Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Strong magneto-optical response enabled by quantum two-level systems

Open Access Open Access

Abstract

The magneto-optical effect breaks time-reversal symmetry, a unique property that makes it indispensable in nonreciprocal optics and topological photonics. Unfortunately, all natural materials have a rather weak magneto-optical response in the optical frequency range, posing a significant challenge to the practical application of many emerging device concepts. Here, we theoretically propose a composite material system that exhibits an intrinsic magneto-optical response orders of magnitude stronger than most magneto-optical materials used today. This is achieved by tailoring the resonant interplay between the quantum electrodynamics of electronic transitions in two-level systems and the classical electromagnetic response of local plasmon resonance.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. INTRODUCTION

Breaking Lorentz reciprocity using the magneto-optical (MO) effect opens the door to a new class of functionality unattainable in reciprocal systems. It directly contributes to the rise of fields of nonreciprocal [18] and topological photonics [918]. While many exciting optical sciences are being discovered, such as optical Weyl points [17,19], quantum Hall effect [13,2027], etc., there has always been a challenge: nature does not offer any material with a strong MO effect in the optical frequency range [28]. In nonreciprocal photonics, the weak MO effect has been the roadblock in miniaturization of optical isolators, holding back the development of fully integrated Si photonics. In emerging topological photonics, demonstration of nontrivial topological states has rarely gone beyond radio frequency [13,29], again because the MO effect of natural materials tails off when transitioning from the radio to the optical frequencies. The weak MO response in natural materials prompted the search for other ways of breaking reciprocity, including, e.g., nonlinearity [3035] and time-dependent modulation [2,68,36]. However, these methods come with their own limitations. The performance of nonlinear methods strongly depends on the intensity of light signal and suffers from dynamical reciprocity [37]. The time-dependent modulation, on the other hand, requires active driving that is both complex and power hungry [5].

The lack of strong MO materials has also motivated extensive research on using nanostructures to enhance the MO effect. These structures slow down light to increase interaction time with the MO materials. It is achieved by a variety of resonant structures [3854], including ring resonators, photonic crystal slabs, and optical nanoresonators. However, these wavelength-scale structural features substantially limit its general applicability in complex geometries. For example, it is unclear how a ring resonator can be made compatible with three-dimensional Weyl photonic crystals [17,55], which come with their own structural specifications.

An ideal enhancement mechanism should use intrinsic material properties instead of extrinsic structures that could impede general usability. Here, we theoretically show that a quantum-classical composite material could exhibit an intrinsically large MO response in the optical frequencies. It uses quantum electronic transition in two-level systems (TLSs)—such as atoms, molecules, and quantum dots—to provide the intrinsic MO response. Under an external magnetic field, Zeeman splitting lifts the degenerate optical transitions for left and right circularly polarized light. When the split is greater than the spectral width of individual transitions, left and right circularly polarized transitions no longer spectrally overlap, leading to a perfect MO response. This effect can be readily achieved in isolated quantum systems such as TLSs, but not in bulk materials, where a large number of states are densely packed in the energy spectrum and overlapping transitions cancel MO response despite Zeeman splitting. Based on magnetized TLS, we further explore ultra-compact local plasmonic resonance to broaden the bandwidth of enhanced MO effect. With a strong intrinsic MO response, the material offers great design flexibility in breaking the Lorentz nonreciprocity in the optical frequency range.

2. MAIN RESULTS

The MO strength of a material can be described by the imaginary off-diagonal component of polarizability. Even for some of the strongest MO materials such as yttrium iron garnet (YIG), the MO response, i.e., the off-diagonal component of the polarizability, is at least three orders of magnitude weaker than the non-MO electromagnetic response, i.e., the diagonal components. On the other hand, Zeeman splitting of electronic transition is well known to induce a strong MO effect [5658]. When a magnetic field splits the excited state of a TLS, the polarizability tensor at the transition frequency σ+ is (details in Section III.C of Supplement 1)

αTLS(0)=α02(1i0i10000),
where α0=3iϵ0λ3/4π2. ϵ0 is vacuum permittivity, and λ is the wavelength of the resonant transition. The polarizability exhibits a perfect MO response: the imaginary off-diagonal component has the same magnitude as the diagonal component.

Figure 1 compares MO materials and a magnetized TLS. Unlike bulk materials, where we can use the Faraday rotation angle to measure the MO strength, here we instead use the polarization of the forward scattered light to characterize the strength of the MO response for a single subwavelength object. Specifically, we consider an incident wave linearly polarized along the x direction. When the material has no MO effect [Fig. 1(a)], the induced polarization p is linearly polarized along the x direction, resulting in linearly polarized scattered light in the forward direction. For an MO material, e.g., YIG with ϵxy=0.06i, the induced polarization p acquires a small component in the y direction with a phase difference of π/2 relative to that in the x direction. This small y component leads to elliptically polarized light, as shown in Fig. 1(b). Because of the weak MO strength, the polarization is highly elongated along the x direction. If YIG’s ϵxy were to increase to 3i, the scattered light would become more circular, as shown in Fig. 1(c). These conclusions are independent of the size of the particles, as long as they are well below the wavelength. The strong MO effect of a magnetized TLS can be seen in Fig. 1(d), where σ± represent transitions coupled to clockwise and anti-clockwise circularly polarized light, respectively. The scattered fields are circularly polarized, indicating a perfect magneto-optical response.

 figure: Fig. 1.

Fig. 1. Schematics of scattered fields and polarizations of forward scattered fields for MO materials with diagonal element ϵxx=6.25 and off-diagonal dielectric constants (a) ϵxy=0, (b) ϵxy=0.06i, and (c) ϵxy=3i; and (d) magnetized TLS. We note that at the deep sub-wavelength scale, the shape and size of the object do not have a significant impact on the results shown above (see confirmation in Section I.A of Supplement 1).

