Open access peer-reviewed chapter

Dyadic Green’s Function for Multilayered Planar, Cylindrical, and Spherical Structures with Impedance Boundary Condition

Written By

Shiva Hayati Raad and Zahra Atlasbaf

Submitted: 04 December 2020 Reviewed: 05 January 2021 Published: 02 December 2021

DOI: 10.5772/intechopen.95834

From the Edited Volume

Electromagnetic Wave Propagation for Industry and Biomedical Applications

Edited by Lulu Wang

Chapter metrics overview

466 Chapter Downloads

View Full Metrics

Abstract

The integral equation (IE) method is one of the efficient approaches for solving electromagnetic problems, where dyadic Green’s function (DGF) plays an important role as the Kernel of the integrals. In general, a layered medium with planar, cylindrical, or spherical geometry can be used to model different biomedical media such as human skin, body, or head. Therefore, in this chapter, different approaches for the derivation of Green’s function for these structures will be introduced. Due to the recent great interest in two-dimensional (2D) materials, the chapter will also discuss the generalization of the technique to the same structures with interfaces made of isotropic and anisotropic surface impedances. To this end, general formulas for the dyadic Green’s function of the aforementioned structures are extracted based on the scattering superposition method by considering field and source points in the arbitrary locations. Apparently, by setting the surface conductivity of the interfaces equal to zero, the formulations will turn into the associated problem with dielectric boundaries. This section will also aid in the design of various biomedical devices such as sensors, cloaks, and spectrometers, with improved functionality. Finally, the Purcell factor of a dipole emitter in the presence of the layered structures will be discussed as another biomedical application of the formulation.

Keywords

  • biomedical
  • dyadic Green’s function
  • integral equation
  • 2D materials
  • sensors
  • Purcell factor

1. Introduction

Planarly, cylindrically and spherically layered media have been widely used to model the human skin, body, or head. In particular, a rectangular slab is proposed to model and analyze skin temperature distribution [1]. Moreover, in the multilayered skin model, three stacked layers are exploited to simulate the performance of the epidermis, dermis, and sub-cutis parts of the skin [2]. In other research, a planarly layered medium has been proposed as the simplified human body model by considering the impact of skin, fat, and muscle in the electromagnetic performance [3]. Cylindrical-shaped equivalent phantom of the skin is alternatively used to characterize the interactions between an antenna and the human body [4]. For a more precise investigation, multilayered cylinders are proposed to model a biological system with different tissues [5]. Considering spherical geometrics, the interactions of a three-layered spherical human head model with a finite-length dipole and a cellular phone helical antenna are investigated using Green’s function [6, 7]. The research is extended to the six-layered model, including skin, fat, bone, dura, CSF, and brain layers of the head [8]. This model has also been used to explore the impact of the shapes and positions of coils in the MRI system [9]. For the analysis of the aforementioned structures, integral equation (IE) based methods have been widely used. Dyadic Green’s function, which is a powerful method to calculate the electromagnetic response for different excitation sources, plays an essential role in the IE method [10, 11, 12].

Different mathematical approaches can be utilized to calculate the dyadic Green’s function. The dyadic Green’s function in a homogeneous environment can be expressed in terms of vector functions M, N, and L. The Green’s function is singular when the field and observation points coincide. In this case, Green’s function can be written as the main part plus the portion proportional to the impulse function. This type of decomposition of the Green’s function is not unique and depends on the shape of the volume that separates the environment [13]. In general, to obtain the dyadic Green’s function in a layered medium, the Green’s function of the homogeneous environment can be used. Then, the effect of inhomogeneity can be considered by adding reflected and transmitted waves to the dyadic Green’s function [14]. In a similar approach, called the scattering superposition method, the scattering Green’s function in each layer is expressed in terms of vector wave functions with the unknown coefficients that are obtained by applying the boundary conditions [15]. Moreover, to calculate the Green’s function of a medium by impedance method, instead of each dielectric layer a transmission line, and instead of each metallic layer a current source can be considered in the equivalent model in the rectangular, cylindrical, and spherical coordinates [15, 16, 17].

Considering medical diagnostics and treatment, planarly, cylindrically and spherically layered media are engineered mainly with plasmonic materials. In this regard, the interaction of an environment with a nanometer-scale dipole emitter is of interest in different biomedical fields. For example, the optical activity of the proteins can be investigated using Green’s tensor approach. Furthermore, a single excited molecule in the vicinity of a metallic structure can be potentially used in the sensors because of behaving as a resonant filter. Moreover, the point source is an appropriate model for the concentrated light sources which are used in medical applications [18]. Following these trends, different plasmonic structures have been presented theoretically and realized in the real environment. For example, the effect of the size of the gold nanoparticles on the decay rate and the energy transfer of dipole emitters is investigated using Green’s tensor formulation and compared with the results obtained from the fabrication [19]. The same analysis is carried out for a metallic cylinder coated with a dielectric layer and the resulted integrals are solved numerically [20]. Also, Fermat’s golden rule is used to connect the imaginary part of the Green’s function to the radiation impedance of the dipole antenna adjacent to a medium, and the Purcell factor is extracted for the planar meta-surface, which is then experimentally characterized in the microwave frequency band [21].

Recently, graphene’s plasmons are proposed as the low loss and reconfigurable alternative to the plasmons of the noble metals. There are two approaches to use graphene in analytical and numerical methods. In the first one, the graphene boundary is modeled with a 2D surface, characterized by its surface conductivity, whereas in the second one, the graphene layer is replaced by a very thin dielectric [22]. Although the latter can be analyzed using the available formulas previously presented for the multilayered dielectric structures in the literature [13, 23], graphene analysis using surface conductivity model has many advantages. First, in the dielectric model, each graphene interface adds an extra layer to the structure. Therefore, when the number of graphene layers is large, the problem becomes very complicated. Secondly, in the dielectric model, it is necessary to compute the special functions of the cylindrical and spherical coordinates with complex arguments, which requires the implementation of specific algorithms for their effective calculation [24], while in the impedance boundary condition method, the surface conductivity of graphene appears as a coefficient for special functions. Third, in the dielectric model, the thickness of the graphene layer is considered to be about 0.335 nm, which is often very small compared to other geometrical parameters and makes the convergence of the analytical functions slow [25]. Also, if the goal is using the dielectric model in numerical methods, it is necessary to use a dense mesh for the equivalent dielectric of the graphene that is not optimal in terms of time and memory [26].

Applying the graphene surface conductivity model for the derivation of the Green’s function has been considered in recent years. For example, for the graphene sheet under electric bias, the dyadic Green’s function is derived by the Hertzian potential and plane-wave expansion methods and the corresponding integrals are solved with the saddle point method [27, 28]. Also, the method is expanded for the analysis of graphene with tensor surface conductivity which can be used to analyze graphene with magnetic bias or spatial dispersion. Romberg’s integration procedure is proposed for the numerical solution of the resulting integrals [29]. In another research, the analysis of the electric dipole in the proximity of the parallel plate waveguide with graphene walls is studied by extracting the Green’s function and calculating the spontaneous emission. It is found that symmetric and asymmetric plasmonic modes lead to a sharp increase in this parameter [30]. As another instance, a point source is taken into account in the vicinity of the infinite cylinder with graphene cover and it is observed that by changing the distance of the source from the cylinder as well as changing the chemical potential of the graphene layer, the Fano resonances can be controlled [31]. In this chapter, the dyadic Green’s function of various planar, cylindrical, and spherical geometries with impedance boundary conditions will be calculated using the scattering superposition method. Specifically, we have focused on the graphene material due to its wide range of applications. Apparently, another 2D material can be considered by replacing the graphene surface conductivity with the surface conductivity of the desired material. The presented formulas can be potentially sued to design various biomedical devices. Moreover, by approaching the surface impedance to zero, these structures can be potentially used to investigate the interaction of the human body with different electromagnetic sources. For the Green’s function calculation of complex media in different coordinates using vector wave functions, the reader is referred to [32, 33, 34, 35].

Advertisement

2. Surface conductivity of graphene material under different conditions

Graphene is a two-dimensional material made of carbon atoms and can be considered in the solution of Maxwell’s equations using surface conductivity boundary condition [23]. Depending on the geometrical and optical conditions, graphene surface conductivity can be isotropic or anisotropic. The purpose of this section is to provide an overview of the graphene surface conductivity under different conditions (electric bias, magnetic bias, and spatial dispersion) and for different geometries (continuous or patterned sheets).

