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

Enhanced normal-incidence Goos-Hänchen effects induced by magnetic surface plasmons in magneto-optical metamaterials

Open Access Open Access

Abstract

Goos-Hänchen (GH) effects at normal incidence are investigated for metamaterials consisting of an array of ferrite rods. A new effective-medium approach is presented and applied to retrieve the effective parameters of the magneto-optical (MO) metamaterials based on a transformation method. Giant normal-incidence Goos-Hänchen (NIGH) shifts on total reflection and enhanced magnetic switching effects are predicted near magnetic surface plasmon (MSP) resonances for structures with small effective permittivity. Numerical simulations are performed and the results are in good agreement with those from the transformation effective-medium approach.

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

1. Introduction

The well-known GH effect is one of the most fundamental phenomena in beam optics. Early in 1947, it was discovered that a light beam of finite width will undergo a lateral shift in the incident plane when totally internally reflected from a dielectric surface. Today this effect has been extended to reflection from or transmission through different materials and structures, such as metal-dielectric nanocomposites [1,2], epsilon-near-zero metamaterials [3], graphene [4], PT-symmetric medium [5,6], topological insulator [7] and various resonant materials [8–14].

In most cases the GH shift only occurs at oblique incidence. The unique exception so far reported is the effect observed in reflection from the surface of magnetic materials near ferromagnetic resonance [8–14]. The permeability of such a magnetic material takes the MO tensor form

μ^=(μriμi0iμiμr0001)
where the z-axis is set to be along the magnetization. When an s-polarized beam (with the electric component along the z-axis) is incident upon the surface of a magnetic material, nontrivial GH shifts can be observed even at normal incidence due to the nonreciprocity in scattering coefficients originated from the broken time reversal symmetry (TRS). Both the sign and magnitude of this NIGH shift can be magnetically controlled, and giant NIGH effects are predicted around a certain frequency fc which satisfies εs=μvg [8,11], where εs denotes the permittivity of the magnetic material, μvg=μr(1Q2) is the so-called Voigt permeability and the ratio Q=μi/μr is the MO Voigt parameter which measures the MO effects of the material.

However, such an interesting NIGH shift, quite large and well tunable, is accompanied by a vanishing reflectivity at the same frequency fc. Away from fc, the reflectivity rises but the shift drops dramatically. For example, the NIGH shift on reflection from the surface of yttrium-iron-garnet (YIG), one of the most conventional MO metamaterials, follows [8]

dr0.0136ηλ0
at an applied magnetic field h0=1000Oe, where η(ffc)/fc is the relative deviation from the frequency fc and λ0 is the incident wavelength. Increasing the applied magnetic field can promote the NIGH effect but the shift enhancement is limited. Theoretical analysis has shown that when the deviation from fc is large enough so that reflectivity reaches 100%, the shift at an applied field h0 drops to a small value given by [8]
drTR=λ0πh0h0+ms
with ms denoting the saturation magnetization. In the strong-filed limit h0>>ms, the shift on total reflection drTR is bounded by a maximum value of λ0/π. Similar NIGH effects were also reported in other MO materials such as antiferromagnetic MnF2 [11].

Is it possible to realize large NIGH shifts while keeping considerable reflectivity? The strong-field limit of drTR suggests that homogenous MO materials with intrinsic ferromagnetic resonance (IFR) cannot serve as good candidates to achieve this goal. In this paper, we propose a new design of MO metamaterials where a MSP resonance with TRS-breaking is utilized as the underlying mechanism to produce large NIGH shifts even at total reflection.

The MSP phenomenon is the magnetic analog of the well-known “surface plasmon” effect in noble metals [15]. Gollub et al. [16] first introduced this concept into the design of metamaterials and develop a split-ring-resonator structure to successfully excite MSPs in a frequency band with negative effective permeability. Similar effects were also reported in subwavelength photonic crystals (PCs) made of ferrite rods [17], where the collective MSP resonance of the whole medium is a phenomenon emerging from the coupled localized-surface-plasmon-resonance (LSPR) modes of individual rods. The MSP effect in these ferrite arrays has attracted considerable interest since it is combined with the TRS breaking effect of magnetized ferrites so that a special bandgap is formed around the MSP resonance which contains optical chiral edge states along the surface of the metamaterial [18,19]. Experiments have successfully demonstrated that such an MO metamaterial working in the MSP bandgap can be utilized to fabricate omnidirectional waveguides with better performance than those based on conventional Bragg gaps of magnetic PCs [20].

By comparing MSP in ferrite arrays and IFR in homogenous ferrites, one can see that they are both magnetically-tunable resonance with broken TRS, but the former, as a metamaterial property, can be controlled by tailoring the composite structure. So it is reasonable to expect that the MSP resonance can induce the NIGH effect for a beam reflected from the surface of a metamaterial consisting of ferrite rods, just like what the IFR do for a homogeneous ferrite, and “good” structures may exist which can exhibit large NIGH shifts well beyond natural MO materials.

In previous studies, an effective-medium approximation (EMA) has been applied to retrieve the effective parameters of these MO metamaterials consisting of ferrite rods [21,22]. We note that MO-EMA method is focused on the characteristics of band structures, which is not sufficient for the discussion of the behavior of beam reflection from the effective MO medium. In nature, it is reduced to the cases for magneto-dielectric composites beyond the long-wavelength approximation [23]. So here we are motivated to present a new effective-medium theory which can be applied to obtain band structures and effective MO parameters simultaneously. The key idea is to establish a correspondence between the scattering coefficients of different orders of a MO material and those of a series of normal dielectrics. Then the problem of the concerned MO metamaterial is transformed into a series of conventional EMA models of normal dielectric. This transformation is not only helpful to obtain the effective parameters of a MO metamaterial in a much easier way, but also offers a deeper insight into the relation between the effective MO effects and the underlying MSP mechanism.