Download Full Size | PDF

However, TLS’s strong intrinsic MO response suffers from one critical issue: the incident frequency must be around one of the transitions σ±, as shown by the inset in Fig. 1(d). Unfortunately, these transitions typically have a very narrow spectral bandwidth γ0. For example, the bandwidth of the D1 transition (5S21/25P21/2) of an Rb87 atom is only 36.1 MHz [59]. To make this strong MO response relevant for optical applications, the bandwidth must be increased by orders of magnitude.

The fundamental limit of bandwidth comes from the small size of the radiating dipoles in TLSs. The relationship between bandwidth and dipole size is given by γ0μ2 [60]. For the D1 transition described above, the effective size of the transition dipole momentum is only 0.08 nm, a rather small size compared to most classical radiators at optical frequencies. Unfortunately, it is rare for a TLS to have a large dipole, due to its small physical sizes. One cannot solve this issue by using super-large TLSs or even bulk materials. As shown in Fig. 2(a), the magnetic field splits otherwise degenerate are states. It is essential to have these non-overlapping transitions for clockwise (orange) and counterclockwise (blue) polarized light in order to create MO response. In large systems or bulk materials, the transitions are often densely packed in the energy spectrum. Although the magnetic field can still split the degenerate states, it cannot easily create non-overlapping transition for clockwise and counterclockwise polarized light, as illustrated in Fig. 2(b). These overlapping transitions make it difficult to realize strong MO response [61] (see Section II in Supplement 1 for further explanation).

 figure: Fig. 2.

Fig. 2. Schematics of σ± transition probabilities for a TLS (a) with and (b) without a magnetic field. (c) and (d) correspond to the bulk system with dense packed σ± transitions with same magnetized state as (a) and (b). When energy levels overlap as shown in (d), the MO response is weak despite the magnetic field.

Download Full Size | PDF

In classical electromagnetics, the dipole moment may not always be limited by small physical size. For example, considering a spherical nanoparticle with a subwavelength size (e.g., 5 nm), the induced dipole moment is [62,63] p3VϵNP1ϵNP+2ϵ0E0, where V and ϵNP are the volume and the permittivity of the nanoparticle, respectively, and E0 is the incident wave. The dipole moment diverges when ϵNP=2, resulting in an effective dipole of infinite size. This effect contributes to local plasmonic resonances that are widely used for sensing [64,65]. In practice, the dipole moment will not diverge due to the loss in real materials. Using this effect, we show a quantum-classical composite system [6674] that can exhibit a magnetized dipole of greatly enhanced size. Figure 3(a) shows one example of such a composite system: it consists of a magnetized TLS placed near a non-magnetic nanoparticle. The distance between the TLS and the nanoparticle is relatively short (<20nm).

 figure: Fig. 3.

Fig. 3. (a) A magnetized TLS (blue dot) placed below a non-magnetic gold nanoparticle of 5 nm radius (details in main text). The distance between them is 8 nm. Black arrows denote the local polarization currents. The red circle shows the polarization of the forward scattering field under a linearly polarized incident wave. (b) Effective broadening of the bandwidth of the σ+ transition as a function of transition wavelength. The bandwidth enhancement is defined by Γ/γ0. (c) and (d) Poynting flux lines around the magnetized TLS-nanoparticle (gold) composite. The frequency of the incident light is tuned to be the frequency of transition σ+, i.e., ω=ω0ΔωB/2. Great contrast can be seen in the optical cross section when under different polarizations, even at the same frequency. All calculations are based on exact method in Eqs. (2) and (3).

Download Full Size | PDF

First, we provide an intuitive understanding of how the composite system helps to increase the effective size of the dipole in a magnetized TLS, and consequently, its bandwidth. We again consider the forward scattering field under a linearly polarized incident wave. The magnetized TLS produces a very strong scattered field around itself. The field as felt by a nearby nanoparticle is thus dominated by this scattered field, which is circularly polarized. As a result, the induced polarization in the nanoparticle is circular, instead of following the incident field to create a linear polarization. Figure 3(a) illustrates the induced polarization current of the nanoparticle. This circulating current produces a circularly polarized scattered field, which further contributes to the MO response of the system. It effectively increases the size of the rotating dipole of the magnetized TLS, leading to a much stronger coupling to the radiation continuum, and thus a broader radiative bandwidth, than a bare TLS alone.

Another perspective to understand bandwidth broadening is the Purcell factor calculated as [71,75] F=Γγ0=μ·ImG(rTLS,rTLS)·μμ·ImG0(rTLS,rTLS)·μ, where G=G0+Gs is the dyadic Green functions [76]. G0,s are the Green’s function in free space and the scattered Green’s function due to the nanoparticle, respectively. The bandwidth enhancement diverges when the permittivity ϵNP2. For real materials, such as gold, the optical loss curtails the divergence but still produces significant bandwidth enhancement. An example of a gold nanoparticle is shown in Figs. 3(a) and 3(b). From the Purcell factor alone, one cannot conclude whether the composite system’s response is dominated by the TLS’s MO response or the non-magnetic response of the nanoparticle. The modified decay rate Γ includes both radiative and non-radiative parts.

Next, we perform quantum electrodynamic modeling of the composite system to explicitly show the MO response. When a plane wave is incident upon a composite TLS-nanoparticle system, the expectation value of the total field operator E^(r) can be written as [74]

E^(r)=E0(r)+E0s(r)+E^TLS(r),
where E0 represents the incident plane wave, and E0s is the scattered field of the nanoparticle in the absence of the TLS. The scattered field operator of the TLS is given by E^TLS(r)=(ω2/ϵ0c2)G(r,rTLS,ω)·d, where the induced dipole moment d can be solved by a Bloch equation under weak excitation [71,74] (also see details in Section III.A of Supplement 1):
d(ω)=1mμm·(E0+E0s)ωωm+δωm+iΓm/2μm=αTLS(eff)·E0,
where m=± denote the transitions σ±, and ω is the frequency of the incident wave. The effective TLS polarizability αTLS(eff) has been dressed by the nanoparticle (see Section III.D in Supplement 1). The transition frequency ωm+δωm includes the Lamb shift δωm=(ω2/ϵ02)μm·ReK(rTLS,rTLS,ω)·μm. The classical electromagnetic response of the nanoparticle is included in the Green’s function [67,72] K(r,r,ω)=G(r,r,ω)(c2/ϵ0ω2)δ(rr)I.