MethodQualityArea
Mechanical exfoliation of highly ordered pyrolytic graphite (HOPG)Very highSmall
Epitaxial growth of graphene on silicon carbide (SiC)MediumLarge (3–4 inches wafers)
Reduction of graphene oxide (rGO)MediumLarge
Chemical vapor deposition (CVD)HighVery large (30 inches)

Table 1.

Comparison of the quality and area of synthesized graphene using different techniques [36].

It should be noted that graphene material is mainly synthesized through four methods, including 1) mechanical exfoliation of highly ordered pyrolytic graphite (HOPG), 2) the epitaxial growth of graphene on silicon carbide (SiC), 3) the reduction of graphene oxide, and 4) chemical vapor deposition (CVD) technique. The comparison of these methods in terms of quality and the fabricated area is provided in Table 1 [36].

In this regard, a monolayer graphene film with metallic electrodes is transferred to a high impedance surface (HIS) which is realized by the printed circuit board (PCB) technology at microwave frequencies. The measurement is conducted in a small microwave chamber [37]. In another research in the same spectrum, an infrared laser is used to etch the CVD grown graphene sheet with predefined periodicity and later investigate the absorption of the designed structure inside a rectangular waveguide [38]. The electron beam lithography is another approach used for pattering the graphene sheet for enhanced light matter interaction [39]. Moreover, a transparent graphene millimeter wave absorber constructed by multiple transfer-etch processing is characterized by reflectometery technique at 140 GHz [40]. Also, CVD-grown graphene is used to enhance the sensitivity of the surface-enhanced Raman spectroscopy (SERS)-based chemical sensor. The measurement is done using a Raman spectrometer at the laser wavelength of 785 nm (red) [41]. In another sensor chip, DNA is hybridized to the graphene-based substrate under UV light with the wavelength of 260 nm. In this sensor, the atomic force microscopy (AFM) is used to ensure the continuity and uniformity of the synthesized graphene, and Raman characterization is used to investigate its quality and the number of layers [42].

2.1 Graphene material under electric bias

When a graphene sheet is under electric bias, its surface conductivity is isotropic and can be approximately calculated by using Kubo’s formulas as [43]:

σintra=2ie2kBT2πω+i/τln2coshμc2kBTE1
σinter=e24×12+1πarctanω2μc2kBTi2πlnω+2μc2ω2μc2+2kBT2E2

In the above equations, T is the temperature, μc is the chemical potential of graphene, is the reduced Planck’s constant, KB is the Boltzmann’s constant, and ω is the angular frequency. At low-THz frequencies, the inter-band contribution of the surface conductivity can be neglected. Also, the graphene layer can be modeled as a dielectric with very low thickness δ and equivalent dielectric constant ε=1iσωε0δ [22]. Note that in the above equations, graphene is assumed to be in the linear region, otherwise, other terms proportional to 1/ω3 and 1/ω4 should be added respectively for the frequency range of ω<2μc and ω2μc [44].

To increase the light-matter interaction of graphene material, the nano-pattering method has been proposed [45]. The surface conductivity of the periodic graphene elements under the electric bias is also isotropic. For the square patches shown in Figure 1(a), by considering the periodicity of D0 and the air gap distance of g, closed-form surface impedance (inverse of the surface conductivity) is as [47]:

Figure 1.

(a) Graphene square patches with the periodicity of D0 and the air gap distance of g and (b) densely packed graphene strips with the width W and the periodicity L [46].

Zs=D0σsD0g+iπ2ωε0ε2+12D0lncscπg2D0E3

The effective surface conductivity of periodic graphene elements with arbitrary pattern can be measured or extracted using the parameter retrieval method through full-wave simulation [48]. It should be noted that for computing the electric field required for each considered chemical potential, approximate equations can be derived as [49]:

μceV=λ1E0λ2,E0V/nm0λ1E0λ2,E0V/nm<0E4

where, λ1=0.3677 and λ2=0.5010. For the chemical potentials in the range of [−1,1] eV, the required bias fields are in the order of several volts per nanometer, which can be implemented practically [29].

The surface conductivity of densely packed graphene strips, as illustrated in Figure 1(b), is anisotropic and can be approximated in the form of a diagonal tensor using the effective medium formulation as [50]:

σxx=WσσCLσC+σyy=σWLσxy=σyx=0E5

In the above equations, W and L are the width and periodicity of the strips, respectively. Also, σ is the surface conductivity of graphene under electric bias, and σc is the static conductivity of the surface. Two very important properties of this environment are the existence of near-zero surface conductivity and hyperbolic dispersion region with the potential applications in 2D lens structures and the spontaneous emission enhancement of the dipole emitters [50, 51].

2.2 Graphene sheet with spatial dispersion effects

When the graphene sheet is placed on a substrate with a high dielectric constant, its surface conductivity is a tensor in which the elements depend on the wave propagation constant in the structure. This is called the spatial dispersion effect and the associated formulae for calculating surface conductivity are [29]:

σxx=σ+αd2dx2+βd2dy2E6
σyy=σ+βd2dx2+αd2dy2E7
σxy=σyx=2βd2dxdyE8

The parameters σ, α, and β are extracted for an unbiased sheet μc=0 thorough perturbation theory [29]. The resulting equations are valid for the electrically biased sheet μc0, as well [52].

2.3 Graphene sheet under magnetic bias

When a graphene sheet is under magnetic bias, its surface conductivity is also a tensor. The diagonal elements of this tensor are equal and the off-diagonal elements are opposite in sign defined as [53].

σxxωB0=σ01iωτωcτ2+1iωτ2E9
σyxωB0=σ0ωcτωcτ2+1iωτ2E10

where σ0=2e2τπ2kBTln2coshμc2kBT and ωc=eB0νF2μc . The approximate formulas for the calculation of the surface impedance of the square graphene elements under magnetic bias, shown in Figure 1(a), are as follows [54]:

Z¯¯p=FGZ¯¯g+i2αμ0ε0εeff1001E11

where FG=0.6D0/D0g3+0.4 and α=k0D0εeffπlnsinπg2D0. Similar to the electrically biased patterned elements with arbitrary shapes, the parameter retrieval method can be used under the applied magnetic bias [55]. Given the discussion of the above three sections, it is observed that assuming the surface conductivity of graphene as:

σ¯¯=σxxσxyσyxσyyE12

All items expressed above can be extracted as a special case.

Advertisement

3. Analysis of graphene-based structures using dyadic Green’s function

Dyadic Green’s functions for the planarly, cylindrically, and spherically layered structures with graphene interfaces will be derived in this section. To this end, the boundary conditions of the continuity and discontinuity of tangential electric and magnetic fields are respectively satisfied regarding the considered surface conductivity model for the graphene.

3.1 Graphene-based planarly layered media

This section aims to obtain dyadic Green’s functions for planar structures with graphene boundaries. This problem can be solved either in the rectangular or cylindrical coordinates. In the first sub-section, a graphene sheet with the tensor surface conductivity boundary condition (TSCBC) is considered and its dyadic Green’s function is calculated in the rectangular coordinates. In this case, the anisotropy of the surface impedance causes the coupling of the transverse electric (TE) and transverse magnetic (TM) fields. As a result of coupling, the number of unknown coefficients in the expansion of the dyadic Green’s function is increased concerning the electrically biased sheet. In the second part of this section, a graphene-dielectric stack with an arbitrary number of layers is investigated considering the electric bias for the graphene sheets. This problem is solved in the cylindrical coordinates to simplify the calculation of the resulted Sommerfeld integrals.

3.1.1 Graphene sheet with the tensor surface conductivity boundary condition

The purpose of this section is to obtain the dyadic Green’s function of a graphene sheet with the tensor surface conductivity in the interface of half-spaces, as shown in Figure 2(a). The constitutive parameters of the top and bottom regions are considered as (ε1, μ1) and (ε2, μ2), respectively. Without losing the generality of the problem, the source is assumed to be in the first environment and the graphene boundary is considered in z = 0 interface. Dyadic Green’s function of this structure will be calculated using the scattering superposition method. For this purpose, the Green’s function in each region of the problem is written in the form of the Green’s function in the absence and presence of structure. Thus [23]:

Figure 2.

(a) Graphene sheet with the tensor surface conductivity boundary condition at the interface of half-spaces [46] and (b) graphene-dielectric stack.