The new method based on the transformation of EMA (TEMA) is applied to calculate the NIGH shift in the MSP gap of a square array of ferrite rods. Both the results for structures in the quasi-static limit and structures with significant size effects are discussed. It is shown that, unlike the NIGH effect in homogeneous MO materials, the reflected shift near the MSP resonance from the surface of the array is dominated by the value of effective permittivity instead of the magnetic field. Then a large NIGH shift is predicted even on total reflection when the structure parameters (such as rod radius or lattice constant) are adjusted to produce a small effective permittivity near the MSP resonance. Numerical simulations are performed to calculate the intensity and shift of reflected beams and the results coincide well with those from the TEMA. Giant NIGH shifts (i.e. shift larger than the incident wavelength) [24] well beyond the strong-field limit in homogeneous ferrites and enhanced magnetic-switch effects of the shift sign are demonstrated by tailoring the metamaterial structure as theoretically predicted.

2. The TEMA method

Consider a periodic square lattice as illustrated in Fig. 1(a) which consists of cylindrical ferrite rods of radius rs separated by distance a. The whole lattice is embedded in a uniform dielectric background of relative permittivity εb and permeability μb. The effective electromagnetic parameters of a metamaterial in this structure can be retrieved by solving the scattering problem of a coated particle embedded in the effective medium [22,23]. As shown in Fig. 1(b), an individual component particle of the metamaterial is surrounded by a coating layer of the background medium. The radius r0 of the coated particle is determined by the lattice constant a according to the relation πr02=a2. The total scattering cross section of such a coated particle embedded in the effective medium is given by [23]

Csca=4kem|bem|2
where ke refers to the wave number in the effective medium and bem is the Mie scattering coefficient of the m-th order. When ker0<<1, Csca is dominated by the terms of m=0,±1. Since the coated particle represents the microstructure of the composite, the scattering from it vanishes when it is surrounded by the effective medium, which yields
bem=0form=0,±1
For particles of normal dielectric, the effective permittivity εe and permeability μe are respectively obtained by solving the equation of m=0 and the equation of m=1 (or m=1 because of the symmetry between the m=±1 terms in the non-MO case). The results were compared with those from band calculations in [23], Wu et al. and good agreement was found near the Γ-point (keff0) even for particle spacing comparable to the incident wavelength.

 figure: Fig. 1

Fig. 1 (a) Structure of a MO metamaterial consisting of a square array of ferrite rods. (b) A conventional EMA model where an individual ferrite rod surrounded by a shell of the background medium is embedded in the MO effective medium. (c) Non-MO transformed effective media of order m = 0, 1, −1.

Download Full Size | PDF

When the metamaterial is composed of MO rods magnetized along the cylinder axis (the z axis), both the particle permeability and the effective permeability are in the tensor form in Eq. (1), namely

μ^s=(μsriμsi0iμsiμsr0001),μ^e=(μeriμei0iμeiμer0001)
Note that in the concerned 2D problem where an s-polarized wave travels in a MO material, one needs three parameters to characterize the electromagnetic properties of the MO material. The parameters may be chosen as (ε,μr,μi) or, equivalently, (ε,μvg,Q), where Q=μi/μr and μvg=μr(1Q2) is the Voigt permeability which directly relates to the wave vector in the MO material kMO=εμvgω/c.

The core region (r<rs) and the effective-medium region (r>r0) in Fig. 1(b) are characterized by (εs,μs,vg,Qs) and (εe,μe,vg,Qe), respectively. For an s-polarized incident wave of frequency ω, the electric field in each region can be expanded in terms of Bessel functions

Ee(r)=m[AemJm(ker)+BemHm(1)(ker)]eimφe^z,r>r0Eb(r)=m[AbmJm(kbr)+BbmHm(1)(kbr)]eimφe^z,rs<r<r0Es(r)=m[AsmJm(ksr)]eimφe^z,r<rs
where (r,φ) are the polar coordinates of the position vector r. Jm(x) and Hm(1)(x), respectively, are Bessel functions and the first-type Hankel functions. And the three wave numbers are given by ke=εeμe,vgω/c, kb=εbμbω/c, ks=εsμs,vgω/c.

The expressions of magnetic field can be obtained by the Maxwell’s equations:

μ0μbHb/t=×Ebandμ0μ^αHα/t=×Eα(α=s,e).

The relations between the coefficients in Eq. (6) are determined by the boundary conditions of E and H. At the boundary r=rs, we have

AbmJm(kbrs)+BbmHm(1)(kbrs)=Jm(ksrs)AbmJm'(kbrs)+BbmHm(1)'(kbrs)=ξsmJm'(ksrs)
Here Asm has set to be unit. ξsm is defined as
ξsm=[1mQsJm(ksrs)ksrsJm'(ksrs)]zbzs,vg
with zb=μb/εb denoting the background impedance while zs,vg=μs,vg/εs the impedance of a non-MO medium of scalar permittivity εs and scalar permeability μs,vg. Accordingly, the scattering coefficient of the m-th order for a MO particle (εs,μs,vg,Qs) in the background medium is obtained to be
bsm=BbmAbm=ξsmJm(kbrs)Jm'(ksrs)Jm(ksrs)Jm'(kbrs)Jm(ksrs)Hm(1)'(kbrs)ξsmJm'(ksrs)mHm(1)(kbrs)
For Qs=0, the permeability tensor μ^s of the particle reduces to scalar, and Eq. (9) reduces to the result for a particle of normal dielectric (εs,μs), i.e.,
b¯sm=ξ¯sJm(kbrs)Jm'(ksrs)Jm(ksrs)Jm'(kbrs)Jm(ksrs)Hm(1)'(kbrs)ξ¯sJm'(ksrs)Hm(1)(kbrs)
with ξ¯s=zb/zs representing the impedance ratio between the background and the non-MO particle.

Comparing Eqs. (9) and (10), it is easy to find the scattering coefficient of the m-th order of the MO particle is equivalent to that of a non-MO particle (εs(m),μs(m)), where εs(m) and μs(m) satisfy

εs(m)μs(m)=εsμs,vgorns(m)=ns,vgμs(m)εs(m)=11mQsFm(ksrs)μs,vgεsorzs(m)=zs,vg1mQsFm(ksrs)
The function Fm(ksrs) is expressed as Fm(x)=Jm(x)/[xJm'(x)]. In addition, the equivalent non-MO particle for the m-th order transformation can also be characterized by its refractive index ns(m) and impedance zs(m), where ns,vg and zs,vg are the refractive index and the impedance of the non-MO medium (εs,μs,vg).

