Absence Of Unidirectionally Propagating Surface Plasmon-polaritons At .

1y ago
3 Views
1 Downloads
884.66 KB
6 Pages
Last View : 1m ago
Last Download : 3m ago
Upload by : Mollie Blount
Transcription

PENAbsence of unidirectionally propagatingsurface plasmon-polaritons at nonreciprocalmetal-dielectric interfaces1234567890():,;Siddharth BuddhirajuIan A.D. Williamson1*,Yu Shi1, Alex Song 1, Casey Wojcik1, Avik Dutt 1 & Shanhui Fan 11,Momchil Minkov1,In the presence of an external magnetic field, the surface plasmon polariton that exists at themetal-dielectric interface is believed to support a unidirectional frequency range near thesurface plasmon frequency, where the surface plasmon polariton propagates along one butnot the opposite direction. Recent works have pointed to some of the paradoxical consequences of such a unidirectional range, including in particular the violation of the timebandwidth product constraint that should otherwise apply in general in static systems. Herewe show that such a unidirectional frequency range is nonphysical using both a generalthermodynamic argument and a detailed calculation based on a nonlocal hydrodynamicDrude model for the metal permittivity. Our calculation reveals that the surface plasmonpolariton at metal-dielectric interfaces remains bidirectional for all frequencies.1 GinztonLaboratory, Department of Electrical Engineering, Stanford University, Stanford, CA, USA. *email: sbuddhi@stanford.edu; shanhui@stanford.eduNATURE COMMUNICATIONS (2020)11:674 https://doi.org/10.1038/s41467-020-14504-9 www.nature.com/naturecommunications1