G¯¯e11=G¯¯e0R¯R¯+G¯¯es11R¯R¯z>0E13
G¯¯e21=G¯¯es21R¯R¯z<0E14

where G¯¯es11 and G¯¯es21 are respectively the scattering Green’s function in regions 1 and 2 and they are expanded in terms of M¯ and N¯ vector wave functions. Also, G¯¯e0 is the free-space Green’s function which can be computed using the Gm method as:

G¯¯e0R¯R¯=1k12ẑẑδR¯R¯+dkxdkyC1×M¯h1M¯h1+N¯h1N¯h1z<zE15

where C1=i8π2h1kx2+ky2 . Using the scalar wave function ψk¯=expikxx+ikyy+ikzz, it can be readily found that:

M¯k¯=ikyx̂kxŷψk¯E16
N¯k¯=kzkjkxx̂+kyŷ+kx2+ky2kjẑψk¯E17

The parameters kx, ky, and kz are the wavenumbers in x, y, and z directions, respectively, and kj shows the wavenumber for j = 1, 2. These parameters are not independent and are related to each other via kz=±kj2kx2ky2=±kj2kρ2 . The vector wave function M¯ represents the electric field of TE modes and the vector wave function N¯ shows the electric field of the TM waves. In the structure under consideration, the anisotropy of surface conductivity leads to the coupling of TE and TM fields. Therefore [56]:

G¯¯es11R¯R¯=dkxdkyC1×a1M¯h1+a1N¯h1M¯h1+b1N¯h1+b1M¯h1N¯h1E18
G¯¯es21R¯R¯=dkxdkyC1×a2M¯h2+a2N¯h2M¯h1+b2N¯h2+b2M¯h2N¯h1E19

The unknown coefficients a1, a1,b1,b1,a2, a2, b2, and b2 will be obtained by applying the boundary conditions. Using the self and mutual orthogonality of the vector wave functions, the above equations can be divided into two systems of equations, each with four unknown coefficients. The boundary conditions on the electric and magnetic Green’s functions respectively state that:

ẑ×G¯¯e11G¯¯e21=0E20
ẑ××G¯¯e11iωμ1×G¯¯e21iωμ2=σ¯¯.G¯¯e21E21

After applying the above boundary conditions and removing the coupling effect from the tangential components of the electric field, and by defining, A=σxxkx+σxyky, B=σyxkx+σyyky, C=σxxkyσxykx, and, D=σyykxσyxky, the unknown coefficients of the TE waves, a1=ΔTE1ΔTE and a1=ΔTE2ΔTE, can be obtained as [46]:

ΔTE=ikρ2P+Q++P+Akx+Bkyh1k1+Q+Cky+Dkx+ω2ih1k1BC+ADE22
ΔTE1=ikρ2Q+P+Akx+BkyPh1k1ωiCky+DkxQ+ω2ih1k1BC+ADE23
ΔTE2=ω2h1μ1DkyCkxE24

where, P+=h1μ1+h2μ2, Q+=k1μ1+k22h1k1h2μ2, P=h1μ1h2μ2, and Q=k1μ1k22h1k1h2μ2. By defining the TM waves expansion coefficients as b1=ΔTM1ΔTM and b1=ΔTM2ΔTM, it can be shown:

ΔTM=ikρ2Q+P++Q+Cky+Dkx+h1k1P+Akx+Bky+iω2h1k1AD+BCE25
ΔTM1=ikρ2QP++h1k1P++Akx+BkyDkx+CkyQ+iω2h1k1AD+BCE26
ΔTM2=2ωh1μ1AkyBkxE27

Other unknown coefficients can be obtained using decoupling equations in [46]. To validate the obtained coefficients, the structure of Figure 1(a) consisting of square patches with D0 = 5 μm, g = 0.5 μm, μc = 0.5 eV and τ = 0.5 ps with plane wave illumination is considered under electric and magnetic biases. Since Green’s function coefficients are the same as reflection and transmission coefficients of the plane wave, the results of Green’s function are compared with the results of the circuit model as [54]:

S31=22+η0σd2+η0σd2+η0σ02E28
S41=2η0σ02+η0σd2+η0σ02E29

Figure 3 shows the magnitude of the transmission coefficient by considering the electric bias for the graphene layer. The results of the two methods are identical, and because of the absence of electromagnetic coupling under electric bias, the transmission coefficient due to mutual coupling is zero. In Figure 4 the same results are illustrated for the applied magnetic bias of 0.5 Tesla. There is good agreement in the magnitude and phase of the transmission coefficient in the abovementioned two methods. Also, by finding the poles of the coefficients, the electromagnetic wave propagation constants for the electrically and magnetically biased graphene sheets can be obtained which are in full agreement with [28, 57], respectively. The correctness of the extracted coefficients confirms the validity of the dyadic Green’s function formulation.

Figure 3.

The magnitude of the transmission coefficient for the graphene nano-patch with the parameters D0 = 5 μm, g = 0.5 μm, μc = 0.5 eV, and τ = 0.5 ps [46].

Figure 4.

(a) The magnitude and (b) phase of the transmission coefficient for the graphene nano-patch with D0 = 5 μm, g = 0.5 μm, μc = 0.5 eV, ps. τ =0.5, and B0 = 0.5 T [46].

3.1.2 Graphene-dielectric stack

Dyadic Green’s function for an N-layer dielectric environment has been previously formulated using the scattering superposition method [58]. In this section, the above equations are extended to the environment with the electrically biased graphene boundaries, as shown in Figure 2(b). The graphene boundary can be either continuous or periodically patterned as discussed in section 2. To start the analysis, the layers are numbered by starting from the top layer, and an arbitrary field point i and source point j are assumed. The problem is solved in the cylindrical coordinates. Since the cylindrical wave functions are discussed in detail in the next section, they are not mentioned here. The dyadic Green’s function can be expanded as [58]:

G¯¯esijrr=i4πdhn=02δn0λhj1δiNM¯hi1δj1AMijM¯hj+1δjNBMijM¯hj+1δiNN¯hi1δj1ANijN¯hj+1δjNBNijN¯hj+1δi1M¯hi1δj1CMijM¯hj+1δjNDMijM¯hj+1δi1N¯hi×1δj1CNijN¯hj+1δjNDNijN¯hjE30

In the above equations, M¯ and N¯ are vector wave functions in the cylindrical coordinate system, and AM,Nij, BM,Nij, CM,Nij, DM,Nij are unknown coefficients. The Kranoker delta function is used in the field expansion to generalize the formulation for the arbitrary locations of the field and source points. By applying the boundary conditions on the tangential components of the electric field at the interface of the arbitrary two layers, denoted by i and i + 1, it can be shown that:

AMijBMijOiANijOiBNijeihizi+CMijDMij+δjiOiCNijOiDNij+δjieihizi=AMi+1j+δi+1jBMi+1jOi+1ANi+1j+δi+1jOi+1BNi+1jeihi+1zi+CMi+1jDMi+1jOi+1CNi+1jOi+1DNi+1jeihi+1ziE31

where Oi=hiki,Pi=hiμi,Qi=kiμi. The boundary condition on the tangential components of the magnetic field yields:

Pi+1AMi+1j+δi+1jPi+1BMi+1jQi+1ANi+1j+δi+1jQi+1BNi+1jeihi+1ziPi+1CMi+1jPi+1DMi+1jQi+1CNi+1jQi+1DNi+1jeihi+1ziPiAMijPiBMijQiANijQiBNijeihizi+PiCMijPiDMij+δjiQiCNijQiDNij+δjieihizi=iωσi+1iAMijBMijPiANijPiBNijeihizi+iωσi+1iCMijDMij+δjiPiCNijPiDNij+δjieihiziE32

By re-writing the coefficients as a matrix:

AM,Ni+1j+δi+1jBM,Ni+1jCM,Ni+1jDM,Ni+1j=1TFiH,VRFiH,VTFiH,VRPiH,VTPiH,V1TPiH,VAM,NijBM,NijCM,NijDM,Nij+δijE33

the outgoing and incoming reflection and transmission coefficients can be defined and used to extract the recursive relations. The coefficients for M¯ sources are:

RFiH=μihi+1μi+1hi+gμihi+1+μi+1hi+gE34
RPiH=μihi+1μi+1higμihi+1+μi+1higE35
TFiH=2μihi+1μihi+1+μi+1hi+gE36
TPiH=2μihi+1μihi+1+μi+1higE37

The coefficients for N¯ sources are:

RFiV=μihiki+12ki2μi+1hi+1+ghihi+1μihiki+12+μi+1hi+1ki2+ghihi+1E38
RPiV=μihiki+12μi+1ki2hi+1ghihi+1μihiki+12+μi+1ki2hi+1ghihi+1E39
TPiV=2μi+1kiki+1hi+1μihiki+12+μi+1ki2hi+1ghihi+1E40
TFiV=2μihi+1kiki+1μihiki+12+μi+1hi+1ki2+ghihi+1E41

The superscripts H and V respectively denote TE and TM sources. Also, subscripts F and P are used to show the incoming and outgoing waves, respectively. The procedure of extracting the unknown coefficients using (34)-(41) is discussed in [58]. To validate the proposed formulas, a parallel plate waveguide with graphene walls is considered. To extract the characteristic equation using the proposed formulations, it is necessary to force the denominator of the coefficients equal to zero. For this three-layer medium:

T1=T2.T1=.1TF2Heih2h1dRF2HTF2Heih2h1dRP2HTP2Heih2+h1d1TP2Heih2h1d.1TF1HRF1HTF1HRP1HTP1H1TP1HE42

From which it can be deduced that BM,N11=T121T111. By setting T111 equal to zero and assuming that the medium (1) and (3) are the same, and also defining h=iωμ1σ, for the H coefficients it can be concluded that:

eih2d1eih2d+1h1h2+h=0h2+h+jtanh2d2h1=0E43

This procedure is repeatable for V sources. Also, to calculate the reflection coefficient from a multilayer structure, it is necessary to consider the field and source points in region 1. In this case, the only non-zero coefficient in Green’s function expansion is BM,N11 coefficient representing the plane wave reflection coefficient from the multilayer structure.