At the boundary r=r0, the coefficients for m=0,±1 satisfy

AbmJm(kbr0)+BbmHm(1)(kbr0)=AemJm(ker0)AbmJm'(kbr0)+BbmHm(1)'(kbr0)=AemξemJm'(ker0)
Here ξem=[1mQeFm(ker0)]zbze,vgwith the impedance of a non-MO dielectric (εe,μe,vg) ze,vg=μe,vg/εe. In above equations, we have applied the EMA conditions bem=Bem/Aem=0 for m=0,±1. From Eq. (12) we have

1ξemJm(ker0)Jm'(ker0)=Jm(kbr0)+bsmHm(1)(kbr0)Jm'(kbr0)+bsmHm(1)'(kbr0)

If the metamaterial is composed of non-MO particles (Qs=0, Qe=0), the effective permittivity ε¯e and the effective permeability μ¯e can be obtained from the m=0 equation and the m=1 equation (or m=1), respectively, as

ε¯e=2εbkbr0J0'(kbr0)+b¯s0H0(1)'(kbr0)J0(kbr0)+b¯s0H0(1)(kbr0),μ¯e=μbkbr0J1(kbr0)+b¯s1H1(1)(kbr0)J1'(kbr0)+b¯s1H1(1)'(kbr0).
It is easy to check that the above results can be rewritten into the conventional EMA equations as given in [23], Wu et al.

For the metamaterial composed of MO particles (εs,μs,vg,Qs), we have

ξe0=zbze,vg,ξe1=(1Qe)zbze,vg,ξe,1=(1+Qe)zbze,vg
for ker0<<1. By substituting Eq. (15) into Eq. (13) and comparing the results with those for the non-MO case, it is found that the effective parameters for the MO metamaterial are given by
εe=ε¯e|(εs(0),μs(0)),μe,vg1Qe=μ¯e|(εs(1),μs(1)),μe,vg1+Qe=μ¯e|(εs(1),μs(1))
Here ε¯e|(εs(m),μs(m)) and μ¯e|(εs(m),μs(m)) are the effective permittivity and permeability of a metamaterial composed of the equivalent particles of the m-th order as defined in Eq. (11) [see Fig. 1(c)]. Let εe(0)=ε¯e|(εs(0),μs(0)), μe(+)=μ¯e|(εs(1),μs(1)), μe()=μ¯e|(εs(1),μs(1)), we have

εe=εe(0),2μe,vg=1μe(+)+1μe(),Qe=μe(+)μe()μe(+)+μe()

Equation (17) establishes a correspondence between the concerned MO metamaterial and the three non-MO metamaterials defined according to the transformation relations given by Eq. (11). Accordingly we have a three-step strategy to solve the problem of MO-metamaterials described as follows.

Step 1. Apply Eq. (11) to determine εs(m) and μs(m) for m=0,±1 from the properties of the original MO particles.

Step 2. Replace the MO-particles by the non-MO ones of (εs(m),μs(m)), without changing the underlying structure (characterized by rs and a), and find the effective parameters εe(0), μe(+), μe() by using the conventional EMA results Eq. (14).

Step 3. Substituting (εe(0),μe(+),μe()) into Eq. (17) to obtain the effective parameters (εe,μe,vg,Qe) for the original MO metamaterials.

If necessary, one can also deduce the effective permeability tensor.

μ^e=(μeriμei0iμeiμer0001)
where μer=μe(+)+μe()2 and μei=μe(+)μe()2, with

μe(+)=μbkbr0J1(kbr0)+bs,1H1(1)(kbr0)J1'(kbr0)+bs,1H1(1)'(kbr0),μe()=μbkbr0J1(kbr0)+bs,1H1(1)(kbr0)J1'(kbr0)+bs,1H1(1)'(kbr0).

Here we would like to mention that an obvious benefit of this new strategy is to avoid the complicated calculation of scattering when the MO tensors are involved. The results are easily extended to the cases beyond the long-wavelength limit kbr0<<1 by treating the transformed media with Eq. (14), the general EMA formula for non-MO metamaterials which is valid even when kbr0>1 [23]. And, since the EMA approach for non-MO cases has been thoroughly discussed and extensively applied, to transform an MO problem to non-MO ones is helpful to gain insight into the phenomena such as NIGH effects occurring in MO metamaterials.

3. NIGH shifts

Now we begin to consider a typical subwavelength photonic crystal consisting of YIG rods in a vacuum background (εb=1,μb=1). As illustrated in Fig. 2, an electrically z-polarized paraxial beam is perpendicularly incident along the x-axis upon the surface of such a square array of YIG rods. When the effective medium approximation is applied, the NIGH shift of the reflected beam is determined by the effective parameters as [8,11]

de=λ0πQeμe,vgεe
The values of the effective parameters can be obtained by applying the transformation strategy to this case, where the magnetic permeability tensor of YIG magnetized along the z-axis follows the typical dispersion of ferrites [20]:
μsr=1+ωm(ω0+iαω)(ω0+iαω)2ω2,μsi=ωmω(ω0+iαω)2ω2
Here, ω is the frequency of the incident light.ω0, ωm are the magnetic resonance frequencies given by ω0=2πγh0, ωm=2πγms with h0 and ms denoting the applied magnetic field and the saturated magnetization, respectively. The damping factor α is quite small due to the low loss of YIG and will be neglected in the following theoretical analysis but discussed later in numerical simulations. Other material parameters of YIG are chosen to be: γ=2.8×103GHz/Oe, ms=1800Gauss, εs=14.5 [20].

 figure: Fig. 2

Fig. 2 Schematic figure of the geometric arrangement to calculate the NIGH shifts. The incident beam is electrically z-polarized, normally incident upon the surface of the square lattice of YIG rods.

