–1736. (doi:10.1126/science.289.5485.1734)3. Li J, Chan CT. 2004 Double-negative acoustic metamaterial. Phys. Rev. E 70, 055602.

    (doi:10.1103/PhysRevE.70.055602)4. Milton GW, Briane M, Willis JR. 2006 On cloaking for elasticity and physical equations with a

    transformation invariant form. New J. Phys. 8, 248–267. (doi:10.1088/1367-2630/8/10/248)5. Ding Y, Liu Z, Qiu C, Shi J. 2007 Metamaterial with simultaneously negative bulk modulus

    and mass density. Phys. Rev. Lett. 99, 093904. (doi:10.1103/PhysRevLett.99.093904)6. Yao S, Zhou X, Hu GK. 2008 Experimental study on negative effective mass in a 1D mass-

    spring system. New J. Phys. 10, 043020. (doi:10.1088/1367-2630/10/4/043020)7. Auriault JL, Boutin C. 2012 Long wavelength inner-resonance cut-off frequencies in elastic

    composite materials. Int. J. Solids Struct. 49, 3269–3281. (doi:10.1016/j.ijsolstr.2012.07.002)8. Zhu R, Liu XN, Hu GK, Sun CT, Huang GL. 2014 Negative refraction of elastic waves

    at the deep-subwavelength scale in a single-phase metamaterial. Nat. Commun. 5, 5510.(doi:10.1038/ncomms6510)

    9. Yu Z, Fan S. 2009 Complete optical isolation created by indirect interband photonic transitions.Nat. Photonics 3, 91–94. (doi:10.1038/nphoton.2008.273)

    10. Zanjani MB, Davoyan AR, Mahmoud AM, Engheta N, Lukes JR. 2014 One-way phononisolation in acoustic waveguides. Appl. Phys. Lett. 104, 081905. (doi:10.1063/1.4866590)

    11. Zanjani MB, Davoyan AR, Engheta N, Lukes JR. 2015 NEMS with broken T symmetry:graphene based unidirectional acoustic transmission lines. Sci. Rep. 5, 09926. (doi:10.1038/srep09926)

    12. Fleury R, Sounas DL, Sieck CF, Haberman MR, Alù A. 2014 Sound isolation and giantlinear nonreciprocity in a compact acoustic circulator. Science 343, 516–519. (doi:10.1126/science.1246957)

    13. Swinteck N, Matsuo S, Runge K, Vasseur JO, Lucas P, Deymier PA. 2015 Bulk elasticwaves with unidirectional backscattering-immune topological states in a time-dependentsuperlattice. J. Appl. Phys. 118, 063103. (doi:10.1063/1.4928619)

    14. Trainiti G, Ruzzene M. 2016 Non-reciprocal elastic wave propagation in spatiotemporalperiodic structures. New J. Phys. 18, 083047. (doi:10.1088/1367-2630/18/8/083047)

    15. Nassar H, Xu XC, Norris AN, Huang GL. 2017 Modulated phononic crystals: Non-reciprocal wave propagation and Willis materials. J. Mech. Phys. Solids 101, 10–29.(doi:10.1016/j.jmps.2017.01.010)

    16. Liang B, Yuan B, Cheng JC. 2009 Acoustic diode: Rectification of acoustic energy flux in one-dimensional systems. Phys. Rev. Lett. 103, 104301. (doi:10.1103/PhysRevLett.103.104301)

    17. Mousavi SH, Khanikaev AB, Wang Z. 2015 Topologically protected elastic waves in phononicmetamaterials. Nat. Commun. 6, 8682. (doi:10.1038/ncomms9682)

    18. Lu J, Qiu C, Ke M, Liu Z. 2016 Valley vortex states in sonic crystals. Phys. Rev. Lett. 116, 093901.(doi:10.1103/PhysRevLett.116.093901)

    19. Lu J, Qiu C, Ye L, Fan X, Ke M, Zhang F, Liu Z. 2016 Observation of topological valleytransport of sound in sonic crystals. Nat. Phys. 13, 369–374. (doi:10.1038/nphys3999)