ARTICLENATURE COMMUNICATIONS rude model. We start with a brief review of the dispersionrelation of the surface-plasmon polariton at the metal–dielectricinterface. Throughout the paper, for simplicity, we refer to anysystem with a strong plasmonic response as a metal. In additionto the usual free-electron metals, such “metals” also includeheavily doped semiconductors that exhibit a plasmonic responseat infrared wavelengths. Within the local Drude model, in thepresence of a static magnetic field B0 ¼ B0 y, the frequency(ω)dependent dielectric function of metal has the following form:ω2pϵm ðωÞ¼1 ϵ1ðω þ iγ0 Þ2 ω2c0γ01 þ i ω0Bðω þ iγ0 Þ2 ω2c B@ 0ωðω þ iγ Þ i ωωc2001i ωωc0γ1 þ i ω0CC;Að1ÞaDielectricx dyB0zMetal m ( )b0.8 (in p)In the past 2 decades, there have been significant developmentsin the field of plasmonics, which explores surface-plasmonpolaritons that exist at metal–dielectric interfaces to achievenanoscale control of light1–3. Most plasmonic structures satisfythe Lorentz reciprocity theorem4. On the other hand, in thepresence of an external magnetic field, the behavior of surfaceplasmon polaritons becomes nonreciprocal. Such nonreciprocalsurface-plasmon polaritons have generated substantial interest5–8since they represent a fundamentally different regime of lightpropagation, having potential importance for applications such assensing and information processing.A particularly significant effect of nonreciprocal plasmons is theexistence of a unidirectional frequency range. Such unidirectionalfrequency ranges have been shown to occur in topologicallynontrivial metal–metal interfaces9–14, where the unidirectionalbehavior is linked to the topology of the bandstructure, or inmetal–dielectric systems15–18. In this paper, we will discussmetal–dielectric systems. With the metal described by the Drudemodel, in the presence of an external magnetic field, there exists afrequency range where the surface-plasmon polariton can onlypropagate along one direction. In a recent paper18, it was notedthat the existence of such unidirectional surface-plasmon polaritons can lead to resonator structures that violate the timebandwidth product constraint that should otherwise apply ingeneral in static systems19. Subsequently, refs. 20,21 argued thatbreaking the time-bandwidth product should not be possiblebased on a coupled-mode theory analysis, and that doing so mayviolate the second law of thermodynamics. Motivated by theseconsiderations, it becomes important to re-examine the fundamental physical assumptions that give rise to unidirectionallypropagating surface-plasmon polaritons at metal–dielectric interfaces. Within the local Drude model, the existence of the unidirectional frequency range depends on the behavior of the modelin the limit of large wavevectors. On the other hand, it has longbeen known22 that nonlocal effects become important in this limit.In this paper, we show that there should not be a unidirectionalfrequency range in the spectrum of the surface-plasmon polaritonat metal–dielectric interfaces once nonlocal effects are considered.Instead, there will always be propagating modes in both directions. We illustrate this by detailed calculations based on thehydrodynamic model for the metal dielectric function. We alsopresent a general thermodynamic argument to show that themain conclusion of the paper, i.e., the absence of unidirectionallypropagating surface-plasmons polaritons at nonreciprocalmetal–dielectric interfaces, should hold for any physical nonlocalmodel of the dielectric function.0.60.4 c 00.20.0–20 c 0.2 p–100K (in kp)1020Fig. 1 Surface-plasmon-polariton dispersion relation in the local Drudemodel. a An interface between a dielectric and a metal described by theDrude model. b Dispersion relation of surface-plasmon-polaritonpropagation at the interface in the absence (green curve) and in thepresence (blue curve) of an external magnetic field B0 as indicated in a. K isthe wavevector component parallel to the interface, kp ωp/c, and ωc eB0 m is the cyclotron frequency.where ϵ is the dielectric response from the bound electrons andions, ωp the plasma frequency, γ0 the phenomenological scattering loss rate, and ωc eB0 m the cyclotron frequency, withfundamental charge e and effective mass of the free carriers m.Consider the metal–dielectric interface shown in Fig. 1a. WhenB0 0, in the near-lossless limit of γ0 0, the dispersion relationω(K) of the surface-plasmon polariton is shown by the greencurve in Fig. 1b, where K is the wavevector parallel to theinterface. Here we assume that ϵ 1, and that the dielectric isair with ϵd 1. Since the system is reciprocal, we have ω(K) ω( K). In the K limit, the frequency of the surfaceplasmon-polaritonapproaches the surface-plasmon frequencypffiffiffiωsp ¼ ωp 2.When B0 0, again assuming the near-lossless limit, thedispersion relation for the same interface system is shown by theblue curve in Fig. 1b. Since the system is no longer reciprocal, wehave ω(K) ω( K). Moreover, the surface-plasmon frequenciesfor the forward and backward directions are unequal at the K limit, opening a unidirectional frequency range aroundωsp where the surface plasmon polariton propagates only alongthe positive-K direction. The existence of such a unidirectionalfrequency range is the key to the unusual time-bandwidthproduct behavior reported in ref. 18. On the other hand, asobserved from Fig. 1b, the existence of such a unidirectionalrange is strongly dependent on the behavior of the Drude modelin the K limit. And yet, it is well known that the localDrude model of Eq. (1) is no longer adequate in this limit22, andinstead the spatially dispersive or “nonlocal” behavior of theelectromagnetic response of the metal must be considered. Forexample, it was also shown11 that topological effects incontinuous media need to be described taking into accountnonlocality. Therefore, to understand the potential physics ofsuch unidirectional propagation, it is essential to consider theeffect of the nonlocal dielectric function.NATURE COMMUNICATIONS (2020)11:674 https://doi.org/10.1038/s41467-020-14504-9 www.nature.com/naturecommunications