Download Full Size | PDF

3.1. Quasi-static structures

We first investigate structures composed of extremely fine particles so that kbr0<<1 and ksrs<<1 are satisfied simultaneously, i.e. structures in the quasi-static limit. Then Eq. (14), the general EMA result for a non-MO metamaterial, can be approximately replaced by the well-known Maxwell-Garnett equations

ε¯e=pεs+(1p)εbandμ¯eμbμ¯e+μb=pμsμbμs+μb
where p=(rs/r0)2 represents the particle concentration. The transformation relations given by Eq. (11) are also simplified in this quasi-static approximation (QSA). For m=0,±1, we have
εs(m)=εs(1mQs)andμs(m)=μs,vg1mQs
Combining Eqs. (20), (21) and (16) yields the analytic expressions of the three effective parameters of the MO metamaterials in the QSA, where the effective permittivity εe is the direct average of the component dielectric constants,
εe=pεs+(1p)εb
while μe,vg=2μe(+)μe()/(μe(+)+μe()) and Qe=(μe(+)μe())/(μe(+)+μe()) are determined by substituting
μe(±)μb=μs,vg(±)+fPμbfPμs,vg(±)+μb
with fP=1p1+p and μs,vg(±)=μs,vg1Qs.

Figure 3 shows the variance of μe(±) and μe,vg in the QSA with the Voigt permeability μs,vg of MO particles for Qs=0,0.5,2. The particle concentration has been set to be p=0.1. For Qs=0, i.e. the case of non-MO particles, the three permeability curves coincide since μe,vg=μe(+)=μe() when no MO effects are included. Typical permeability resonance is produced around μs,vg=1, a signature of the occurrence of the MSP resonanceas reported in the previous studies [16,18,19]. These MSP resonances of the total composite correspond to the poles in μe,vg. For a single rod or a structure in the limit p0 when Qs=0, the MSP condition reduces to the conventional form μe,vg=1. While for non-vanishing Qs, the resonance of μe,vg is split into two sub-bands, either of which corresponds to the resonance regime of μe(+) or μe(). The distance between the sub-bands is enlarged with strengthened MO effects. Such a resonance splitting due to MO effects is a direct result of the imbalance between the scattering coefficients bs,1 and bs,1 caused by the broken time-reversal symmetry of MO particles [25].

 figure: Fig. 3

Fig. 3 Permeability μe(+) (thin red line), μe() (thin blue line) and μe,vg (thick black line) for various Qs when p=0.1 in a quasi-static structure.

Download Full Size | PDF

Band gaps, i.e., the negative-μe,vg regions as shown in Fig. 3, are formed around each MSP resonance. These gap regions are also called “reststrahl regions” since the reflectivity is extremely high in them. EMA works quite well near the gap end with μe,vg=0 (since ker0<<1 is ensured in this region) and the NIGH shift can be appropriately estimated by

de(±)=λ0πQe(±)μe,vgεe±λ0π1εe
Here the superscript +/ corresponds to the sub-band with μe(+)=0 or μe()=0, and the second step is obtained by substituting Qe(±)1 near μe(±)=0 according to Eq. (17). Accordingly, a giant NIGH shift in such a MO metamaterial is only realized when the effective permittivity approaches zero.

By comparing Eq. (24) and Eq. (3), it is found the MSP-induced shift de(±) is dominated by the effective permittivity rather than driven by strong magnetic fields as is the shift drTR associated with IFR. But the two shifts share the same upper limit λ0/π whenever the QSA can be applied to the MO metamaterial, since the effective permittivity in this approximation cannot be lowered to less than 1 according to Eq. (22). In other words, the key factor to break the restriction on the NIGH shift is to introduce proper size-effects beyond QSA into the MO structures. In the following sub-section, we will illustrate how to realize large NIGH shifts in “coarser” MO structures when particle-size and spacing are carefully adjusted to achieve small effective permittivity around MSP resonance.

3.2. Size effects

Nontrivial size effects on MSP bands and gaps have been demonstrated, though not emphasized, in previous studies of band structures for subwavelength YIG-PCs. A typical example is the experimental work of omnidirectional waveguides based on a square array of YIG rods with rs=0.0016m and a=0.0145m [20]. It was shown that, a MSP-induced gap is opened around fg=9.6GHz when the YIG rods are magnetized at h0=2980Oe. Though this gap exhibits distinctive MSP features, the observed frequency range deviates obviously from the predicted electrostatic resonance in the QSA. As shown in Fig. 3, the QSA leads to the prediction of the formation of two MSP gaps near the resonance of μe(+) and μe(), respectively, which can be estimated by the condition

μs,vg(±)=μs,vg1Qs=1
Substituting the dispersion of YIG permeability into this equation, we obtain two resonance frequencies in the QSA for YIG rods as
fr(±)=±(f0+fm2)
with f0=ω0/(2π) and fm=ωm/(2π). At h0=2980Oe, fr() is negative, so that fr(+)=10.9GHz is the unique physical solution, which explains why only a single MSP-gap is observed in the experiment. But the deviation more than 1000 between fg and fr(+) suggests that this structure is well beyond the simple QSA with significant size effects.

The red-shift of the MSP gap can be well-explained when we apply the TEMA method to this structure. The dispersion relations of the parameters (εs(m),μs(m)) of the component particles for the m-th order transformed medium (m=0,±1) are obtained by substituting the YIG permittivity and permeability into Eq. (11). The frequency dependence of εe(0), μe(+), μe() are then calculated and illustrated in Fig. 4, together with the relevant scattering features of individual component particles for each transformed medium. It is shown that dense and sharp resonances between 10GHz and 11GHz will occur in all the three spectra, which are bulk modes originated from the singularity of the Voigt permeability μs,vg of YIG, and irrelevant to the concerned MSP band. The LSPR peak of the first-order particle, which is at fr(+)=10.9GHz in the QSA, has now evolved into a broad Mie resonance around 9.6GHz due to the large particle radius, leading to a pole in μe(+) near this frequency [Fig. 4(b)]. The large particle size also results in the resonance of εe(0) associated with a monopole peak around 8.0GHz for the zeroth-order particles [Fig. 4(a)]. The shift of LSPR modes and the occurrence of monopole resonance are not surprising but common size effects extensively observed in non-MO particles [25–28]. But in our case of MO metamaterials, the two effects can be combined and controlled to produce an MSP gap with large NIGH shift.

 figure: Fig. 4

Fig. 4 Frequency dependence of the effective parameters and the scattering coefficients for the transformed media of the m-th order. (a) εe(0) and bs(0) for the m=0 medium; (b) μe(+) and bs(+) for the m = 1 medium; (c) μe() and bs() for the m=1 medium; Here bs(0) is the zeroth-order scattering coefficient of an individual particle for m = 0 while bs(±) the first-order coefficients for m=±1. The right panels are the partial enlarged drawing of the high frequency part.