Under a linearly polarized incident wave, we calculate the polarization of the scattered field in the forward direction for the cases studied in Fig. 3(a). The polarization is almost circular, indicating strong MO response. The induced current calculated on the surface of the nanoparticle is indeed circulating, as shown in Fig. 3(a). The circular polarization of a TLS near a nanoparticle can be described by the degree of circular polarization (DOCP), which levels off to +100% at transition frequency σ+ (see details in Section III.D of Supplement 1). It indicates the polarization of this modified TLS is perfectly circular. The strong MO response can also be seen under circularly polarized incidence. At the σ+ transition frequency, the clockwise polarization strongly excites the composite, while the other, orthogonal polarization barely interacts with the composite [as shown as the Poynting flux flows in Figs. 3(c) and 3(d)].

Next, we consider a TLS with a specific radiative linewidth γ0=1GHz [77] around 550 nm wavelength. To allow some spacing between the TLS and the gold particle, we embed the TLS in a uniform background material with a dielectric constant ϵb=2.25. The permittivity of gold nanoparticle ϵNP is given by the Drude model with non-local effect and Landau damping (Section III.F in Supplement 1). In general, as the TLS-nanoparticle spacing decreases, the bandwidth broadens, while the peak value of the off-diagonal polarizability decreases. The spacing is 8 nm, and the radius of the gold particle is 5 nm. In practice, such structures can be produced using core-shell structures [7880]. We apply a finite magnetic field corresponding to a Zeeman splitting of ΔωB=200γ0. Here, we use full-wave electromagnetic modeling in Eqs. (2) and (3). The Mie scattering field E0s and the dyadic Green’s function G(r,r,ω) can be numerically calculated [76,81]. Figure 4(a) shows the off-diagonal component of the polarizability αxy with and without the nanoparticle. Two peaks correspond to the two Zeeman-splitting transitions. The ratio |αxy|/|αxx| remains close to 1 at resonant frequencies of transitions σ±, because the bandwidth broadening is less than the Zeeman splitting and there is almost no overlap between these two transitions. The nanoparticle broadens the bandwidth by 40 times, and the bandwidth broadening also varies with the direction of incident light. Here, we obtain greater enhancement in Fig. 4(a) than that in Fig. 3(b) because the incident direction is normal to the axis that connects the TLS and the center of the nanoparticle.

 figure: Fig. 4.

Fig. 4. (a) Off-diagonal component of the polarizability |αxy| with (blue) and without (gray) nanoparticle (NP). A TLS is at left side of nanoparticle with same distance and radius as Fig. 3(a). (b) Inset shows a quantum-classical composite system with multiple TLSs. The blue circle and red curves correspond to the off-diagonal component of polarizability |αxy| for 1 TLS with frequency distribution over a bandwidth of Δω=200γ0 and 25 TLSs with Δω=1250γ0, respectively. Due to randomization and averaging, the split peaks as shown in (a) disappear. (c) Faraday rotation of linearly polarized light passing through a disk filled with magnetized TLS-nanoparticle composites. Inset shows the zoom-in view of the glass material filled with composites. (d) and (e) Polarizations of outgoing fields for the composite and conventional YIG (ϵxx=6.25 and ϵxy=0.06i) of the same geometry, respectively. (f) Faraday rotation angles of the composite and YIG as a function of the frequency. Upper and lower red circles correspond to (d) and (e), respectively. The Zeeman splitting energy is fixed at 200γ0.

Download Full Size | PDF

The enhanced bandwidth of 40 GHz is quite useful for communication applications. It would be important to further enhance the bandwidth to the THz range to enable broader applications. This cannot be achieved by reducing the spacing between the TLS and the nanoparticle, as the MO response will be weakened by the strong ohmic loss in metal.

Next, we use multiple TLSs to further broaden the bandwidth by at least one order of magnitude. Specifically, a cluster of TLSs is randomly placed on a sphere centered around the nanoparticle, as shown by the inset in Fig. 4(b). The transition frequencies of these TLSs are not necessarily identical, and are randomly distributed in a spectral range Δω. Because of this frequency distribution, multiple transitions at different frequencies enable further enhancement of the MO bandwidth. As a result, it can be greater than the Zeeman splitting. We perform quantum electrodynamic modeling that fully incorporates the electrodynamic interactions among all components in the composite system with the radiative correction to the dipole approximation for nanoparticles; the method is shown in Section III.F of Supplement 1. We compare the MO response for the cases of one and 25 TLSs, which are shown by the blue and red curves in Fig. 4(b), respectively. The spectral bandwidth for the ensemble case is defined as the area of the off-diagonal polarizability spectrum divided by its maximum value |αxy(ω)|dω/max|αxy(ω)|, and it reaches 1300γ0 for the 25 TLSs case, which is equivalent to 1.3 THz for TLSs. Over this broadened bandwidth, the ratio between the diagonal and off-diagonal polarizability remains high at around 1/2. The polarizability tensor of multiple TLSs is evaluated by αi=1NTLSαTLS(i), where NTLS is the number of TLSs. αTLS(i) is the dressed polarizability of the ith TLS, and it is solved by considering all radiative interactions (see details in Section III.F of Supplement 1). Such a direct sum is a reasonable estimate due to the deep subwavelength scale of the composite. Given the randomness of the location and transition frequency of the TLSs, we average the result for 1000 random simulations, each having a different random configuration. We also note that without the nanoparticle, a cluster of magnetized TLSs alone would still exhibit extremely narrow bandwidths.