ARTICLENATURE COMMUNICATIONS https://doi.org/10.1038/s41467-020-14504-9 nð3Þ¼ ðnvÞ; twhere β is the nonlocal parameter proportional to the Fermivelocity vF35,3β2 ¼ 5ω þ 13 iγ 2v :ω þ iγ Fð4ÞLinearizing Eqs. (2) and (3), and defining the free-electroncurrent J en0v, where n0 is the equilibrium electron density, asingle equation can be obtained for J in the frequency domain aseβ2 ð JÞ þ ωðω þ iγÞJ ¼ iωðω2p ϵ1 ϵ0 E J B0 Þ;ð5Þmwhere B0 is an externally applied dc magnetic field. This equationis coupled with Maxwell’s equations, written using the E field asω2ð6Þϵ E iωμ0 J:c2 1Unlike in the local model, the presence of the nonlocal termβ2 ( ) in this model requires an additional boundarycondition30 to determine the dispersion relation. Here, theadditional boundary condition required to solve Eqs. (5) and(6) is E ¼ ¼ 0;J nð7Þ is the unit vector normal to the metal–dielectricwhere ninterface. This has the effect of imposing an infinite potential wellfor the electron gas at the metal–dielectric boundary.To illustrate the effect of nonlocality on the nonreciprocalbehavior of the surface-plasmon polaritons, we consider aninterface where the dielectric layer is silicon (ϵd 11.68) and themetallic layer is n-doped InSb, a material commonly used indemonstrating magneto-optical plasmonic effects15,18. This interface was previously used in ref. 18, with the InSb layer treatedusing the local Drude model. The InSb layer has ϵ 15.6 andplasma frequency ωp 2π (2 1012 Hz). A constant dc magnetic field of B0 0.2 T is applied in the y direction to breakreciprocity. Owing to a rather small conductivity effective massfor electrons, a large value31 of β 1.07 106 ms 1 is obtained at300 K. Thus, the effect of nonlocality, which was not consideredin ref. 18, is in fact prominent in the dielectric response of InSb. Inorder to highlight the difference between the hydrodynamica1.0 (in p)0.80.60.4Local0.2Nonlocal 0–100–500K (in kp)50100b0.8 (in p)Hydrodynamic Drude model. There exist many treatments ofnonlocality, such as those based on the hydrodynamic model23,24,the random phase approximation25,26, and a quantum-correctedmodel27. The description of plasmonic properties using thesemodels is closely related to the emerging area of quantum plasmonics, where the quantum nature of the electron gas plays asignificant role27–29. Here, we briefly discuss the hydrodynamicmodel, a simple analytic model that has often been used todescribe nonlocal response in deep subwavelength metallicstructures23,30, and recently in nanoparticles made of dopedsemiconductors such as indium antimonide (InSb)31. We referreaders to refs. 30,32 and references therein for a detailed overviewof nonlocality in surface-plasmon polaritons as well as a derivation of the hydrodynamic model.In this model, the collective motion of electrons is describedusing a density n(r, t), a velocity v(r, t), and an energy functionalthat can be appropriately chosen to describe the internalkinetic energy as well as interactions. We follow ref. 30 to employthe Thomas–Fermi approximation for the energy functional. Theequations of motion of the free carriers in this approximation aregiven by33,34 ve nð2Þþ γv þ ðv Þv ¼ ðE þ v BÞ β2; tmn0.60.4Re (K )0.2Im (K ) 0 0.025 p–102–100100102K (in kp)Fig. 2 Dispersion relations in the local and hydrodynamic Drude models.Dispersion relation for the interface considered in ref. 18, in the local (blue)and nonlocal (red) models. a A flat dispersion relation is obtained in thelossless (γ0 0) local Drude model, resulting in a unidirectional gapindicated by the dotted gray lines. This gap is removed in the hydrodynamicDrude model with a high-K counter-propagating wave. b Real (solid line)and imaginary (dotted line) parts of K when γ0 0.025ωp in the local(blue) and the nonlocal (red) models. Landau damping is self-consistentlyincorporated in the nonlocal model. ReðKÞ ImðKÞ in the unidirectional gapin the local model, while ReðKÞ ImðKÞ in the nonlocal model. Note thatK kp is in log scale.model and the local Drude model, we first set γ0 0. Using theseparameters, we solve Eqs. (5) and (6) for surface-plasmonpolariton dispersion relation at the Si–InSb interface. The redcurve in Fig. 2a depicts the dispersion relation in thehydrodynamic model, and the blue curve in the local Drudemodel. The dispersion relations from the two models are almostthe same for small K, but deviate as K becomes larger. Inparticular, within the hydrodynamic model, there is no longer aunidirectional frequency range. At every frequency, there are botha propagating and a counter-propagating mode. We also notethat the predictions between the local and nonlocal models startto deviate for K 0.4 μm 1. Thus, in this system, nonlocal effectsbecome important even for surface-plasmon waves with wavelength on the submicron scale, in contrast with standardplasmonic metals where nonlocal effects are important onlywhen the plasmon wavelength is at the nanoscale.The qualitative difference between the hydrodynamic modeland the local Drude model persists over a wide range of loss ratesγ in Eq. (2). The loss in a surface-plasmon polariton arises notonly from scattering, but also surface-induced Landaudamping36,37. Unlike bulk plasmons where Landau dampingoccurs only for K ω vF, surface modes also experience Landaudamping at smaller wavevectors owing to confinement in theNATURE COMMUNICATIONS (2020)11:674 https://doi.org/10.1038/s41467-020-14504-9 www.nature.com/naturecommunications3