Download Full Size | PDF

Figure 5(a) shows the frequency dependence of εe and μe,vg for the array of YIG rods calculated by substituting the results of εe(0), μe(+) and μe() into Eq. (17). The resonance of μe(+) from the shifted LSPR mode contributes a zero-point of μe,vg at fm=9.685GHz while the monopole resonance of εe(0) leads to a zero-point of εe at fe=9.255GHz. In the frequency range between the two zero-points, ke2=εeμe,vgω2/c2 is negative and an MSP gap takes place. Note that |ker0| is so close to zero [Fig. 5(b)] that the TEMA is an appropriate approximation in the whole gap region, and a large NIGH shift is expected since the value of εe is now considerably suppressed due to the monopole resonance at lower frequencies. Figure 5(c) gives the calculated NIGH shift (solid line) in the MSP-gap region, where the two edges of the gap, fe and fm, are labeled by vertical dashed lines. It is shown that the normalized NIGH shift dr/λ0 reaches up to nearly 0.7 at the gap end fm with μe,vg=0, more than twice the QSA limit of 1/π.

 figure: Fig. 5

Fig. 5 Comparison between the TEMA results and numerical simulations for different lattice constants. (a) & (d) & (g): Frequency dependences of εe and μe,vg for the array of YIG rods calculated from the TEMA method, a0=0.0145m. (b) & (e) & (h): Comparison of the MSP-gap positions. Dashed lines: (100ker0)2 vs f from the TEMA method; Solid lines with squares: the field magnitude at the center of the reflected beam from simulations. (c) & (f) & (i): Comparison of the NIGH shifts. Solid lines: results from the TEMA method; Squares: results from the numerical simulations using Comsol Multiphysics (a normally-incident Gaussian beam with finite width).

Download Full Size | PDF

The NIGH shift can be further increased by adjusting the lattice constant a. Figure 5 compares the calculated MSP gap and the associated NIGH shift for a=a0 [Figs. 5(a)-5(c)], 0.9a0 [Figs. 5(d)-5(f)] and 0.8a0 [Figs. 5(g)-5(i)] with a0=0.0145m. Slight broadening of the εe(0) and μe(+) resonances can be observed since the coupling between particles is strengthened with the decrease of lattice constant. As a consequence, a narrower MSP gap with lower εe, and hence, larger NIGH shift, is formed in structures with smaller lattice constant. When a is decreased to 0.8a0, the normalized NIGH shift dr/λ0 at the gap end with μe,vg=0 rises to be nearly 1.5, which is the so-called “giant GH effect”, i.e. a GH-shift larger than the incident wavelength.

To verify the above TEMA results, we performed a finite-element analysis of these YIG arrays using the software Comsol Multiphysics 4.3a. In the simulations the normally-incident beam is a z-polarized Gaussian beam with a finite width. The electric-field profile of the reflected beam is calculated and the reflected GH-shift is obtained by comparing the central positions between the reflected beam and the incident beam [Fig. 6(a)], where the incident beam is set to be centrally located around the x-axis with half-width w=6λ0. When the frequency approaches the gap end with μe,vg=0, large NIGH shifts are observed as predicted in the TEMA. The distributions of electric field at different frequencies in the gap region are illustrated in Figs. 6(b)-6(d). With the frequency increasing, the field pattern in MO rods evolves into a typical surface dipole mode, corresponding to the dipole resonance shown in Fig. 4. The frequency dependences of the centroid field amplitude |Er| for the reflected beam in the simulations are presented in Figs. 5(b), 5(e) and 4(h), together with the values of (ker0)2 calculated from the TEMA method. It is shown that the incident beam in simulations is totally reflected for frequencies within the TEMA-predicted gap region (the region where (ker0)2 is negative), with a reflected shift in good agreement with the TEMA result as illustrated in Figs. 5(c), 5(f) and 5(i).

 figure: Fig. 6

Fig. 6 (a) The electric-field amplitude of the reflected beam from numerical simulations for different incident frequencies with (dashed curves) and without (solid curves) absorption. Blue: f = 9.3GHz; Yellow: f = 9.5GHz; Red: f = 9.65GHz. (b) & (c) & (d): The patterns of electric fields for each frequency. Structure parameters are chosen to be rs=0.0016m and a=0.0145m.

Download Full Size | PDF

So far we have neglected the absorption of YIG in both the effective-medium analysis and the numerical simulations. Results for more realistic materials are presented as well in Fig. 6(a) (dashed lines) where the YIG damping factor α=γdH/2ωwith linewidth dH=30Oe [20] is included in the calculation of the electric-field profiles of the reflected beam. The non-vanishing damping factor is shown to result in a moderate or even slight decrease of reflectivity with the NIGH shift almost unchanged. This holds even at frequencies very close to the MSP resonance. For example, at f=9.65GHz for a structure of rs=0.0016m and a=0.0145m, the central field of the reflected beam is only reduced from 0.9 without absorption to 0.8 with absorption, while the reflected shift remains to be 0.617λ0. The reflected beam is still “bright” enough for the observation of large NIGH shifts when absorption occurs. Two reasons account for this result. First, the NIGH shift is generally insensitive to the absorption of MO materials, as confirmed in previous studies of various samples [8,10,11]. Secondly, the MSP resonance originates from the surface mode of MO particles, so that much less absorption will be observed than those due to conventional bulk modes [17,20].