The TLS-nanoparticle composite shown here has a typical radius of around 10 nm, which is small enough that a large number of composite units can be incorporated into a bulk material. It can then be fabricated into a complex structure with little geometrical constraint. So far, we have analyzed only a single composite. To support the above claim, we now provide an example of such a bulk material with greatly enhanced MO response.

We consider a glass material with a refractive index of 1.5. To enable a strong MO response in this glass, we embed magnetized TLS-nanoparticle composites. Each composite has a 5 nm gold nanoparticle and 10 TLSs that are 8 nm away from the nanoparticle surface. We perform full electrodynamic simulation of the material. This is a computationally expensive simulation, considering that we model the response of each TLS and nanoparticle without using any effective medium assumption. Accordingly, we must limit the size of the material to make the computation feasible.

Here, we consider a disk with a radius of 550 nm and height of 45 nm, filled with about 300 composites randomly distributed [Fig. 4(c)]. In practice, we can use more composites and more TLSs to further increase the MO strength. We consider the Faraday rotation of a Gaussian beam passing through the material (details of modeling shown in Section V of Supplement 1). The beam wrist is half the wavelength, so the beam can maximally fit onto the disk. The Zeeman splitting caused by external magnetic field is ΔωB=200γ0, as before. Light transmitted through the disk is generally elliptically polarized. The Faraday rotation angle is estimated as θF=(1/2)atan[2Re(Ey/Ex)/(1|Ey/Ex|2)], where Ex,y are the electrical fields of the transmitted light in the x and y directions, respectively. Despite the ultra-thin thickness of the disk, the Faraday rotation reaches 30 deg, as shown in Fig. 4(b). In contrast, for a YIG disk of the same size, the Faraday rotation angle is 0.5 deg [Figs. 4(d)4(f)]. The Faraday rotation of the glass remains almost two orders of magnitude greater than that of the YIG material for a spectral bandwidth of 0.1 nm. Here, the Faraday rotation through the YIG disk is performed by using finite-element methods to solve the Maxwell’s equations.

3. DISCUSSION

Magnetized TLS-nanoparticles can realize an intrinsic MO response that is two orders of magnitude stronger than that of natural materials such as YIG. In contrast, all existing enhancement methods do not modify any fundamental property of a material. Their reliance on extrinsic structure features substantially limits general usability. The magnetized TLS-nanoparticle is also within the reach of experimental realization. For example, CdSe/CdTe quantum dots can be easily fabricated [82,83] and their transition frequency tuned by size. The Zeeman splitting energy of semiconductor quantum dots caused by the spin splitting has been expressed as ΔωB=gμBB/ [84], where g is the Landé g-factor. The Zeeman splitting in our calculation ΔωB=200γ0 corresponds to an external magnetic field B=0.56T. In practice, many TLSs have a cross section well below 3λ2/2π. We show in Section III.E of Supplement 1 that the current on the nanoparticle can still be circularly polarized. In addition, it is quite feasible to fabricate nanoparticles that are attached with quantum dots. Examples can be seen in Ref. [69].

Due to computation power limitations, we show only an example with operation at around 550 nm wavelength and a bandwidth of 0.1 nm. However, in practice, both the operation wavelength and bandwidth can be greatly extended. The operation wavelength is determined by the local plasmon resonance of the spherical gold particle. By tuning core-shell structures, the local plasmon resonance can be realized in the full optical (400–800 nm) wavelength range [7880].

It is also straightforward to broaden the bandwidth to tens or even hundreds of nanometers. To do so, we can use the inhomogeneous broadening of the TLS-nanoparticle composites: the radii of the nanoparticles and the sizes of the TLSs can be made to distribute over a broad spectral bandwidth. In addition, the spectral bandwidth also scales linearly with the number of TLSs, providing another way to enhance the bandwidth (see details in Section IV of Supplement 1). The optical loss of the composite material is contributed mainly by the absorption by nanoparticles. The example shown in Fig. 4 has 20% optical absorption with a Faraday rotation angle of 20 deg. This is remarkably good, considering that the disk is only 45 nm thick. The loss is small even compared to the all-dielectric on-chip isolator [4]. The loss can also be tuned by the density of the composite in the hosting material. Lastly, we note that the MO response has been reported for a noble metal nanoparticle coated with transition metal [85,86]. Its MO response is two orders of magnitude weaker than the magnetized TLS-nanoparticle shown here (details in Section I.B of Supplement 1).

Funding

National Science Foundation (NSF) (EFRI NewLAW Award 1641109); Defense Advanced Research Projects Agency (DARPA) (DETECT program).

 

See Supplement 1 for supporting content.

REFERENCES

1. Z. Yu, Z. Wang, and S. Fan, “One-way total reflection with one-dimensional magneto-optical photonic crystals,” Appl. Phys. Lett. 90, 121133 (2007). [CrossRef]  

2. Z. Yu and S. Fan, “Complete optical isolation created by indirect interband photonic transitions,” Nat. Photonics 3, 91–94 (2009). [CrossRef]  

3. A. B. Khanikaev, S. H. Mousavi, G. Shvets, and Y. S. Kivshar, “One-way extraordinary optical transmission and nonreciprocal spoof plasmons,” Phys. Rev. Lett. 105, 126804 (2010). [CrossRef]  

4. L. Bi, J. Hu, P. Jiang, D. H. Kim, G. F. Dionne, L. C. Kimerling, and C. Ross, “On-chip optical isolation in monolithically integrated non-reciprocal optical resonators,” Nat. Photonics 5, 758–762 (2011). [CrossRef]  

5. H. Lira, Z. Yu, S. Fan, and M. Lipson, “Electrically driven nonreciprocity induced by interband photonic transition on a silicon chip,” Phys. Rev. Lett. 109, 033901 (2012). [CrossRef]  