NATURE COMMUNICATIONS https://doi.org/10.1038/s41467-020-14504-9is the loss rate from Landau damping. Here, Fx,z(q) is the Fouriertransform of the electric field Ex,z(x) in the metal, and q isnormalized to the onset of Landau damping, i.e., to ω vF. Since γsdepends on the field profiles which in turn depend on γs, we solvefor the damping rate and the fields in a self-consistent manner.In Fig. 2b, we plot the dispersion relation for γ0 0.025ωp inblue for the local model and red for the nonlocal model. In thenonlocal model, the effective loss rate is given by γ0 γsdescribed above. The solid lines represent ReðKÞ, while thedotted lines represent ImðKÞ. Within the local model, whilepropagation is not strictly unidirectional in the presence ofloss, the counter-propagating mode is significantly overdamped(ReðKÞ ImðKÞ) in the unidirectional frequency range,marked by the gray dotted lines. On the other hand, thecounter-propagating mode continues to remain underdamped(ReðKÞ ImðKÞ) in the nonlocal model. Only for substantiallyhigh values of loss (γ0 0.05ωp) does the dispersion relation inthe nonlocal model return approximately to its local form, inwhich case damping is high enough that the propagation of thesurface-plasmon polariton is no longer apparent. The analysishere indicates that the effect of nonlocality on nonreciprocalsurface-plasmon polaritons should be pronounced for a widerange of values of loss.In order to numerically demonstrate the effect of nonlocalityon nonreciprocal photon transport, we re-examine the structureshown in Fig. 3a, which was first considered in ref. 18. Thestructure is two-dimensional and consists of the Si–InSb interfaceas discussed above, subject to a static out-of-plane magnetic field.Such an interface thus behaves as a nonreciprocal plasmonicwaveguide. The waveguide is surrounded by a metal region,which serves both to truncate the waveguide at one end, as well asto eliminate any radiation losses. In ref. 18, the surroundingregion was assumed to be silver. Here for simplicity we assume asurrounding region made of a perfect electric conductor (PEC),which makes very little difference to the simulations sincethe operating frequency, in the far-infrared region, is far belowthe plasma frequency of silver. Ref. 18 treats the InSb layerusing the local dielectric function of Eq. (1). The choice of themagnetic field along the y direction results in a unidirectionalfrequency range where there is a surface-plasmon polaritonpropagating toward the truncation, but not in the oppositedirection. Consequently, at a frequency inside the unidirectionalrange, ref. 18 shows that the electromagnetic field will propagatetoward and be trapped at the truncation, with no leakage either inthe backward direction, or through radiation losses. Such atrapping effect appears to lead to the violation of the timebandwidth product constraint.On the other hand, as we have discussed above, the nonlocalbehavior is in fact intrinsic and rather significant in the dielectricresponse of InSb. Therefore, we extend the finite-differencefrequency-domain method38 to include the nonlocal response asdescribed by Eq. (5) for the InSb region, and resimulate thestructure in Fig. 3a. To highlight the fact that there is a backwardpropagating mode even in the lossless system, we assume γ 0.We excite the waveguide mode by placing a line source normal tothe interface. We choose ω 0.7ωp, a frequency that is inside theunidirectional range of the local model. In Fig. 3b, we plot thefield distribution of Hz, the z-component of the magnetic field ofthe excited surface-plasmon-polariton mode. We observe asignificant excitation of backward propagating surface-plasmon4axSibzyB0InSb40 1Line sourcex (in μm)direction normal to the interface. Following ref. 36, we write γ γ0 γs, where γ0 is the loss rate from scattering andR 1 3q jF x ðqÞj2 dq3πωγs ¼ð8ÞR1 12 0 jF x ðqÞj2 þ jF z ðqÞj2 dq2000–10cSz (a.u.)ARTICLE204060z (in μm)8010030–3S, totalE H contributionJ* ·J contribution–6–90204060z (in μm)80100Fig. 3 Numerical simulation of a truncated waveguide using thehydrodynamic model. A finite-difference frequency-domain (FDFD)algorithm is used to obtain the numerical results, at ω 0.7ωp. a Thestructure as considered in ref. 18. b Field profile of Hz generated by theindicated line source, clearly depicting a backward propagating mode with asignificantly smaller wavevector than the forward propagating mode. c Afinite Poynting flux in the backward direction relative to the line source, anda zero Poynting flux forward owing to the PEC to the right. The blue curve isthe total Poynting vector from Eq. (9), while the red and yellow curve are itselectromagnetic and kinetic terms, respectively, from Eq. (9).polariton, as well as significant reflection at the truncation, inconsistency with our dispersion relation analysis as shown inFig. 2. In the Supplementary Materials, we provide movies tocompare the field evolution in the local (Supplementary Movies 1and 2) and nonlocal (Supplementary Movies 3 and 4) models inthe presence of losses. Even in the presence of losses, a strongreflection into the high-K backward propagating mode is seen inthe nonlocal model. On the other hand, no such effect ofbackward propagation is visible in the local model. Thesesimulations indicate that the difference between the predictionsof the local and nonlocal models are qualitatively different even inthe presence of losses.To further highlight the contrast between the local andnonlocal models, we note that, for γ 0, the local Drude modelwould predict that within the unidirectional frequency range,there is a net energy flux toward the truncation, as ref. 18 shows.On the other hand, we compute the Poynting vector flux alongthe z-direction in the nonlocal model. The time-averagedPoynting vector S in the hydrodynamic model can be derivedby combining the linearized forms of Eqs. (2) and (3) with thePoynting theorem to obtain"#1β2 J ð JÞ :ð9ÞS ¼ Re E H þ i 22ωωp ϵ1 ϵ0We show the Poynting flux along the z-direction in Fig. 3c.The total Poynting flux (blue curve) within the hydrodynamicmodel contains contributions from both the electromagnetic fieldNATURE COMMUNICATIONS (2020)11:674 https://doi.org/10.1038/s41467-020-14504-9 www.nature.com/naturecommunications