3.2 Graphene-based cylindrical structures

In this section, the dyadic Green’s function of a cylindrical structure with the tensor surface conductivity boundary condition will be extracted. Later, different examples of guiding and scattering problems are provided to investigate the validity of the formulation. In general, in cylindrical structures, TE and TM modes are coupled, which leads to the complexity of mathematical relations. Therefore, the generalization of the formulation to the multilayered cylinders is not considered here. Note that graphene sheets can be wrapped around cylindrical particles due to the presence of van der Waals force [59]. For this purpose, tape-assist transfer under micromanipulation and spin-coating methods are proposed [60].

3.2.1 Dyadic Green’s function for a cylinder with tensor surface conductivity boundary condition

The dyadic Green’s function of the single-layer cylinder with the tensor surface conductivity boundary condition, as considered in Figure 5., will be extracted in the following. The interior region of the cylinder is made of dielectric material and its cover is considered as a full tensor surface conductivity. To solve the problem, the vector wave functions are defined as [23]:

Figure 5.

(a) A monolayer cylinder with a 2D cover with the tensor surface conductivity boundary condition and (b) its special cases constructed by densely packed strips and square nano-patches [61].

M¯μh=eimϕeihzimZmμrrr̂Zmμrrϕ̂E44
N¯μh=1kjeimϕeihzihZmμrrr̂hmrZmμrϕ̂μ2ZmμrẑE45

In the above equations, Zm is the cylindrical Bessel function in the inner layer and the cylindrical Hankel function in the outer layer, both with the orders of m. The wavenumber in the radial direction is μ and the wavenumber along the length is h. The free-space Green’s function in the cylindrical coordinates is [23]:

G¯¯e0R¯R¯=1k2r̂r̂δR¯R¯+i8πdhm=1η2×M¯ηhM¯η1h+N¯ηhN¯η1hR<RE46

In the cylindrical coordinate system, TE and TM modes are coupled and the expansion of the fields are [23].

G¯¯es11R¯R¯=i8πdhm=1η2×AηM¯η1h+BηN¯η1hM¯η1h+CηN¯η1h+DηM¯η1hN¯η1hE47
G¯¯es21R¯R¯=i8πdhm=1η2×aξM¯ξh+bξN¯ξhM¯η1h+cξN¯ξh+dξM¯ξhN¯η1hE48

where Aη, Bη, Cη, and Dη are the unknown coefficients of the DGF expansion in region 1. Also, aξ, bξ, cξ, and dξ are DGF expansion coefficients in the region 2. The boundary condition on the tangential components of the magnetic Green’s function is given by [61]:

r̂××G¯¯e11μ1×G¯¯e21μ2=σ¯¯ϕz.G¯¯e21E49

By applying the above-mentioned boundary condition along with the boundary condition regarding the continuity of the electric Green’s function to (47)(48), the system of equations to determine the unknown coefficients can be obtained as:

Hm1ηaa1k1hmaHm1ηaJmξaa1k2hmaJmξa01k1η2Hm1ηa01k2ξ2Jmξa1μ1hmaHm1ηak1μ1Hm1ηaa1μ2hmaJmξa+σ1k2μ2Jmξaa+σ21μ1η2Hm1ηa01μ2ξ2Jmξa+σ3σ4×AηBηaξbξ=Jmηaa01μ1hmaJmηa1μ1η2JmηaE50
1k1hmaHm1ηaHm1ηaa1k2hmaJnξaJmξaa1k1η2Hm1ηa01k2ξ2Jmξa0k1μ1Hm1ηaa1μ1hmaHm1ηak2μ2Jmξaa+σ1'1μ2hmaJmξa+σ2'01μ1η2Hm1ηaσ3'1μ2ξ2Jmξa+σ4'×CηDηcξdξ=1k1hmaJmηa1k1η2Jmηak1μ1Jmηaa0.E51

where:

σ1=iωσJmξaa,σ2=σzzk2ξ2Jmξa+iωσhmk2aJmξa
σ3=iωσϕϕJmξaa,σ4=iωσϕϕ1k2hmaJmξa+iωσϕz1k2ξ2Jmξa
σ1=iωσzz1k2ξ2Jmξa+iωσhmk2aJmξa,σ2=iωσJmξaa
σ3=iωσϕϕ1k2hmaJmξa+iωσϕz1k2ξ2Jmξa,σ4=iωσϕϕJmξaa

Nullifying the determinant of the matrices (50)(51) for m = 0, each of the above matrices can be separated as the multiplication of:

ωε2ξJ1ξaJ0ξaωε1ηH11ηaH01ηa+iσd=0E52
ξωμ2J0ξaJ1ξaηωμ1H01ηaH11ηa+iσd=0E53

For the graphene shell under magnetic bias, both matrices in (50)(51) result in the following equation for the propagation constant of hybrid TE and TM waves:

ωε2ξH11ξaH01ξaωε1ηJ1ηaJ0ηa+iσd×ηωμ1J0ηaJ1ηaξωμ2H01ξaH11ξa+iσd=σ02E54

Which is in agreement with [57]. Also, the total scattering cross-section (TSCS) of a graphene-coated cylinder with the parameters a = 50 μm, ε2 = 2.4, τ = 1 ps, μc = 0.25 eV under electric bias is calculated for both TE and TM polarizations in Figure 6, and compared with the results of the CST2017 software package. As can be seen, both methods have resulted in the same results.

Figure 6.

TSCS for an infinite cylinder covered with electrically biased graphene shell with a = 50 μm, ε2 = 2.4, τ = 1 ps, and μc = 0.25 eV considering (a) TE and (b) TM polarization for the incident wave [61].

As another example, the densely packed graphene strips with the parameters L = 420 nm, W = 400 nm, μc = 0.5 eV, and τ = 1 ps are considered around the dielectric cylinder as in Figure 5. It is assumed that the strips are wrapped around a hollow cylinder with the radius of a = 50 μm. Figure 7 shows the TSCS of this structure for the magnetic biases with the strength in the range of 20-40 T. As observed, by increasing the magnetic bias, the resonant frequency of the surface plasmons blue shifts. The associated planar structure behaves as a hyperbolic meta-surface [62]. Under locally flat consideration of the curvature, this structure can also be considered as a hyperbolic medium. In the cylindrical geometries, hyperbolic meta-surfaces can be obtained using graphene-dielectric stacks [63]. The advantage of this hyperbolic structure is its two-dimensional nature and reconfigurability. It is also demonstrated that covering the surface of nanotubes with the hyperbolic meta-surface increases the interaction of the light with dipole emitters [64].

Figure 7.

TSCS for an infinite length hollow cylinder (a = 50 μm) coated with densely packed graphene strips with L = 420 nm, W = 400 nm, μc = 0.5 eV, and τ = 1 ps considering different magnetic biases [61].

Finally, as Figure 5(b) illustrates, graphene-based square patches around the cylinder are considered under magnetic bias. Geometrical and optical parameters are as follows: τ = 1 ps, g = 0.5 μm, and D = 0.5 μm and the TSCS is illustrated in Figure 8. for B0 = 0 T and B0 = 10 T. As can be seen, by changing the magnetic bias, the optical state changes from the maximum scattering to the minimum scattering. Such capability has recently been proposed by using a phase change material for switching between these two situations [65]. In this structure, the operating frequency can also be adjusted by changing the electric bias of graphene.

Figure 8.

TSCS for an infinite cylinder coated with graphene patches with D = 0.5 μm, g = 0.5 μm, and τ = 0.5 ps under different electric and magnetic biases [61].