Though all the shifts obtained above are positive (along the + y-axis), negative NIGH shifts can be easily achieved by reversing the direction of the applied magnetic field from ( + z)-axis to (-z)-axis. This magnetic-switch of the sign of reflected shift is an important feature of the NIGH effect in MO materials [10,11] and can be remarkably enhanced in MO metamaterials with the MSP resonance. Figure 7 shows the electric-field patterns at different magnetic fields for a structure of a=0.8a0 and rs=0.0016m illuminated by an incident beam of half-width w=λ0 at frequency f=9.75GHz. Here we have adopted “narrow” beams to provide a clearer illustration of the beam shifts. It is shown that the reflected shift is switched from 1.02λ0 to 1.02λ0 when the applied field is changed to opposite direction. Again, both the results for the cases of zero-absorption [Figs. 7(a)-7(c)] and nan-vanishing absorption [Figs. 7(d)-7(f)] are presented and the inclusion of absorption leads to a slightly-weaker reflected beam with almost the same shift.

 figure: Fig. 7

Fig. 7 Enhanced magnetic switch of NIGH shifts near the MSP resonance. (a)-(c): Patterns of electric field amplitude for the incident beam and the reflected beams without absorption; (d)-(f): Patterns of electric field amplitude for the incident beam and the reflected beams with absorption. Structure parameters are chosen to be a = 0.8a0 and rs = 0.0016m. Incident frequency is set to be f = 9.75GHz.

Download Full Size | PDF

4. Conclusion

In conclusion, our analysis based on the TEMA method reveals that enhanced NIGH shifts on total reflection can be realized in the MO metamaterials when particle size and spacing are carefully adjusted to produce a vanishing effective permittivity around the MSP resonance. The theoretical prediction is verified by numerical simulations. Giant NIGH shifts and enhanced magnetic switching effects near the MSP resonance [29] are demonstrated and the effects of absorption are discussed. We hope this study could invoke more interests in such a special GH shift which is rooted in the combination of MSP resonances and TRS breaking, and provide new clues to practical applications of NIGH effects with designs of metamaterials.

Here some comments are in order. Through homogenization of an array of gyromagnetic cylinders, photonic Chern insulator may be created [30]. Such investigation achieves the electromagnetic analogs of quantum spin Hall systems, which provides a platform for us to explore the interplay between disorder and topology on GH shift of such metamaterials like photonic Chern insulator.

Funding

National Natural Science Foundation of China (Grant No. 1137422, No. 11774252); National Science of Jiangsu Province (Grant No. BK20161210); Qing Lan project; PAPD of Jiangsu Higher Education Institutions.

References and links

1. B. Zhao and L. Gao, “Temperature-dependent Goos-Hänchen shift on the interface of metal/dielectric composites,” Opt. Express 17(24), 21433–21441 (2009). [CrossRef]   [PubMed]  

2. D. L. Gao and L. Gao, “Goos–Hänchen shift of the reflection from nonlinear nanocomposites with electric field tenability,” Appl. Phys. Lett. 97(4), 041903 (2010). [CrossRef]  

3. Y. Xu, C. T. Chan, and H. Chen, “Goos-Hänchen effect in epsilon-near-zero metamaterials,” Sci. Rep. 5(1), 8681 (2015). [CrossRef]   [PubMed]  

4. X. Li, P. Wang, F. Xing, X. D. Chen, Z. B. Liu, and J. G. Tian, “Experimental observation of a giant Goos-Hänchen shift in graphene using a beam splitter scanning method,” Opt. Lett. 39(19), 5574–5577 (2014). [CrossRef]   [PubMed]  

5. Y. L. Chuang and R. K. Lee, “Giant Goos-Hänchen shift using PT symmetry,” Phys. Rev. A 92(1), 013815 (2015). [CrossRef]  

6. P. Ma and L. Gao, “Large and tunable lateral shifts in one-dimensional PT-symmetric layered structures,” Opt. Express 25(9), 9676–9688 (2017). [CrossRef]   [PubMed]  

7. Q. D. Jiang, H. Jiang, H. Liu, Q. F. Sun, and X. C. Xie, “Topological Imbert-Fedorov shift in Weyl semimetals,” Phys. Rev. Lett. 115(15), 156602 (2015). [CrossRef]   [PubMed]  

8. W. Yu, H. Sun, and L. Gao, “Magnetic control of Goos-Hänchen shifts in a yttrium-iron-garnet film,” Sci. Rep. 7, 45866 (2017). [CrossRef]   [PubMed]  

9. P. Gruszecki, J. Romero-Vivas, Y. S. Dadoenkova, N. N. Dadoenkova, I. L. Lyubchanskii, and M. Krawczyk, “Goos-Hänchen effect and bending of spin wave beams in thin magnetic films,” Appl. Phys. Lett. 105(24), 242406 (2014). [CrossRef]  

10. F. Lima, T. Dumelow, J. A. P. Da Costa, and E. L. Albuquerque, “Lateral shift of far infrared radiation on normal incidence reflection off an antiferromagnet,” EPL 83(1), 17003 (2008). [CrossRef]  

11. F. Lima, T. Dumelow, E. L. Albuquerque, and J. A. P. Da Costa, “Power flow associated with the Goos-Hänchen shift of a normally incident electromagnetic beam reflected off an antiferromagnet,” Phys. Rev. B 79(15), 155124 (2009). [CrossRef]  

12. F. Lima, T. Dumelow, E. L. Albuquerque, and J. A. P. Da Costa, “Nonreciprocity in the Goos–Hänchen shift on oblique incidence reflection off antiferromagnets,” J. Opt. Soc. Am. B 28(2), 306–313 (2011). [CrossRef]  

13. R. Macêdo, R. L. Stamps, and T. Dumelow, “Spin canting induced nonreciprocal Goos-Hänchen shifts,” Opt. Express 22(23), 28467–28478 (2014). [CrossRef]   [PubMed]  

14. R. Macêdo and T. Dumelow, “Tunable all-angle negative refraction using antiferromagnets,” Phys. Rev. B 89(3), 035135 (2014). [CrossRef]  