ARTICLENATURE COMMUNICATIONS https://doi.org/10.1038/s41467-020-14504-9 (in p)0.80.60.4Re ( )0.2 0 0.025 p–102Im ( )–100100K (in kp)102Fig. 4 Dispersion relations in the real-wavevector complex-frequencypicture. The dispersion relations in the real wavevector and complexfrequency picture for the surface-plasmon polariton in the local Drudemodel (blue) is contrasted with that in the hydrodynamic model (red). Thesolid and dotted lines represent the real and imaginary parts of thefrequency, respectively, for a loss rate of γ0 0.025ωp. In the local model,the flat asymptote in the limit of K persists even upon the inclusion oflosses, resulting in an infinite number of states. By contrast, in thehydrodynamic model, the unidirectional frequency gap is closed in the realwavevector complex-frequency picture as well as the complex-wavevectorreal-frequency picture (shown in Fig. 2b).(E H*, red curve) and the kinetic energy of the free carriers(J ð JÞ, yellow curve), unlike in the local model. Since there is aPEC termination, the total Poynting vector to the right of thesource must be zero. We see that this is indeed the case, with theforward propagating electromagnetic component being canceledexactly by the counter-propagating kinetic component, an effectarising from the nonlocal term J*( J). Similarly, a negativevalue of Poynting flux is observed to the left of the source, alsoconfirming the excitation of the high-K mode propagatingbackward.General thermodynamic argument. In the local Drude model,the surface plasmon has a dispersion relation ω(K) that asymptotically approaches a constant in the limit of K , whichresults in an infinite number of states in a finite frequency range.From Fig. 1, this asymptotic behavior is apparent when the Drudemodel is lossless. However, this is also the case in the presence oflosses in the Drude model: the dispersion relation relevant tocomputing the number of states in the presence of losses isobtained by setting a real wavevector and solving for a complexfrequency39. It was shown39,40 that such a dispersion relationpresents a flat asymptote at the surface-plasmon frequency evenin the presence of losses. In Fig. 4, we plot the surface-plasmonpolariton dispersion relation in the presence of losses for a realwavevector and complex frequency for the local Drude model inblue, with the real part of the frequency shown by the solidcurve and the imaginary part by the dotted curve. Note that forthe local model, the flat asymptotes persist even in the presence oflosses in the real-wavevector complex-frequency picture. Inmetal–dielectric systems, this asymptotic behavior is key to theexistence of the unidirectional frequency range when a magneticfield is applied. However, since the asymptotic behavior alsoimplies an infinite number of states in a finite frequency range,the thermal energy contained in the system diverges to infinity atany nonzero temperature. Any physical system should not haveinfinite thermal electromagnetic energy density at a finite temperature. Thus, the dispersion relation of the local Drude modeland the resulting unidirectional frequency range are not physical.Further, our prediction that a unidirectional frequency range doesnot arise when a more realistic nonlocal model is used shouldtherefore hold true independent of the details