6. D. L. Sounas, C. Caloz, and A. Alù, “Giant non-reciprocity at the subwavelength scale using angular momentum-biased metamaterials,” Nat. Commun. 4, 2407 (2013). [CrossRef]  

7. N. A. Estep, D. L. Sounas, J. Soric, and A. Alù, “Magnetic-free non-reciprocity and isolation based on parametrically modulated coupled-resonator loops,” Nat. Phys. 10, 923–927 (2014). [CrossRef]  

8. F. Ruesink, M.-A. Miri, A. Alù, and E. Verhagen, “Nonreciprocity and magnetic-free isolation based on optomechanical interactions,” Nat. Commun. 7, 13662 (2016). [CrossRef]  

9. Z. Wang, Y. Chong, J. D. Joannopoulos, and M. Soljačić, “Reflection-free one-way edge modes in a gyromagnetic photonic crystal,” Phys. Rev. Lett. 100, 013905 (2008). [CrossRef]  

10. A. B. Khanikaev, S. H. Mousavi, W.-K. Tse, M. Kargarian, A. H. MacDonald, and G. Shvets, “Photonic topological insulators,” Nat. Mater. 12, 233–239 (2013). [CrossRef]  

11. L. Lu, J. D. Joannopoulos, and M. Soljačić, “Topological photonics,” Nat. Photonics 8, 821–829 (2014). [CrossRef]  

12. B. Zhen, C. W. Hsu, L. Lu, A. D. Stone, and M. Soljačić, “Topological nature of optical bound states in the continuum,” Phys. Rev. Lett. 113, 257401 (2014). [CrossRef]  

13. S. A. Skirlo, L. Lu, Y. Igarashi, Q. Yan, J. Joannopoulos, and M. Soljačić, “Experimental observation of large Chern numbers in photonic crystals,” Phys. Rev. Lett. 115, 253901 (2015). [CrossRef]  

14. L. Lu, C. Fang, L. Fu, S. G. Johnson, J. D. Joannopoulos, and M. Soljačić, “Symmetry-protected topological photonic crystal in three dimensions,” Nat. Phys. 12, 337–340 (2016). [CrossRef]  

15. W. Gao, M. Lawrence, B. Yang, F. Liu, F. Fang, B. Béri, J. Li, and S. Zhang, “Topological photonic phase in chiral hyperbolic metamaterials,” Phys. Rev. Lett. 114, 037402 (2015). [CrossRef]  

16. B. Yang, Q. Guo, B. Tremain, L. E. Barr, W. Gao, H. Liu, B. Béri, Y. Xiang, D. Fan, A. P. Hibbins, and S. Zhang, “Direct observation of topological surface-state arcs in photonic metamaterials,” Nat. Commun. 8, 97 (2017). [CrossRef]  

17. L. Lu, L. Fu, J. D. Joannopoulos, and M. Soljačić, “Weyl points and line nodes in gyroid photonic crystals,” Nat. Photonics 7, 294–299 (2013). [CrossRef]  

18. B. Bahari, A. Ndao, F. Vallini, A. El Amili, Y. Fainman, and B. Kanté, “Nonreciprocal lasing in topological cavities of arbitrary geometries,” Science 358, 636–640 (2017). [CrossRef]  

19. J. Noh, S. Huang, D. Leykam, Y. Chong, K. P. Chen, and M. C. Rechtsman, “Experimental observation of optical Weyl points and Fermi arc-like surface states,” Nat. Phys. 13, 611–617 (2017). [CrossRef]  

20. F. Haldane and S. Raghu, “Possible realization of directional optical waveguides in photonic crystals with broken time-reversal symmetry,” Phys. Rev. Lett. 100, 013904 (2008). [CrossRef]  

21. S. Raghu and F. Haldane, “Analogs of quantum-Hall-effect edge states in photonic crystals,” Phys. Rev. A 78, 033834 (2008). [CrossRef]  

22. M. Hafezi, E. A. Demler, M. D. Lukin, and J. M. Taylor, “Robust optical delay lines with topological protection,” Nat. Phys. 7, 907–912 (2011). [CrossRef]  

23. M. Hafezi, S. Mittal, J. Fan, A. Migdall, and J. Taylor, “Imaging topological edge states in silicon photonics,” Nat. Photonics 7, 1001–1005 (2013). [CrossRef]  

24. K. Y. Bliokh, D. Smirnova, and F. Nori, “Quantum spin Hall effect of light,” Science 348, 1448–1451 (2015). [CrossRef]  

25. V. Peano, C. Brendel, M. Schmidt, and F. Marquardt, “Topological phases of sound and light,” Phys. Rev. X 5, 031011 (2015). [CrossRef]  

26. X. Cheng, C. Jouvaud, X. Ni, S. H. Mousavi, A. Z. Genack, and A. B. Khanikaev, “Robust reconfigurable electromagnetic pathways within a photonic topological insulator,” Nat. Mater. 15, 542–548 (2016). [CrossRef]  

27. O. Zilberberg, S. Huang, J. Guglielmon, M. Wang, K. P. Chen, Y. E. Kraus, and M. C. Rechtsman, “Photonic topological boundary pumping as a probe of 4D quantum Hall physics,” Nature 553, 59–62 (2018). [CrossRef]  

28. S. Wittekoek, T. J. Popma, J. Robertson, and P. Bongers, “Magneto-optic spectra and the dielectric tensor elements of bismuth-substituted iron garnets at photon energies between 2.2–5.2 eV,” Phys. Rev. B 12, 2777–2788 (1975). [CrossRef]  

29. Z. Wang, Y. Chong, J. Joannopoulos, and M. Soljačić, “Observation of unidirectional backscattering-immune topological electromagnetic states,” Nature 461, 772–775 (2009). [CrossRef]  