It is essential to note that in the cylindrical structures, the spatial domain Green’s function consists of an integral on the real axis. Due to the existence of poles in the integration path, the integration path is usually deformed into a triangular shape. Also, a common method for calculating the spatial domain Green’s function is the generalized pencil of function (GPOF) method in which the Green’s function is expanded in terms of the complex exponential functions [66]. The unknowns can be found via the algorithm provided in [67]. Also, as mentioned earlier, the dyadic Green’s function of the graphene-based multilayered cylindrical structures will not be considered here. For calculating the scattering cross-section of graphene-based multilayer cylindrical structures, a simple approach based on the transfer matrix method (TMM) is proposed that will be suitable for establishing novel optical devices [68, 69].

3.3 Graphene-based spherically layered medium

In this section, a multi-layered spherical structure with the graphene boundaries is considered and its Green’s function is extracted by assuming different locations for the source and observation points. The relationship between the Green’s function expansion coefficients and the modified Mie-Lorentz coefficients is exhibited to discuss how to solve the scattering problems using the Green’s function. Scattering analysis of graphene-based layered structures is of great importance in the design of novel optical devices [70]. Finally, the procedure for calculating the Purcell factor is considered as an important application. Instances of experimentally realized graphene-coated spherical particles can be found in [71, 72], where improved template method and hydrothermal method are proposed for the synthesis. Also, transmission electron microscopy (TEM) and field emission scanning electron microscopy (FE-SEM) are used for characterization.

3.3.1 Dyadic Green’s function of a graphene-based spherically layered structure

Let us consider an N-layer spherical medium with the graphene boundaries as shown in Figure 9. The purpose of this section is to compute the dyadic Green’s function of this structure with the assumption of arbitrary locations for the field and source points. For this purpose, the Green’s function in each layer is expanded in terms of vector wave functions with unknown coefficients. These functions are calculated using the scalar wave function of [74]:

Figure 9.

Spherically layered medium with graphene boundaries (a) 2D and (b) 3D views [73].

φmnrθϕ=znkprPnmcosθeimϕE55

where zn. represents the spherical Bessel or Hankel functions of order n and Pnm. is the associated Legendre function with degree n and order m. It can be readily shown that:

M¯mnkp=znkpreimϕimsinθPnmcosθθ̂dPnmcosθϕ̂E56
N¯mnkp=nn+1kprznkprPnmcosθeimϕr̂+1kprdrznkprdrdPnmcosθθ̂+imsinθPnmcosθϕ̂eimϕE57

The above vector functions are self and mutually orthogonal. This feature will be used to decouple the equation when computing the unknown coefficients.

As mentioned earlier, in the scattering superposition method, the dyadic Green’s function is written as the sum of the free-space and scattering Green’s functions. The free-space Green’s function is related to the source in an infinite homogeneous medium while the scattering Green’s function is due to the source in the presence of the layered medium. The expansion of the Green’s function for the spherical structure with N concentric layers, assuming that the source and field points are respectively located in the desired layers with the labels p and q, can be written as [23]:

Ĝ¯¯epqrr=Ĝ¯¯0epqrrδqp+Ĝ¯¯espqrrE58

The free-space Green’s function can be obtained using the residue theorem as [23]:

Ĝ¯¯0err=r̂r̂kq2δrr+ikq4πn=1m=nn2δm02n+1nn+1×nm!n+m!×M¯mn1kqM¯mnkq+N¯mn1kqN¯mnkqr>rM¯mnkqM¯1mnkq+N¯mnkqN¯1mnkqr<rE59

Moreover, scattering Green’s function in each layer can be expanded as [75]:

Ĝ¯¯espqrr=ikq4πn=1m=nn2δm02n+1nn+1nm!n+m!×1δpNM¯mn11δq1AHpqM¯mn+1δqNBHpqM¯1mn+1δpNN¯mn11δq1AVpqN¯mn+1δqNBVpqN¯1mn+1δp1M¯mn1δq1CHpqM¯mn+1δqNDHpqM¯1mn+1δp1N¯mn1δq1CVpqN¯mn+1δqNDVpqN¯1mnE60

where AH,Vpq, BH,Vpq, CH,Vpq, and DH,Vpq are the unknown coefficients of the Green’s function to be obtained. In the above formulation, the superscript (1) represents that the spherical Hankel functions of the first type are chosen to represent the spherical vector functions. For the other vector wave functions, the first-kind Bessel function should be selected. As observed, in the spherical structures the TE and TM waves are not coupled. Thus, the expansion of the fields only includes the interaction of the functions M¯ and M¯′ as well as the functions N¯ and N¯′. Moreover, to solve the problem for the structures with the arbitrary number of layers, the Kronecker delta function is used in the expansions of the fields. In the middle layers of the spherical structure, the electric field is a linear combination of the Bessel and Hankel functions, whereas in the outermost layer, only the Hankel functions, and in the innermost layer, only the Bessel functions will exist. So we have used 1δp1 and 1δpN functions. To determine the delta functions related to the source terms, free-space Green’s function can be used. For q = 1 (external layer), the source functions related to the second criterion of the free-space Green’s function is used. Also, for q = N (internal layer) the source functions related to the first criterion should be used. In the middle layers, a linear combination of different source functions must be used.

It should be noted that by using the addition theorem in Legendre functions, the internal series in the Green’s function expansion can be eliminated [11]. To select the number of terms required for the convergence of the external series, the conditions of the problem must be considered. In the other words, in the structures whose electrical size is very much smaller than that of the wavelength, the term n = 1 is sufficient for the convergence [76]. Also, in the case where the distance between the source and observation points is large, the series can be truncated in the number Nt=x+3x+2 , where x = k0R1. Otherwise, the convergence of the series is weak, and a large number of terms in the range of 20 k0R1 should be considered. In this case, the series acceleration techniques will be highly efficient in terms of computational efficiency [77, 78].

Boundary conditions on tangential components of electric and magnetic Green’s functions in the interface of two adjacent layers are [73]:

r̂×Ĝ¯¯epq=r̂×Ĝ¯¯ep+1qE61
1μp+1r̂××Ĝ¯¯ep+1q1μpr̂××Ĝ¯¯epq=σp+1pr̂×r̂×G¯¯epqE62

which respectively lead to the following equations:

ξnppAHpqξnppBHpqξnppAVpqξnppBVpq+ψnppCHpqψnppDHpq+δpqψnppCVpqψnppDVpq+δpq=ξnp+1pAHp+1q+δp+1qξnp+1pBHp+1qξnp+1pAVp+1q+δp+1qξnp+1pBVp+1q+ψnp+1pCHp+1qψnp+1pDHp+1qψnp+1pCVp+1qψnp+1qDVp+1qE63
kp+1μp+1ξnp+1pAHp+1p+δp+1qξnp+1pBHp+1pξnp+1pAVp+1q+δp+1qξnp+1pBVp+1q+kp+1μp+1ψnp+1pCHp+1qψnp+1pDHp+1qψnp+1pCVp+1qψnp+1pDVp+1qkpμpξnppAHpqξnppBHpqξnppAVpqξnppBVpqkpμpψnppCHpqψnppDHpq+δpqψnppCVpqψnppDVpq+δpq=σp+1pξnppAHpqξnppBHpqξnppAVpqξnppBVpq+σp+1pψnppCHpqψnppDHpq+δpqψnppCVpqψnppDVpq+δpqE64

In the above equations, ψnpq=jnkpRq, ξnpq=hn1kpRq, ψnpq=1/ρdρjnρρ=kpRq and ξnpq=1/ρdρhn1ρρ=kpRq. To obtain recursive relations for unknown coefficients of Green’s function, separating the above equations is necessary. Therefore [75]:

AH,Vp+1q+δp+1qBH,Vp+1q=1TFpH,VAH,VpqBH,Vpq+RFpH,VTFpH,VCH,VpqDH,Vpq+δpqE65
CH,Vp+1qDH,Vp+1q=RPpH,VTPpH,VAH,VpqBH,Vpq+1TPpH,VCH,VpqDH,Vpq+δpqE66

where the reflection and transmission coefficients for TE waves are computed as:

RFpH=kp+1μpψnp+1pψnppkpμp+1ψnppψnp+1p+gψnppψnp+1pkp+1μpψnp+1pξnppkpμp+1ξnppψnp+1p+gξnppψnp+1pE67
RPpH=kp+1μpξnp+1pξnppkpμp+1ξnppξnp+1p+gξnppξnp+1pkp+1μpξnp+1pψnppkpμp+1ψnppξnp+1p+gψnppξnp+1pE68
TFpH=kp+1μpψnp+1pξnp+1pψnp+1pξnp+1pkp+1μqψnp+1pξnppkpμp+1ξnppψnp+1p+gξnppψnp+1pE69
TPpH=kp+1μpξnp+1pψnp+1pψnp+1pξnp+1pkp+1μpξnp+1pψnppkpμp+1ψnppξnp+1p+gψnppξnp+1pE70

Also, the reflection and transmission coefficients for the TM waves are:

RFpV=kp+1μpψnp+1pψnppkpμp+1ψnppψnp+1p+gψnppψnp+1pkp+1μpψnp+1pξnppkpμp+1ξnppψnp+1p+gξnppψnp+1pE71
RFpV=kp+1μpψnp+1pψnppkpμp+1ψnppψnp+1p+gψnppψnp+1pkp+1μpψnp+1pξnppkpμp+1ξnppψnp+1p+gξnppψnp+1pE72
TFpV=kp+1μpψnp+1pξnp+1pξnp+1pψnp+1pkp+1μpψnp+1pξnppkpμp+1ξnppψnp+1p+gξnppψnp+1pE73
TPpV=kp+1μpξnp+1pψnp+1pψnp+1pξnp+1pkp+1μpξnp+1pψnppkpμp+1ψnppξnp+1p+gψnppξnp+1pE74

In the above formulas, the subscripts F and P represent the outgoing and incoming waves, respectively. The symbols TPFpH and RPFpH express the transmission and reflection of TE waves (due to the presence of superscript H), whereas the expression TPFpV and RPFpV express the transmission and reflection of the TM waves (due to the presence of superscript V). Using the matrix form:

AH,Vp+1q+δp+1qBH,Vp+1qCH,Vp+1qDH,Vp+1q=1TFpH,VRFpH,VTFpH,VRPpH,VTPpH,V1TPpH,VAH,VpqBH,VpqCH,VpqDH,Vpq+δpqE75

Recursive relations can use to start through: AH,VNq=BH,VNq=CH,V1q=DH,V1q=0.

It should be noted that due to the difficulty of constructing multiple concentric graphene shells, one can consider each boundary as a dielectric or perfect electric conductor (PEC) in the optical design. For example, an optical absorber consisting of a metal-dielectric spherical resonator whose outermost layer is coated with graphene is shown to enhance the absorption of the resonator [79]. To design optical structures such as absorber and invisible cloaks using Green’s function formulation, it is necessary to transfer the point source to infinity to resemble a plane wave. In this case, considering the expansion of Green’s function, it can be observed that the only unknown coefficients of expansion are BH,Vpq and DH,Vpq coefficients. Using the convolution integral it can be shown that [23]:

G¯¯0e.J+G¯¯es11.J=G¯¯es21.JE76

The source dependency is the same in all of the above Green’s functions and can be simplified from both sides of the equation. It is observed that the equation is converted into the equation resulting from the Mie analysis of the spherically layered structures. By directly starting from the Mie-Lorentz theory, the same results can be obtained [80]. Moreover, due to the sub-wavelength nature of the localized graphene plasmons, the final formulas can be simplified using the polynomial approximation of the special functions [81]. These structures can also be used as the building blocks of optical meta-materials [82].

3.3.2 Purcell factor and energy transfer between donor-acceptor emitters

As mentioned earlier, one of the important applications of Green’s function is studying the interaction of dipole emitters in the vicinity of nanostructures. For this purpose, a vertical dipole in the vicinity of the graphene-based spherical structure, which was introduced in the previous section, is considered and its Purcell factor is calculated using the Green’s function. Assuming that the field point and observation points for the dipole with moment d0=d0r̂ are in the same location of r=Δ>R1, θ=0, and ϕ=0, by calculating the scattered field using the convolution integral and the use of [18]:

Γtotald0Γ0=1+6πε0k13d02Imd0.Escad0rE77

where symbol Im represents the imaginary part of the complex function. The decay rate can be calculated. For this purpose, using relationships:

Pnmcosθθ0sinmθ=1m=00m0E78
dPnmcosθθ0msinm1θ=1m=10m1E79

The scattered field can be calculated. Thus [73]:

ΓtotalΓ0=132nnn+12n+1znk1Δk1Δ2BV11E80

As can be seen, the above equation is in full agreement with [83] which has extracted the decay rate for a core-shell plasmonic sphere, whereas in this case, it is necessary to use the Mie coefficients of graphene-based structure, namely, BV11 . Using the derived formulas, the positions of the dipole can be considered arbitrarily. The transferred energy between the donor-acceptor pairs can be calculated straightforwardly using the Green’s function G as [84]:

ΓωΓ0ω=dA.GrArBω.dB2dA.G0rArBω.dB2E81

where the subscript 0 refers to the free-space parameters and dA and dB are respectively the dipole moments of the acceptor and donor.

Advertisement

4. Conclusion

In conclusion, dyadic Green’s function extraction for planarly, cylindrically, and spherically layered medium based on scattering superposition method is a unified approach to deal with a wide range of electromagnetic problems in the realm of biomedicine. Specifically, the interaction of the human skin, body, and head with the electromagnetic sources with arbitrary distributions can be studied. Moreover, by engineering the constitutive parameters of the layers, a variety of novel devices for medical diagnostics and treatment can be proposed. Plasmonic metals and 2D materials are two main categories of such materials. For the sake of efficient analytical analysis, the impedance boundary condition is satisfied in the case of 2D materials to be used in the design of compact devices.