of the nonlocalmodel, since any valid nonlocal correction must remove thediverging number of states in the local Drude model and therebyremove the asymptotic behavior. As an example, the nonlocalhydrodynamic model considered in this paper indeed removesthe flat asymptotic behavior in the real-wavevector complex-frequency picture, shown by the red curve in Fig. 4.Related to the general thermodynamic argument above, it wasshown37 that the leading order correction from quantumplasmonics to the surface-plasmon polariton in the local Drudemodel has a model-independent form of ω ωsp CK for someconstant C. Moreover, it was argued that the leading order nonlocalcorrection to the dynamics of the electron gas is OðK 2 Þ41 regardlessof the microscopic model of nonlocality, with this correction alsobeing the origin of the slope of the dispersion relation for large K inFig. 2. These results concur with our observation above that theprediction of ω ωsp in the large-K limit from the local Drudemodel is unphysical, and thus any effect that relies upon suchasymptotic behavior needs to be examined carefully.DiscussionIn this paper, we show that the unidirectional frequency range,which is predicted for a metal–dielectric interface where the freeelectron metal under a static magnetic field is described using alocal Drude model, is nonphysical. We present a general argument from thermodynamic considerations and illustrate theargument with an explicit calculation using a more realisticnonlocal hydrodynamic model for the metal. Our results heresuggest that the anomalous time-bandwidth product predicted byref. 18, which arises as a direct consequence of the existence of theunidirectional frequency range, is not physical either. Moregenerally, our work highlights the importance of using a morerealistic permittivity model, such as those derived from quantumplasmonics considerations27–29, to understand nonreciprocalplasmonic effects.MethodsNumerical simulation. The field and Poynting flux in Fig. 3b, c were obtained bysolving Eqs. (5) and (6) using the finite-difference frequency-domain method38.Data availabilityThe data that support the findings of this study are available from the correspondingauthor upon reasonable request.Received: 26 September 2019; Accepted: 8 January 2020;References1.2.3.4.5.6.7.Maier, S. A. & Atwater, H. A. Plasmonics: localization and guiding ofelectromagnetic energy in metal/dielectric structures. J. Appl. Phys. 98, 10(2005).Ozbay, E. Plasmonics: merging photonics and electronics at nanoscaledimensions. Science 311, 189–193 (2006).Schuller, J. A. et al. Plasmonics for extreme light concentration andmanipulation. Nat. Mater. 9, 193–204 (2010).Jalas, D. et al. What is - and what is not - an optical isolator. Nat. Photo. 7,579–582 (2013).Belotelov, V., Doskolovich, L. & Zvezdin, A. Extraordinary magneto-opticaleffects and transmission through metal-dielectric plasmonic systems. Phys.Rev. Lett. 98, 077401 (2007).Temnov, V. V. et al. Active magneto-plasmonics in hybrid metal-ferromagnetstructures. Nat. Photonics 4, 107–111 (2010).Chin, J. Y. et al. Nonreciprocal plasmonics enables giant enhancement of thinfilm faraday rotation. Nat. Commun. 4, 1599 (2013).NATURE COMMUNICATIONS (2020)11:674 https://doi.org/10.1038/s41467-020-14504-9 www.nature.com/naturecommunications5