30. A. Guo, G. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V. Aimez, G. Siviloglou, and D. Christodoulides, “Observation of PT-symmetry breaking in complex optical potentials,” Phys. Rev. Lett. 103, 093902 (2009). [CrossRef]  

31. H. Ramezani, T. Kottos, R. El-Ganainy, and D. N. Christodoulides, “Unidirectional nonlinear PT-symmetric optical structures,” Phys. Rev. A 82, 043803 (2010). [CrossRef]  

32. L. Feng, M. Ayache, J. Huang, Y.-L. Xu, M.-H. Lu, Y.-F. Chen, Y. Fainman, and A. Scherer, “Nonreciprocal light propagation in a silicon photonic circuit,” Science 333, 729–733 (2011). [CrossRef]  

33. L. Fan, J. Wang, L. T. Varghese, H. Shen, B. Niu, Y. Xuan, A. M. Weiner, and M. Qi, “An all-silicon passive optical diode,” Science 335, 447–450 (2012). [CrossRef]  

34. M. C. Rechtsman, J. M. Zeuner, Y. Plotnik, Y. Lumer, D. Podolsky, F. Dreisow, S. Nolte, M. Segev, and A. Szameit, “Photonic Floquet topological insulators,” Nature 496, 196–200 (2013). [CrossRef]  

35. B. Peng, Ş. K. Özdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. Fan, F. Nori, C. M. Bender, and L. Yang, “Parity-time-symmetric whispering-gallery microcavities,” Nat. Phys. 10, 394–398 (2014). [CrossRef]  

36. K. Fang, Z. Yu, and S. Fan, “Realizing effective magnetic field for photons by controlling the phase of dynamic modulation,” Nat. Photonics 6, 782–787 (2012). [CrossRef]  

37. Y. Shi, Z. Yu, and S. Fan, “Limitations of nonlinear optical isolators due to dynamic reciprocity,” Nat. Photonics 9, 388–392 (2015). [CrossRef]  

38. V. Belotelov, L. Doskolovich, and A. Zvezdin, “Extraordinary magneto-optical effects and transmission through metal-dielectric plasmonic systems,” Phys. Rev. Lett. 98, 077401 (2007). [CrossRef]  

39. J. B. González-Díaz, A. García-Martín, G. Armelles, D. Navas, M. Vázquez, K. Nielsch, R. B. Wehrspohn, and U. Gösele, “Enhanced magneto-optics and size effects in ferromagnetic nanowire arrays,” Adv. Mater. 19, 2643–2647 (2007). [CrossRef]  

40. A. Berger, R. A. de la Osa, A. K. Suszka, M. Pancaldi, J. M. Sáiz, F. Moreno, H. P. Oepen, and P. Vavassori, “Enhanced magneto-optical edge excitation in nanoscale magnetic disks,” Phys. Rev. Lett. 115, 187403 (2015). [CrossRef]  

41. M. Kataja, T. Hakala, A. Julku, M. Huttunen, S. van Dijken, and P. Törmä, “Surface lattice resonances and magneto-optical response in magnetic nanoparticle arrays,” Nat. Commun. 6, 7072 (2015). [CrossRef]  

42. N. Maccaferri, L. Bergamini, M. Pancaldi, M. K. Schmidt, M. Kataja, S. V. Dijken, N. Zabala, J. Aizpurua, and P. Vavassori, “Anisotropic nanoantenna-based magnetoplasmonic crystals for highly enhanced and tunable magneto-optical activity,” Nano Lett. 16, 2533–2542 (2016). [CrossRef]  

43. C. Min, L. Yang, and G. Veronis, “Microcavity enhanced optical absorption in subwavelength slits,” Opt. Express 19, 26850–26858 (2011). [CrossRef]  

44. G. Armelles, J. B. González-Díaz, A. García-Martín, J. M. García-Martín, A. Cebollada, M. U. González, S. Acimovic, J. Cesario, R. Quidant, and G. Badenes, “Localized surface plasmon resonance effects on the magneto-optical activity of continuous Au/Co/Au trilayers,” Opt. Express 16, 16104–16112 (2008). [CrossRef]  

45. V. Belotelov, I. Akimov, M. Pohl, V. Kotov, S. Kasture, A. Vengurlekar, A. V. Gopal, D. Yakovlev, A. Zvezdin, and M. Bayer, “Enhanced magneto-optical effects in magnetoplasmonic crystals,” Nat. Nanotechnol. 6, 370–376 (2011). [CrossRef]  

46. J. Y. Chin, T. Steinle, T. Wehlus, D. Dregely, T. Weiss, V. I. Belotelov, B. Stritzker, and H. Giessen, “Nonreciprocal plasmonics enables giant enhancement of thin-film Faraday rotation,” Nat. Commun. 4, 1599 (2013). [CrossRef]  

47. L. E. Kreilkamp, V. I. Belotelov, J. Y. Chin, S. Neutzner, D. Dregely, T. Wehlus, I. A. Akimov, M. Bayer, B. Stritzker, and H. Giessen, “Waveguide-plasmon polaritons enhance transverse magneto-optical Kerr effect,” Phys. Rev. X 3, 041019 (2013). [CrossRef]  

48. K. Lodewijks, N. Maccaferri, T. Pakizeh, R. K. Dumas, I. Zubritskaya, J. Åkerman, P. Vavassori, and A. Dmitriev, “Magnetoplasmonic design rules for active magneto-optics,” Nano Lett. 14, 7207–7214 (2014). [CrossRef]  

49. J. Valente, J.-Y. Ou, E. Plum, I. J. Youngs, and N. I. Zheludev, “A magneto-electro-optical effect in a plasmonic nanowire material,” Nat. Commun. 6, 7021 (2015). [CrossRef]  

50. X. Luo, M. Zhou, J. Liu, T. Qiu, and Z. Yu, “Magneto-optical metamaterials with extraordinarily strong magneto-optical effect,” Appl. Phys. Lett. 108, 131104 (2016). [CrossRef]  