    20. Fleury R, Sounas DL, Haberman MR, Alù A. 2015 Nonreciprocal acoustics. Acoust. Today 11,14–21.

    21. Khanikaev AB, Fleury R, Mousavi SH, Alù A. 2015 Topologically robust sound propagationin an angular-momentum-biased graphene-like resonator lattice. Nat. Commun. 6, 8260.(doi:10.1038/ncomms9260)

    22. Yang Z, Gao F, Shi X, Lin X, Gao Z, Chong Y, Zhang B. 2015 Topological acoustics. Phys. Rev.Lett. 114, 114301. (doi:10.1103/PhysRevLett.114.114301)

    23. Chen ZG, Wu Y. 2016 Tunable topological phononic crystals. Phys. Rev. Appl. 5, 054021.(doi:10.1103/PhysRevApplied.5.054021)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://dx.doi.org/doi:10.1126/science.289.5485.1734http://dx.doi.org/doi:10.1103/PhysRevE.70.055602http://dx.doi.org/doi:10.1088/1367-2630/8/10/248http://dx.doi.org/doi:10.1103/PhysRevLett.99.093904http://dx.doi.org/doi:10.1088/1367-2630/10/4/043020http://dx.doi.org/doi:10.1016/j.ijsolstr.2012.07.002http://dx.doi.org/doi:10.1038/ncomms6510http://dx.doi.org/doi:10.1038/nphoton.2008.273http://dx.doi.org/doi:10.1063/1.4866590http://dx.doi.org/doi:10.1038/srep09926http://dx.doi.org/doi:10.1038/srep09926http://dx.doi.org/doi:10.1126/science.1246957http://dx.doi.org/doi:10.1126/science.1246957http://dx.doi.org/doi:10.1063/1.4928619http://dx.doi.org/doi:10.1088/1367-2630/18/8/083047http://dx.doi.org/doi:10.1016/j.jmps.2017.01.010http://dx.doi.org/doi:10.1103/PhysRevLett.103.104301http://dx.doi.org/doi:10.1038/ncomms9682http://dx.doi.org/doi:10.1103/PhysRevLett.116.093901http://dx.doi.org/doi:10.1038/nphys3999http://dx.doi.org/doi:10.1038/ncomms9260http://dx.doi.org/doi:10.1103/PhysRevLett.114.114301http://dx.doi.org/doi:10.1103/PhysRevApplied.5.054021http://rspa.royalsocietypublishing.org/

  • 22

    rspa.royalsocietypublishing.orgProc.R.Soc.A473:20170188

    ...................................................

    24. Slater JC. 1958 Interaction of waves in crystals. Rev. Mod. Phys. 30, 197–222. (doi:10.1103/RevModPhys.30.197)

    25. Simon JC. 1960 Action of a progressive disturbance on a guided electromagnetic wave. IRETrans. Microw. Theory Tech. 8, 18–29. (doi:10.1109/TMTT.1960.1124657)

    26. Louisell W. 1960 Coupled mode and parametric electronics. New York, NY: Wiley.27. Hessel A, Oliner AA. 1961 Wave propagation in a medium with a progressive sinusoidal

    disturbance. IRE Trans. Microw. Theory Tech. 9, 337–343. (doi:10.1109/TMTT.1961.1125340)28. Cassedy ES, Oliner AA. 1963 Dispersion relations in time-space periodic media: Part I-stable

    interactions. Proc. IEEE 51, 1342–1359. (doi:10.1109/PROC.1963.2566)29. Cassedy ES, Oliner AA. 1963 Dispersion relations in time-space periodic media: Part II-

    unstable interactions. Proc. IEEE 51, 1154–1168. (doi:10.1109/PROC.1963.2566)30. Lurie KA. 2007 An introduction to the mathematical theory of dynamic materials. New York, NY:

    Springer.31. Milton GW, Mattei O. 2017 Field patterns: a new mathematical object. Proc. R. Soc. A 473,