References

  1. 1. J. R. Monds and A. G. McDonald, “Determination of skin temperature distribution and heat flux during simulated fires using Green’s functions over finite-length scales,” Applied Thermal Engineering, vol. 50, no. 1, pp. 593–603, 2013
  2. 2. J. Schmitt, G. Zhou, E. Walker, and R. Wall, “Multilayer model of photon diffusion in skin,” JOSA A, vol. 7, no. 11, pp. 2141–2153, 1990
  3. 3. A. Christ, T. Samaras, A. Klingenböck, and N. Kuster, “Characterization of the electromagnetic near-field absorption in layered biological tissue in the frequency range from 30 MHz to 6000 MHz,” Physics in Medicine Biology, vol. 51, no. 19, p. 4951, 2006
  4. 4. N. Chahat, M. Zhadobov, L. Le Coq, S. I. Alekseev, and R. Sauleau, “Characterization of the interactions between a 60-GHz antenna and the human body in an off-body scenario,” IEEE Transactions on Antennas Propagation, vol. 60, no. 12, pp. 5958–5965, 2012
  5. 5. S. Caorsi, M. Pastorino, and M. Raffetto, “Analytic SAR computation in a multilayer elliptic cylinder for bioelectromagnetic applications,” Bioelectromagnetics: Journal of the Bioelectromagnetics Society, The Society for Physical Regulation in Biology Medicine, The European Bioelectromagnetics Association, vol. 20, no. 6, pp. 365–371, 1999
  6. 6. K. S. Nikita, G. S. Stamatakos, N. K. Uzunoglu, and A. Karafotias, “Analysis of the interaction between a layered spherical human head model and a finite-length dipole,” IEEE Transactions on Microwave Theory Techniques, vol. 48, no. 11, pp. 2003–2013, 2000
  7. 7. S. Koulouridis and K. S. Nikita, “Study of the coupling between human head and cellular phone helical antennas,” IEEE Transactions on electromagnetic Compatibility, vol. 46, no. 1, pp. 62-70, 2004
  8. 8. J. Kim and Y. Rahmat-Samii, “Implanted antennas inside a human body: Simulations, designs, and characterizations,” IEEE Transactions on microwave theory techniques, vol. 52, no. 8, pp. 1934-1943, 2004
  9. 9. F. Liu and S. Crozier, “Electromagnetic fields inside a lossy, multilayered spherical head phantom excited by MRI coils: models and methods,” Physics in Medicine Biology, vol. 49, no. 10, p. 1835, 2004
  10. 10. F. Keshmiri and C. Craeye, “Moment-method analysis of normal-to-body antennas using a Green's function approach,” IEEE transactions on antennas propagation, vol. 60, no. 9, pp. 4259–4270, 2012
  11. 11. J. S. Meiguni, M. Kamyab, and A. Hosseinbeig, “Theory and experiment of spherical aperture-coupled antennas,” IEEE Transactions on Antennas and Propagation, vol. 61, no. 5, pp. 2397–2403, 2013
  12. 12. H. Khodabakhshi and A. Cheldavi, “Irradiation of a six-layered spherical model of human head in the near field of a half-wave dipole antenna,” IEEE transactions on microwave theory techniques, vol. 58, no. 3, pp. 680-690, 2010
  13. 13. W. C. Chew, Waves and fields in inhomogeneous media. IEEE press, 1995
  14. 14. J. Sun, C.-F. Wang, J. L.-W. Li, and M.-S. Leong, “Mixed potential spatial domain Green’s functions in fast computational form for cylindrically stratified media,” Progress In Electromagnetics Research, vol. 45, pp. 181–199, 2004
  15. 15. T. Itoh, “Spectral domain immitance approach for dispersion characteristics of generalized printed transmission lines,” IEEE transactions on Microwave Theory and Techniques, vol. 28, no. 7, pp. 733–736, 1980
  16. 16. M. Thiel and A. Dreher, “Dyadic Green's function of multilayer cylindrical closed and sector-structures for waveguide, microstrip-antenna, and network analysis,” IEEE transactions on microwave theory and techniques, vol. 50, no. 11, pp. 2576–2579, 2002
  17. 17. T. V. B. Giang, M. Thiel, and A. Dreher, “A unified approach to the analysis of radial waveguides, dielectric resonators, and microstrip antennas on spherical multilayer structures,” IEEE transactions on microwave theory and techniques, vol. 53, no. 1, pp. 404–409, 2005
  18. 18. L. Novotny and B. Hecht, Principles of nano-optics. Cambridge university press, 2012
  19. 19. C. A. Marocico, X. Zhang, and A. L. Bradley, “A theoretical investigation of the influence of gold nanosphere size on the decay and energy transfer rates and efficiencies of quantum emitters,” The Journal of chemical physics, vol. 144, no. 2, p. 024108, 2016
  20. 20. V. Karanikolas, C. A. Marocico, and A. L. Bradley, “Spontaneous emission and energy transfer rates near a coated metallic cylinder,” Physical Review A, vol. 89, no. 6, p. 063817, 2014
  21. 21. K. Rustomji, R. Abdeddaim, C. M. de Sterke, B. Kuhlmey, and S. Enoch, “Measurement and simulation of the polarization-dependent Purcell factor in a microwave fishnet metamaterial,” Physical Review B, vol. 95, no. 3, p. 035156, 2017
  22. 22. A. Vakil, “Transformation optics using graphene: one-atom-thick optical devices based on graphene,” 2012
  23. 23. C.-T. Tai, Dyadic Green functions in electromagnetic theory. Institute of Electrical & Electronics Engineers (IEEE), 1994
  24. 24. L.-W. Cai, “On the computation of spherical Bessel functions of complex arguments,” Computer Physics Communications, vol. 182, no. 3, pp. 663–668, 2011
  25. 25. D. Bhattacharya, B. Ghosh, P. K. Goswami, and K. Sarabandi, “Evaluation of Efficient Green’s Functions for Spherically Stratified Media,” IEEE Transactions on Antennas and Propagation, vol. 66, no. 3, pp. 1613–1618, 2018
  26. 26. Y. S. Cao, L. J. Jiang, and A. E. Ruehli, “An equivalent circuit model for graphene-based terahertz antenna using the PEEC method,” IEEE Transactions on Antennas and Propagation, vol. 64, no. 4, pp. 1385–1393, 2016
  27. 27. A. Y. Nikitin, F. J. Garcia-Vidal, and L. Martin-Moreno, “Analytical expressions for the electromagnetic dyadic Green’s function in graphene and thin layers,” IEEE Journal of Selected Topics in Quantum Electronics, vol. 19, no. 3, pp. 4600611–4600611, 2012
  28. 28. G. W. Hanson, “Dyadic Green’s functions and guided surface waves for a surface conductivity model of graphene,” Journal of Applied Physics, vol. 103, no. 6, p. 064302, 2008
  29. 29. G. W. Hanson, “Dyadic Green’s functions for an anisotropic, non-local model of biased graphene,” IEEE Transactions on antennas and propagation, vol. 56, no. 3, pp. 747–757, 2008
  30. 30. M. Cuevas, “Surface plasmon enhancement of spontaneous emission in graphene waveguides,” Journal of Optics, vol. 18, no. 10, p. 105003, 2016
  31. 31. T. J. Arruda, R. Bachelard, J. Weiner, and P. W. Courteille, “Tunable Fano resonances in the decay rates of a pointlike emitter near a graphene-coated nanowire,” Physical Review B, vol. 98, no. 24, p. 245419, 2018
  32. 32. E. L. Tan and S. Y. Tan, “A unified representation of the dyadic Green’s functions for planar, cylindrical and spherical multilayered biisotropic media,” Progress In Electromagnetics Research, vol. 20, pp. 75–100, 1998
  33. 33. E. L. Tan and S. Y. Tan, “Coordinate-independent dyadic formulation of the dispersion relation for bianisotropic media,” IEEE Transactions on Antennas Propagation, vol. 47, no. 12, pp. 1820–1824, 1999
  34. 34. E. L. Tan and S. Y. Tan, “Concise spectral formalism in the electromagnetics of bianisotropic media,” Progress In Electromagnetics Research, vol. 25, pp. 309–331, 2000
  35. 35. E. Tan, “Vector wave function expansions of dyadic Green's functions for bianisotropic media,” IEE Proceedings-Microwaves, Antennas propagation, vol. 149, no. 1, pp. 57–63, 2002
  36. 36. W.-B. Lu, H. Chen, and Z.-G. Liu, “A review of microwave devices based on CVD-grown graphene with experimental demonstration,” EPJ Applied Metamaterials, vol. 6, p. 8, 2019
  37. 37. J. Zhang, Z. Liu, W. Lu, H. Chen, B. Wu, and Q. Liu, “A low profile tunable microwave absorber based on graphene sandwich structure and high impedance surface,” International Journal of RF Microwave Computer-Aided Engineering, vol. 30, no. 2, p. e22022, 2020
  38. 38. D. Yi, X.-C. Wei, and Y.-L. Xu, “Tunable microwave absorber based on patterned graphene,” IEEE Transactions on Microwave Theory Techniques, vol. 65, no. 8, pp. 2819–2826, 2017
  39. 39. Z. Fang et al., “Active tunable absorption enhancement with graphene nanodisk arrays,” vol. 14, no. 1, pp. 299–304, 2014
  40. 40. B. Wu et al., “Experimental demonstration of a transparent graphene millimetre wave absorber with 28% fractional bandwidth at 140 GHz,” Scientific reports, vol. 4, p. 4130, 2014
  41. 41. C. Srichan et al., “Highly-sensitive surface-enhanced Raman spectroscopy (SERS)-based chemical sensor using 3D graphene foam decorated with silver nanoparticles as SERS substrate,” Scientific reports, vol. 6, p. 23733, 2016