51. G. Armelles, B. Caballero, A. Cebollada, A. Garcia-Martin, and D. Meneses-Rodríguez, “Magnetic field modification of optical magnetic dipoles,” Nano Lett. 15, 2045–2049 (2015). [CrossRef]  

52. A. Christofi, Y. Kawaguchi, A. Alù, and A. B. Khanikaev, “Giant enhancement of Faraday rotation due to electromagnetically induced transparency in all-dielectric magneto-optical metasurfaces,” Opt. Lett. 43, 1838–1841 (2018). [CrossRef]  

53. A. Leviyev, B. Stein, A. Christofi, T. Galfsky, H. Krishnamoorthy, I. Kuskovsky, V. Menon, and A. Khanikaev, “Nonreciprocity and one-way topological transitions in hyperbolic metamaterials,” APL Photon. 2, 076103 (2017). [CrossRef]  

54. S. H. Mousavi, A. B. Khanikaev, J. Allen, M. Allen, and G. Shvets, “Gyromagnetically induced transparency of metasurfaces,” Phys. Rev. Lett. 112, 117402 (2014). [CrossRef]  

55. L. Lu, Z. Wang, D. Ye, L. Ran, L. Fu, J. D. Joannopoulos, and M. Soljačić, “Experimental observation of Weyl points,” Science 349, 622–624 (2015). [CrossRef]  

56. G. Labeyrie, C. Miniatura, and R. Kaiser, “Large Faraday rotation of resonant light in a cold atomic cloud,” Phys. Rev. A 64, 033402 (2001). [CrossRef]  

57. Y. Shen, M. Bradford, and J.-T. Shen, “Single-photon diode by exploiting the photon polarization in a waveguide,” Phys. Rev. Lett. 107, 173902 (2011). [CrossRef]  

58. K. Pandey, C. Kwong, M. Pramod, and D. Wilkowski, “Linear and nonlinear magneto-optical rotation on the narrow strontium intercombination line,” Phys. Rev. A 93, 053428 (2016). [CrossRef]  

59. D. A. Steck, “Rubidium 87 D Line Data,” (2001).

60. J. D. Jackson, “Classical Electrodynamics,” (1999).

61. P. Pershan, “Magneto-optical effects,” J. Appl. Phys. 38, 1482–1490 (1967). [CrossRef]  

62. K. L. Kelly, E. Coronado, L. L. Zhao, and G. C. Schatz, The Optical Properties of Metal Nanoparticles: the Influence of Size, Shape, and Dielectric Environment (2003).

63. P. C. Chaumet, A. Rahmani, F. de Fornel, and J.-P. Dufour, “Evanescent light scattering: the validity of the dipole approximation,” Phys. Rev. B 58, 2310–2315 (1998). [CrossRef]  

64. S. Nelson, K. S. Johnston, and S. S. Yee, “High sensitivity surface plasmon resonace sensor based on phase detection,” Sens. Actuators B 35, 187–191 (1996). [CrossRef]  

65. T. Okamoto, I. Yamaguchi, and T. Kobayashi, “Local plasmon sensor with gold colloid monolayers deposited upon glass substrates,” Opt. Lett. 25, 372–374 (2000). [CrossRef]  

66. W. Zhang, A. O. Govorov, and G. W. Bryant, “Semiconductor-metal nanoparticle molecules: hybrid excitons and the nonlinear Fano effect,” Phys. Rev. Lett. 97, 146804 (2006). [CrossRef]  

67. A. Trügler and U. Hohenester, “Strong coupling between a metallic nanoparticle and a single molecule,” Phys. Rev. B 77, 115403 (2008). [CrossRef]  

68. A. O. Govorov, “Semiconductor-metal nanoparticle molecules in a magnetic field: spin-plasmon and exciton-plasmon interactions,” Phys. Rev. B 82, 155322 (2010). [CrossRef]  

69. B. P. Khanal, A. Pandey, L. Li, Q. Lin, W. K. Bae, H. Luo, V. I. Klimov, and J. M. Pietryga, “Generalized synthesis of hybrid metal-semiconductor nanostructures tunable from the visible to the infrared,” ACS Nano 6, 3832–3840 (2012). [CrossRef]  

70. C. Van Vlack, P. T. Kristensen, and S. Hughes, “Spontaneous emission spectra and quantum light-matter interactions from a strongly coupled quantum dot metal-nanoparticle system,” Phys. Rev. B 85, 075303 (2012). [CrossRef]  

71. X.-W. Chen, V. Sandoghdar, and M. Agio, “Coherent interaction of light with a metallic structure coupled to a single quantum emitter: from superabsorption to cloaking,” Phys. Rev. Lett. 110, 153605 (2013). [CrossRef]  

72. A. González-Tudela, P. Huidobro, L. Martín-Moreno, C. Tejedor, and F. García-Vidal, “Theory of strong coupling between quantum emitters and propagating surface plasmons,” Phys. Rev. Lett. 110, 126801 (2013). [CrossRef]  

73. A. Delga, J. Feist, J. Bravo-Abad, and F. Garcia-Vidal, “Quantum emitters near a metal nanoparticle: strong coupling and quenching,” Phys. Rev. Lett. 112, 253601 (2014). [CrossRef]  

74. J. Yang, M. Perrin, and P. Lalanne, “Analytical formalism for the interaction of two-level quantum systems with metal nanoresonators,” Phys. Rev. X 5, 021008 (2015). [CrossRef]  

75. C. Girard, O. J. Martin, G. Léveque, G. C. des Francs, and A. Dereux, “Generalized Bloch equations for optical interactions in confined geometries,” Chem. Phys. Lett. 404, 44–48 (2005). [CrossRef]  

76. C.-T. Tai, Dyadic Green Functions in Electromagnetic Theory (Institute of Electrical & Electronics Engineers (IEEE), 1994).