23.24.25.26.27.28.29.30.31.32.33.34.35.6NATURE COMMUNICATIONS https://doi.org/10.1038/s41467-020-14504-9Davoyan, A. R. & Engheta, N. Nonreciprocal rotating power flow withinplasmonic nanostructures. Phys. Rev. Lett. 111, 047401 (2013).Seshadri, S. Excitation of surface waves on a perfectly conducting screencovered with anisotropic plasma. IRE Trans. Microw. Theory Tech. 10,573–578 (1962).Ishimaru, A. Unidirectional waves in anisotropic media. in ElectromagneticTheory and Antennas. 591–601 (Pergamon Press, New York, 1963).Silveirinha, M. G. Chern invariants for continuous media. Phys. Rev. B 92,125153 (2015).Silveirinha, M. G. Quantized angular momentum in topological opticalsystems. Nat. Commun. 10, 349 (2019).Van Mechelen, T. & Jacob, Z. Unidirectional Maxwellian spinwaves. Nanophotonics, 8, 1399–1416 (2019).Gangaraj, S. A. H. & Monticone, F. Do truly unidirectional surface plasmonpolaritons exist? Optica 6, 1158–1165 (2019).Brion, J., Wallis, R., Hartstein, A. & Burstein, E. Theory of surfacemagnetoplasmons in semiconductors. Phys. Rev. Lett. 28, 1455 (1972).Yu, Z., Veronis, G., Wang, Z. & Fan, S. One-way electromagnetic waveguideformed at the interface between a plasmonic metal under a static magneticfield and a photonic crystal. Phys. Rev. Lett. 100, 023902 (2008).Shen, L., You, Y., Wang, Z. & Deng, X. Backscattering-immune one-waysurface magnetoplasmons at terahertz frequencies. Opt. Express 23, 950–962(2015).Tsakmakidis, K. et al. Breaking Lorentz reciprocity to overcome the timebandwidth limit in physics and engineering. Science 356, 1260–1264 (2017).Yanik, M. F. & Fan, S. Stopping light all optically. Phys. Rev. Lett. 92, 083901(2004).Tsang, M. Quantum limits on the time-bandwidth product of an opticalresonator. Opt. Lett. 43, 150–153 (2018).Mann, S. A., Sounas, D. L. & Alù, A. Nonreciprocal cavities and the timebandwidth limit. Optica 6, 104–110 (2019).Agranovich, V.M. & Ginzburg, V. Crystal Optics with Spatial Dispersion,and Excitons, Vol. 42 (Springer Science and Business Media, 2013).Raza, S., Christensen, T., Wubs, M., Bozhevolnyi, S. I. & Mortensen, N. A.Nonlocal response in thin-film waveguides: loss versus nonlocality andbreaking of complementarity. Phys. Rev. B 88, 115401 (2013).Ciracì, C. et al. Probing the ultimate limits of plasmonic enhancement. Science337, 1072–1074 (2012).Feibelman, P. J. Microscopic calculation of surface-plasmon dispersion anddamping. Phys. Rev. B 9, 5077 (1974).Garcia de Abajo, F. J. Nonlocal effects in the plasmons of strongly interactingnanoparticles, dimers, and waveguides. J. Phys. Chem. C 112, 17983–17987(2008).Esteban, R., Borisov, A. G., Nordlander, P. & Aizpurua, J. Bridging quantumand classical plasmonics with a quantum-corrected model. Nat. Commun. 3,825 (2012).Tame, M. S. et al. Quantum plasmonics. Nat. Phys. 9, 329 (2013).Jacob, Z. & Shalaev, V. M. Plasmonics goes quantum. Science 334, 463–464(2011).Raza, S., Bozhevolnyi, S. I., Wubs, M. & Mortensen, N. A. Nonlocal opticalresponse in metallic nanostructures. J. Phys. 27, 183204 (2015).Maack, J. R., Mortensen, N. A. & Wubs, M. Size-dependent nonlocal effects inplasmonic semiconductor particles. Europhys. Lett 119, 17003 (2017).Pitarke, J., Silkin, V., Chulkov, E. & Echenique, P. Theory of surface plasmonsand surface-plasmon polaritons. Rep. Prog. Phys. 70, (2006).Aers, G. & Boardman, A. The theory of semiconductor magnetoplasmonpolariton surface modes: Voigt geometry. J. Phys. C 11, 945 (1978).Sarma, S. D. & Quinn, J. Collective excitations in semiconductor superlattices.Phys. Rev. B 25, 7603 (1982).Halevi, P. Hydrodynamic model for the degenerate free-electron gas:generalization to arbitrary frequencies. Phys. Rev. B 51, 7497 (1995)