  42. 42. J. B. Maurya and Y. K. Prajapati, “Experimental Demonstration of DNA Hybridization Using Graphene-Based Plasmonic Sensor Chip,” Journal of Lightwave Technology, 2020
  43. 43. L. A. Falkovsky, “Optical properties of graphene and IV–VI semiconductors,” Physics-Uspekhi, vol. 51, no. 9, p. 887, 2008
  44. 44. H. Nasari and M. Abrishamian, “Nonlinear terahertz frequency conversion via graphene microribbon array,” Nanotechnology, vol. 27, no. 30, p. 305202, 2016
  45. 45. S. Thongrattanasiri, F. H. Koppens, and F. J. G. De Abajo, “Complete optical absorption in periodically patterned graphene,” Physical review letters, vol. 108, no. 4, p. 047401, 2012
  46. 46. S. H. Raad, Z. Atlasbaf, M. Shahabadi, and J. Rashed-Mohassel, “Dyadic Green’s Function for the Tensor Surface Conductivity Boundary Condition,” IEEE Transactions on Magnetics, vol. 55, no. 11, pp. 1–7, 2019
  47. 47. Y. R. Padooru, A. B. Yakovlev, C. S. Kaipa, G. W. Hanson, F. Medina, and F. Mesa, “Dual capacitive-inductive nature of periodic graphene patches: Transmission characteristics at low-terahertz frequencies,” Physical Review B, vol. 87, no. 11, p. 115401, 2013
  48. 48. A. Andryieuski and A. V. Lavrinenko, “Graphene metamaterials based tunable terahertz absorber: effective surface conductivity approach,” Optics express, vol. 21, no. 7, pp. 9144–9155, 2013
  49. 49. Y. Guo, T. Zhang, W.-Y. Yin, and X.-H. Wang, “Improved hybrid FDTD method for studying tunable graphene frequency-selective surfaces (GFSS) for THz-wave applications,” IEEE Transactions on Terahertz Science and Technology, vol. 5, no. 3, pp. 358–367, 2015
  50. 50. J. S. Gomez-Diaz, M. Tymchenko, and A. Alu, “Hyperbolic plasmons and topological transitions over uniaxial metasurfaces,” Physical Review Letters, vol. 114, no. 23, p. 233901, 2015
  51. 51. J. Gomez-Diaz and A. Alu, “Flatland optics with hyperbolic metasurfaces,” ACS Photonics, vol. 3, no. 12, pp. 2211–2224, 2016
  52. 52. A. Fallahi and J. Perruisseau-Carrier, “Design of tunable biperiodic graphene metasurfaces,” Physical Review B, vol. 86, no. 19, p. 195408, 2012
  53. 53. D. L. Sounas and C. Caloz, “Gyrotropy and nonreciprocity of graphene for microwave applications,” IEEE Transactions on Microwave Theory and Techniques, vol. 60, no. 4, pp. 901–914, 2012
  54. 54. L. Lin, L.-S. Wu, W.-Y. Yin, and J.-F. Mao, “Modeling of magnetically biased graphene patch frequency selective surface (FSS),” in 2015 IEEE MTT-S International Microwave Workshop Series on Advanced Materials and Processes for RF and THz Applications (IMWS-AMP), 2015, pp. 1–3: IEEE
  55. 55. Y. Yermakov et al., “Effective surface conductivity of optical hyperbolic metasurfaces: from far-field characterization to surface wave analysis,” Scientific reports, vol. 8, no. 1, p. 14135, 2018
  56. 56. J. Komijani, J. Rashed-Mohassel, and A. Mirkamali, “Dyadic Green Functions for a Dielectric Layer on a PEMC Plane,” Progress In Electromagnetics Research, vol. 6, pp. 9–22, 2009
  57. 57. J. Gómez-Díaz and J. Perruisseau-Carrier, “Propagation of hybrid transverse magnetic-transverse electric plasmons on magnetically biased graphene sheets,” Journal of Applied Physics, vol. 112, no. 12, p. 124906, 2012
  58. 58. Raad, Shiva Hayati, Zahra Atlasbaf, and Mauro Cuevas. “Dyadic Green's function for the graphene-dielectric stack with arbitrary field and source points.” arXiv preprint arXiv:2104.03581 (2021)
  59. 59. M. Cuevas, M. A. Riso, and R. A. Depine, “Complex frequencies and field distributions of localized surface plasmon modes in graphene-coated subwavelength wires,” Journal of Quantitative Spectroscopy and Radiative Transfer, vol. 173, pp. 26–33, 2016
  60. 60. M. Hajati and Y. Hajati, “Plasmonic characteristics of two vertically coupled graphene-coated nanowires integrated with substrate,” Applied Optics, vol. 56, no. 4, pp. 870–875, 2017
  61. 61. S. H. Raad and Z. Atlasbaf, “Dyadic analysis of a cylindrical wire consisting of a cover with fully-populated surface conductivity tensor,” Optics express, vol. 27, no. 15, pp. 21214–21225, 2019
  62. 62. J. Gomez-Diaz, M. Tymchenko, and A. Alù, “Hyperbolic metasurfaces: surface plasmons, light-matter interactions, and physical implementation using graphene strips,” Optical Materials Express, vol. 5, no. 10, pp. 2313–2329, 2015
  63. 63. K.-H. Kim, Y.-S. No, S. Chang, J.-H. Choi, and H.-G. Park, “Invisible hyperbolic metamaterial nanotube at visible frequency,” Scientific reports, vol. 5, p. 16027, 2015
  64. 64. Y. Mazor and A. Alù, “Angular-momentum selectivity and asymmetry in highly confined wave propagation along sheath-helical metasurface tubes,” Physical Review B, vol. 99, no. 15, p. 155425, 2019
  65. 65. Y. Huang, Y. Shen, C. Min, and G. Veronis, “Switching photonic nanostructures between cloaking and superscattering regimes using phase-change materials,” Optical Materials Express, vol. 8, no. 6, pp. 1672–1685, 2018
  66. 66. Y. Hua and T. K. Sarkar, “Generalized pencil-of-function method for extracting poles of an EM system from its transient response,” 1989
  67. 67. C. Tokgoz and G. Dural, “Closed-form Green's functions for cylindrically stratified media,” IEEE Transactions on Microwave theory and Techniques, vol. 48, no. 1, pp. 40–49, 2000
  68. 68. S. H. Raad, C. J. Zapata-Rodríguez, and Z. Atlasbaf, “Multi-frequency super-scattering from sub-wavelength graphene-coated nanotubes,” JOSA B, vol. 36, no. 8, pp. 2292–2298, 2019
  69. 69. S. H. Raad, C. J. Zapata-Rodríguez, and Z. Atlasbaf, “Graphene-coated resonators with frequency-selective super-scattering and super-cloaking,” Journal of Physics D: Applied Physics, vol. 52, no. 49, p. 495101, 2019
  70. 70. S. H. Raad, Z. Atlasbaf, and M. Cuevas, “Scattering from Multilayered Graphene-Based Cylindrical and Spherical Particles,” in Nanoplasmonics: IntechOpen, 2020
  71. 71. X. Ma et al., “Graphene oxide wrapped gold nanoparticles for intracellular Raman imaging and drug delivery,” vol. 1, no. 47, pp. 6495–6500, 2013
  72. 72. D. Cai, L. Ding, S. Wang, Z. Li, M. Zhu, and H. Wang, “Facile synthesis of ultrathin-shell graphene hollow spheres for high-performance lithium-ion batteries,” Electrochimica Acta, vol. 139, pp. 96-103, 2014
  73. 73. S. H. Raad, Z. Atlasbaf, C. J. Zapata-Rodríguez, M. Shahabadi, and J. Rashed-Mohassel, “Dyadic Green's function for the electrically biased graphene-based multilayered spherical structures,” Journal of Quantitative Spectroscopy Radiative Transfer, vol. 256, p. 107251, 2020
  74. 74. Z.-S. Wu, Z.-J. Li, H. Li, Q.-K. Yuan, and H.-Y. Li, “Off-axis Gaussian beam scattering by an anisotropic coated sphere,” IEEE Transactions on Antennas and Propagation, vol. 59, no. 12, pp. 4740–4748, 2011
  75. 75. L.-W. Li, P.-S. Kooi, M.-S. Leong, and T.-S. Yee, “Electromagnetic dyadic Green's function in spherically multilayered media,” IEEE Transactions on Microwave Theory and Techniques, vol. 42, no. 12, pp. 2302–2310, 1994
  76. 76. M. Farhat, C. Rockstuhl, and H. Bağcı, “A 3D tunable and multi-frequency graphene plasmonic cloak,” Optics express, vol. 21, no. 10, pp. 12592–12603, 2013
  77. 77. F. TING, “Fast solutions of electromagnetic fields in layered media,” 2008
  78. 78. J. D. Shumpert, “Modeling of Periodic Dielectric Structures (electomagnetic Crystals),” University of Michigan, 2001
  79. 79. M. Wan et al., “Strong tunable absorption enhancement in graphene using dielectric-metal core-shell resonators,” Scientific reports, vol. 7, no. 1, p. 32, 2017
  80. 80. S. H. Raad, Z. Atlasbaf, J. Rashed-Mohassel, and M. Shahabadi, “Scattering from graphene-based multilayered spherical structures,” IEEE Transactions on Nanotechnology, vol. 18, pp. 1129–1136, 2019
  81. 81. S. H. Raad and Z. Atlasbaf, “Equivalent RLC ladder circuit for scattering by graphene-coated nanospheres,” IEEE Transactions on Nanotechnology, vol. 18, pp. 212–219, 2019
  82. 82. S. H. Raad and Z. Atlasbaf, “Tunable optical meta-surface using graphene-coated spherical nanoparticles,” AIP Advances, vol. 9, no. 7, p. 075224, 2019
  83. 83. T. J. Arruda, A. S. Martinez, F. A. Pinheiro, R. Bachelard, S. Slama, and P. W. Courteille, “Fano resonances in plasmonic core-shell particles and the Purcell effect,” in Fano Resonances in Optics and Microwaves: Springer, 2018, pp. 445–472
  84. 84. V. D. Karanikolas and A. L. Bradley, “Interactions Between Quantum Emitters in the Presence of Plasmonic Nanostructures,” Trinity College Dublin, 2016

Written By

Shiva Hayati Raad and Zahra Atlasbaf

Submitted: 04 December 2020 Reviewed: 05 January 2021 Published: 02 December 2021