77. S. Stobbe, J. Johansen, P. T. Kristensen, J. M. Hvam, and P. Lodahl, “Frequency dependence of the radiative decay rate of excitons in self-assembled quantum dots: experiment and theory,” Phys. Rev. B 80, 155307 (2009). [CrossRef]  

78. S. Oldenburg, R. Averitt, S. Westcott, and N. Halas, “Nanoengineering of optical resonances,” Chem. Phys. Lett. 288, 243–247 (1998). [CrossRef]  

79. Z. Liang, A. Susha, and F. Caruso, “Gold nanoparticle-based core-shell and hollow spheres and ordered assemblies thereof,” Chem. Mater. 15, 3176–3183 (2003). [CrossRef]  

80. K.-S. Lee and M. A. El-Sayed, “Gold and silver nanoparticles in sensing and imaging: sensitivity of plasmon response to size, shape, and metal composition,” J. Phys. Chem. B 110, 19220–19225 (2006). [CrossRef]  

81. Q. Fu and W. Sun, “Mie theory for light scattering by a spherical particle in an absorbing medium,” Appl. Opt. 40, 1354–1361 (2001). [CrossRef]  

82. D. J. Norris, A. Sacra, C. Murray, and M. Bawendi, “Measurement of the size dependent hole spectrum in CdSe quantum dots,” Phys. Rev. Lett. 72, 2612–2615 (1994). [CrossRef]  

83. D. J. Norris and M. Bawendi, “Measurement and assignment of the size-dependent optical spectrum in CdSe quantum dots,” Phys. Rev. B 53, 16338–16346 (1996). [CrossRef]  

84. V. Kulakovskii, G. Bacher, R. Weigand, T. Kümmell, A. Forchel, E. Borovitskaya, K. Leonardi, and D. Hommel, “Fine structure of biexciton emission in symmetric and asymmetric CdSe/ZnSe single quantum dots,” Phys. Rev. Lett. 82, 1780–1783 (1999). [CrossRef]  

85. P. K. Jain, Y. Xiao, R. Walsworth, and A. E. Cohen, “Surface plasmon resonance enhanced magneto-optics (SuPREMO): Faraday rotation enhancement in gold-coated iron oxide nanocrystals,” Nano Lett. 9, 1644–1650 (2009). [CrossRef]  

86. L. Wang, C. Clavero, Z. Huba, K. J. Carroll, E. E. Carpenter, D. Gu, and R. A. Lukaszew, “Plasmonics and enhanced magneto-optics in core-shell Co-Ag nanoparticles,” Nano Lett. 11, 1237–1240 (2011). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental document

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. Schematics of scattered fields and polarizations of forward scattered fields for MO materials with diagonal element ϵ x x = 6.25 and off-diagonal dielectric constants (a)  ϵ x y = 0 , (b)  ϵ x y = 0.06 i , and (c)  ϵ x y = 3 i ; and (d) magnetized TLS. We note that at the deep sub-wavelength scale, the shape and size of the object do not have a significant impact on the results shown above (see confirmation in Section I.A of Supplement 1).
Fig. 2.
Fig. 2. Schematics of σ ± transition probabilities for a TLS (a) with and (b) without a magnetic field. (c) and (d) correspond to the bulk system with dense packed σ ± transitions with same magnetized state as (a) and (b). When energy levels overlap as shown in (d), the MO response is weak despite the magnetic field.
Fig. 3.
Fig. 3. (a) A magnetized TLS (blue dot) placed below a non-magnetic gold nanoparticle of 5 nm radius (details in main text). The distance between them is 8 nm. Black arrows denote the local polarization currents. The red circle shows the polarization of the forward scattering field under a linearly polarized incident wave. (b) Effective broadening of the bandwidth of the σ + transition as a function of transition wavelength. The bandwidth enhancement is defined by Γ / γ 0 . (c) and (d) Poynting flux lines around the magnetized TLS-nanoparticle (gold) composite. The frequency of the incident light is tuned to be the frequency of transition σ + , i.e., ω = ω 0 Δ ω B / 2 . Great contrast can be seen in the optical cross section when under different polarizations, even at the same frequency. All calculations are based on exact method in Eqs. (2) and (3).
Fig. 4.
Fig. 4. (a) Off-diagonal component of the polarizability | α x y | with (blue) and without (gray) nanoparticle (NP). A TLS is at left side of nanoparticle with same distance and radius as Fig. 3(a). (b) Inset shows a quantum-classical composite system with multiple TLSs. The blue circle and red curves correspond to the off-diagonal component of polarizability | α x y | for 1 TLS with frequency distribution over a bandwidth of Δ ω = 200 γ 0 and 25 TLSs with Δ ω = 1250 γ 0 , respectively. Due to randomization and averaging, the split peaks as shown in (a) disappear. (c) Faraday rotation of linearly polarized light passing through a disk filled with magnetized TLS-nanoparticle composites. Inset shows the zoom-in view of the glass material filled with composites. (d) and (e) Polarizations of outgoing fields for the composite and conventional YIG ( ϵ x x = 6.25 and ϵ x y = 0.06 i ) of the same geometry, respectively. (f) Faraday rotation angles of the composite and YIG as a function of the frequency. Upper and lower red circles correspond to (d) and (e), respectively. The Zeeman splitting energy is fixed at 200 γ 0 .

Equations (3)

Equations on this page are rendered with MathJax. Learn more.

α TLS ( 0 ) = α 0 2 ( 1 i 0 i 1 0 0 0 0 ) ,
E ^ ( r ) = E 0 ( r ) + E 0 s ( r ) + E ^ TLS ( r ) ,
d ( ω ) = 1 m μ m · ( E 0 + E 0 s ) ω ω m + δ ω m + i Γ m / 2 μ m = α TLS ( eff ) · E 0 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.