15. R. W. Damon and J. R. Eshbach, “Magnetostatic modes of a ferromagnet slab,” J. Phys. Chem. Solids 19(3–4), 308–320 (1961). [CrossRef]  

16. J. N. Gollub, D. R. Smith, D. C. Vier, T. Perram, and J. J. Mock, “Experimental characterization of magnetic surface plasmons on metamaterials with negative permeability,” Phys. Rev. B 71(19), 195402 (2005). [CrossRef]  

17. S. Liu, W. Lu, Z. Lin, and S. T. Chui, “Molding reflection from metamaterials based on magnetic surface plasmons,” Phys. Rev. B 84(4), 045425 (2011). [CrossRef]  

18. S. Liu, J. Du, Z. Lin, R. X. Wu, and S. T. Chui, “Formation of robust and completely tunable resonant photonic band gaps,” Phys. Rev. B 78(15), 155101 (2008). [CrossRef]  

19. S. T. Chui, S. Liu, and Z. Lin, “Multiple flat photonic bands with finite Chern numbers,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 88(3), 031201 (2013). [CrossRef]   [PubMed]  

20. J. Lian, J. X. Fu, L. Gan, and Z. Y. Li, “Robust and disorder-immune magnetically tunable one-way waveguides in a gyromagnetic photonic crystal,” Phys. Rev. B 85(12), 125108 (2012). [CrossRef]  

21. S. Liu, W. Chen, J. Du, Z. Lin, S. T. Chui, and C. T. Chan, “Manipulating Negative-Refractive Behavior with a Magnetic Field,” Phys. Rev. Lett. 101(15), 157407 (2008). [CrossRef]   [PubMed]  

22. J. Jin, S. Liu, Z. Lin, and S. T. Chui, “Effective-medium theory for anisotropic magnetic metamaterials,” Phys. Rev. B 80(11), 115101 (2009). [CrossRef]  

23. Y. Wu, J. Li, Z. Q. Zhang, and C. T. Chan, “Effective medium theory for magnetodielectric composites: Beyond the long-wavelength limit,” Phys. Rev. B 74(8), 085111 (2006). [CrossRef]  

24. C. F. Li, “Negative Lateral Shift of a Light Beam Transmitted through a Dielectric Slab and Interaction of Boundary Effects,” Phys. Rev. Lett. 91(13), 133903 (2003). [CrossRef]   [PubMed]  

25. Z. F. Huang, Y. X. Ni, and H. Sun, “Localized surface plasmon resonance and the size effects of magneto-optic rods,” Wuli Xuebao 65, 114202 (2016).

26. G. Shvets and Y. A. Urzhumov, “Engineering the electromagnetic properties of periodic nanostructures using electrostatic resonances,” Phys. Rev. Lett. 93(24), 243902 (2004). [CrossRef]   [PubMed]  

27. S. O’Brien and J. B. Pendry, “Photonic band-gap effects and magnetic activity in dielectric composites,” J. Phys. Condens. Matter 14(15), 4035–4044 (2002). [CrossRef]  

28. X. Fan, W. Zheng, and D. J. Singh, “Light scattering and surface plasmons on small spherical particles,” Light Sci. Appl. 3(6), e179 (2014). [CrossRef]  

29. Y. Poo, S. Liu, C. Xiao, R. X. Wu, and P. Chen, “Field enhancement of nonreciprocal electromagnetic wave supported by magnetic surface plasmon,” Opt. Express 22(22), 27717–27724 (2014). [CrossRef]   [PubMed]  

30. M. Xiao and S. H. Fan, “Photonic Chern insulator through homogenization of an array of particles,” Phys. Rev. B 96(10), 100202 (2017). [CrossRef]  

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 (7)

Fig. 1
Fig. 1 (a) Structure of a MO metamaterial consisting of a square array of ferrite rods. (b) A conventional EMA model where an individual ferrite rod surrounded by a shell of the background medium is embedded in the MO effective medium. (c) Non-MO transformed effective media of order m = 0, 1, −1.
Fig. 2
Fig. 2 Schematic figure of the geometric arrangement to calculate the NIGH shifts. The incident beam is electrically z-polarized, normally incident upon the surface of the square lattice of YIG rods.
Fig. 3
Fig. 3 Permeability μ e ( + ) (thin red line), μ e ( ) (thin blue line) and μ e , v g (thick black line) for various Q s when p = 0.1 in a quasi-static structure.
Fig. 4
Fig. 4 Frequency dependence of the effective parameters and the scattering coefficients for the transformed media of the m-th order. (a) ε e ( 0 ) and b s ( 0 ) for the m = 0 medium; (b) μ e ( + ) and b s ( + ) for the m = 1 medium; (c) μ e ( ) and b s ( ) for the m = 1 medium; Here b s ( 0 ) is the zeroth-order scattering coefficient of an individual particle for m = 0 while b s ( ± ) the first-order coefficients for m = ± 1 . The right panels are the partial enlarged drawing of the high frequency part.
Fig. 5
Fig. 5 Comparison between the TEMA results and numerical simulations for different lattice constants. (a) & (d) & (g): Frequency dependences of ε e and μ e , v g for the array of YIG rods calculated from the TEMA method, a 0 = 0.0145 m . (b) & (e) & (h): Comparison of the MSP-gap positions. Dashed lines: ( 100 k e r 0 ) 2 vs f from the TEMA method; Solid lines with squares: the field magnitude at the center of the reflected beam from simulations. (c) & (f) & (i): Comparison of the NIGH shifts. Solid lines: results from the TEMA method; Squares: results from the numerical simulations using Comsol Multiphysics (a normally-incident Gaussian beam with finite width).
Fig. 6
Fig. 6 (a) The electric-field amplitude of the reflected beam from numerical simulations for different incident frequencies with (dashed curves) and without (solid curves) absorption. Blue: f = 9.3GHz; Yellow: f = 9.5GHz; Red: f = 9.65GHz. (b) & (c) & (d): The patterns of electric fields for each frequency. Structure parameters are chosen to be r s = 0.0016 m and a = 0.0145 m .
Fig. 7
Fig. 7 Enhanced magnetic switch of NIGH shifts near the MSP resonance. (a)-(c): Patterns of electric field amplitude for the incident beam and the reflected beams without absorption; (d)-(f): Patterns of electric field amplitude for the incident beam and the reflected beams with absorption. Structure parameters are chosen to be a = 0.8a0 and rs = 0.0016m. Incident frequency is set to be f = 9.75GHz.