    20160819. (doi:10.1098/rspa.2016.0819)32. Casadei F, Delpero T, Bergamini A, Ermanni P, Ruzzene M. 2012 Piezoelectric resonator

    arrays for tunable acoustic waveguides and metamaterials. J. Appl. Phys. 112, 064902.(doi:10.1063/1.4752468)

    33. Chen YY, Huang GL, Sun CT. 2014 Band gap control in an active elastic metamaterialwith negative capacitance piezoelectric shunting. J. Vib. Acoust. 136, 061008.(doi:10.1115/1.4028378)

    34. Chen YY, Hu GK, Huang GL. 2016 An adaptive metamaterial beam with hybrid shuntingcircuits for extremely broadband control of flexural waves. Smart Mater. Struct. 25, 105036.(doi:10.1088/0964-1726/25/10/105036)

    35. Reed EJ, Soljačić M, Joannopoulos JD. 2003 Reversed Doppler effect in photonic crystals. Phys.Rev. Lett. 91, 89–92. (doi:10.1103/PhysRevLett.91.133901)

    36. Danas K, Kankanala SV, Triantafyllidis N. 2012 Experiments and modeling ofiron-particle-filled magnetorheological elastomers. J. Mech. Phys. Solids 60, 120–138.(doi:10.1016/j.jmps.2011.09.006)

    37. Gump J, Finkler I, Xia H, Sooryakumar R, Bresser WJ, Boolchand P. 2004 Light-induced giantsoftening of network glasses observed near the mean-field rigidity transition. Phys. Rev. Lett.92, 245501. (doi:10.1103/PhysRevLett.92.245501)

    38. Auriault JL, Bonnet G. 1985 Dynamics of periodic elastic composites (Dynamique descomposites elastiques periodiques). Arch. Mech. Stosow. 37, 269–284.

    39. Huang HH, Sun CT, Huang GL. 2009 On the negative effective mass density in acousticmetamaterials. Int. J. Eng. Sci. 47, 610–617. (doi:10.1016/j.ijengsci.2008.12.007)

    40. Mace BR, Manconi E. 2012 Wave motion and dispersion phenomena: veering, locking andstrong coupling effects. J. Acoust. Soc. Am. 131, 1015–1028. (doi:10.1121/1.3672647)

    on June 30, 2017http://rspa.royalsocietypublishing.org/Downloaded from

    http://dx.doi.org/doi:10.1103/RevModPhys.30.197http://dx.doi.org/doi:10.1103/RevModPhys.30.197http://dx.doi.org/doi:10.1109/TMTT.1960.1124657http://dx.doi.org/doi:10.1109/TMTT.1961.1125340http://dx.doi.org/doi:10.1109/PROC.1963.2566http://dx.doi.org/doi:10.1109/PROC.1963.2566http://dx.doi.org/doi:10.1098/rspa.2016.0819http://dx.doi.org/doi:10.1063/1.4752468http://dx.doi.org/doi:10.1115/1.4028378http://dx.doi.org/doi:10.1088/0964-1726/25/10/105036http://dx.doi.org/doi:10.1103/PhysRevLett.91.133901http://dx.doi.org/doi:10.1016/j.jmps.2011.09.006http://dx.doi.org/doi:10.1103/PhysRevLett.92.245501http://dx.doi.org/doi:10.1016/j.ijengsci.2008.12.007http://dx.doi.org/doi:10.1121/1.3672647http://rspa.royalsocietypublishing.org/

    IntroductionWave propagation in non-modulated elastic metamaterialsDiscrete and continuous modelling of a metamaterialFree and forced wave propagation

    Wave propagation in modulated elastic metamaterialsWave-like modulationsPerturbation methodUncoupled modesCoupled modesDirectional wave conversion and transmissionDirectional wave reflectionNumerical simulations of non-reciprocal conversion and reflection

    Short-wavelength modulationsHomogenizationA series representation of the effective massGeneral properties of the effective mass densityExample: a high-amplitude modulation

    ConclusionReferences


Recommended