propagating surface-plasmon polaritons at metal-dielectric inter-faces. Within the local Drude model, the existence of the uni-directional frequency range depends on the behavior of the model in the limit of large wavevectors. On the other hand, it has long been known22 that nonlocal effects become important inthis limit.

Related Documents:

Absence (ESS): Absence Management ABSENCE ESS_EMPLOYEE ABSENCE MANAGEMENT_JA 5/12/2017 PAGE 1 OF 13 Absence Management for Employees For All Employees eligible for some form of paid time off Intro Purpose: To submit, edit or delete an absence request in Bear Trax. How to Access: From the Navigator, under "Me" select "Personal .

Frontline Education's Absence Management Solution Reference Guide for Administrators Table of Contents Internet Feature Section A: Accessing the System Section B: Entering an Absence Section C: Modifying an Absence Section D: Splitting an Absence Section E: Cancelling an Absence Section F: Removing a Substitute from an Absence Section G: Viewing and Editing Preferred Substitutes

Frontline Education's Absence Management Solution Reference Guide for Administrators Table of Contents Internet Feature Section A: Accessing the System Section B: Entering an Absence Section C: Modifying an Absence Section D: Viewing and Editing Preferred Substitutes Section E: Viewing the Daily Report Section F: Viewing the Day of Week Absence Analysis

PAC Absence Entry - Training Guide Visit the PAC Time and Absence page for more information on HR related policies, processes and training. PAC Absence Entry Absence Only Reporters and Timesheet Reporters enter their vacation, sick, personal and other types of absences in through Employee Self Service in PAC.

5 Number of Occasions Score A % absence rate Score B 1 – 3 occasions in a rolling 12 month period 1 2% absence in a rolling 12 month period 1 4 – 6 occasions in a rolling 12 month period 2 2 – 3.9 % absence in a rolling 12 month period 2 7 – 9 occasions in a rolling 12 month period 3 4 – 5.9 % absence in a rolling 12 month period 3 10 - 15 more occasions in a rolling 12 month

Penn State Human Resources Absence Management Department The 331 Building, Suite 136 University Park, PA 16802 Fax: 814-863-6227 Email: absence@psu.edu Absence Management Penn State Human Resources The 331 Building, Suite 136 University Park, PA 16802

Mar 16, 2012 · The purpose of the Absence and Leave Departmental Handbook is to set Departmental absence and leave procedures based upon leave laws, regulations of the Office of Personnel Management and Comptroller General Decisions. The Absence and Leave Departmental Handbook will be an accompanying

in pile foundations for Level 1 earthquake situation. The proposed load factors in the study are a function of the chosen soil investigation/testing and piling method, which is applied to the bending moment in piles. Therefore, better choices of soil investigation/testing and high quality piling method will result in more reasonable design results. Introduction Reliability-based design .