Equations (30)

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

μ ^ = ( μ r i μ i 0 i μ i μ r 0 0 0 1 )
d r 0.0136 η λ 0
d r TR = λ 0 π h 0 h 0 + m s
C s c a = 4 k e m | b e m | 2
b e m = 0 for m = 0 , ± 1
μ ^ s = ( μ s r i μ s i 0 i μ s i μ s r 0 0 0 1 ) , μ ^ e = ( μ e r i μ e i 0 i μ e i μ e r 0 0 0 1 )
E e ( r ) = m [ A e m J m ( k e r ) + B e m H m ( 1 ) ( k e r ) ] e i m φ e ^ z , r > r 0 E b ( r ) = m [ A b m J m ( k b r ) + B b m H m ( 1 ) ( k b r ) ] e i m φ e ^ z , r s < r < r 0 E s ( r ) = m [ A s m J m ( k s r ) ] e i m φ e ^ z , r < r s
μ 0 μ b H b / t = × E b and μ 0 μ ^ α H α / t = × E α ( α = s , e ) .
A b m J m ( k b r s ) + B b m H m ( 1 ) ( k b r s ) = J m ( k s r s ) A b m J m ' ( k b r s ) + B b m H m ( 1 ) ' ( k b r s ) = ξ s m J m ' ( k s r s )
ξ s m = [ 1 m Q s J m ( k s r s ) k s r s J m ' ( k s r s ) ] z b z s , v g
b s m = B b m A b m = ξ s m J m ( k b r s ) J m ' ( k s r s ) J m ( k s r s ) J m ' ( k b r s ) J m ( k s r s ) H m ( 1 ) ' ( k b r s ) ξ s m J m ' ( k s r s ) m H m ( 1 ) ( k b r s )
b ¯ s m = ξ ¯ s J m ( k b r s ) J m ' ( k s r s ) J m ( k s r s ) J m ' ( k b r s ) J m ( k s r s ) H m ( 1 ) ' ( k b r s ) ξ ¯ s J m ' ( k s r s ) H m ( 1 ) ( k b r s )
ε s ( m ) μ s ( m ) = ε s μ s , v g or n s ( m ) = n s , v g μ s ( m ) ε s ( m ) = 1 1 m Q s F m ( k s r s ) μ s , v g ε s or z s ( m ) = z s , v g 1 m Q s F m ( k s r s )
A b m J m ( k b r 0 ) + B b m H m ( 1 ) ( k b r 0 ) = A e m J m ( k e r 0 ) A b m J m ' ( k b r 0 ) + B b m H m ( 1 ) ' ( k b r 0 ) = A e m ξ e m J m ' ( k e r 0 )
1 ξ e m J m ( k e r 0 ) J m ' ( k e r 0 ) = J m ( k b r 0 ) + b s m H m ( 1 ) ( k b r 0 ) J m ' ( k b r 0 ) + b s m H m ( 1 ) ' ( k b r 0 )
ε ¯ e = 2 ε b k b r 0 J 0 ' ( k b r 0 ) + b ¯ s 0 H 0 ( 1 ) ' ( k b r 0 ) J 0 ( k b r 0 ) + b ¯ s 0 H 0 ( 1 ) ( k b r 0 ) , μ ¯ e = μ b k b r 0 J 1 ( k b r 0 ) + b ¯ s 1 H 1 ( 1 ) ( k b r 0 ) J 1 ' ( k b r 0 ) + b ¯ s 1 H 1 ( 1 ) ' ( k b r 0 ) .
ξ e 0 = z b z e , v g , ξ e 1 = ( 1 Q e ) z b z e , v g , ξ e , 1 = ( 1 + Q e ) z b z e , v g
ε e = ε ¯ e | ( ε s ( 0 ) , μ s ( 0 ) ) , μ e , v g 1 Q e = μ ¯ e | ( ε s ( 1 ) , μ s ( 1 ) ) , μ e , v g 1 + Q e = μ ¯ e | ( ε s ( 1 ) , μ s ( 1 ) )
ε e = ε e ( 0 ) , 2 μ e , v g = 1 μ e ( + ) + 1 μ e ( ) , Q e = μ e ( + ) μ e ( ) μ e ( + ) + μ e ( )
μ ^ e = ( μ e r i μ e i 0 i μ e i μ e r 0 0 0 1 )
μ e ( + ) = μ b k b r 0 J 1 ( k b r 0 ) + b s , 1 H 1 ( 1 ) ( k b r 0 ) J 1 ' ( k b r 0 ) + b s , 1 H 1 ( 1 ) ' ( k b r 0 ) , μ e ( ) = μ b k b r 0 J 1 ( k b r 0 ) + b s , 1 H 1 ( 1 ) ( k b r 0 ) J 1 ' ( k b r 0 ) + b s , 1 H 1 ( 1 ) ' ( k b r 0 ) .
d e = λ 0 π Q e μ e , v g ε e
μ s r = 1 + ω m ( ω 0 + i α ω ) ( ω 0 + i α ω ) 2 ω 2 , μ s i = ω m ω ( ω 0 + i α ω ) 2 ω 2
ε ¯ e = p ε s + ( 1 p ) ε b and μ ¯ e μ b μ ¯ e + μ b = p μ s μ b μ s + μ b
ε s ( m ) = ε s ( 1 m Q s ) and μ s ( m ) = μ s , v g 1 m Q s
ε e = p ε s + ( 1 p ) ε b
μ e ( ± ) μ b = μ s , v g ( ± ) + f P μ b f P μ s , v g ( ± ) + μ b
d e ( ± ) = λ 0 π Q e ( ± ) μ e , v g ε e ± λ 0 π 1 ε e
μ s , v g ( ± ) = μ s , v g 1 Q s = 1
f r ( ± ) = ± ( f 0 + f m 2 )
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.