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

Graphene-based near-field optical microscopy: high-resolution imaging using reconfigurable gratings

Open Access Open Access

Abstract

High-resolution and fast-paced optical microscopy is a requirement for current trends in biotechnology and materials industry. The most reliable and adaptable technique so far to obtain higher resolution than conventional microscopy is near-field scanning optical microscopy (NSOM), which suffers from a slow-paced nature. Stemming from the principles of diffraction imaging, we present fast-paced graphene-based scanning-free wide-field optical microscopy that provides image resolution that competes with NSOM. Instead of spatial scanning of a sharp tip, we utilize the active reconfigurable nature of graphene’s surface conductivity to vary the diffraction properties of a planar digitized atomically thin graphene sheet placed in the near field of an object. Scattered light through various realizations of gratings is collected at the far-field distance and postprocessed using a transmission function of surface gratings developed on the principles of rigorous coupled wave analysis. We demonstrate image resolutions of the order of λ0/16 using computational measurements through binary graphene gratings and numerical postprocessing. We also present an optimization scheme based on the genetic algorithm to predesign the unit cell structure of the gratings to minimize the complexity of postprocessing methods. We present and compare the imaging performance and noise tolerance of both grating types. While the results presented in this article are at terahertz frequencies (λ0=10μm), where graphene is highly plasmonic, the proposed microscopy principle can be readily extended to any frequency regime subject to the availability of tunable materials.

© 2017 Optical Society of America

1. INTRODUCTION

Imaging devices with resolutions beyond diffraction limit have always been challenging to realize but are highly anticipated in the fields of biomedical optics, material science, and nanofabrication [14]. Information regarding subwavelength features of an object usually resides in the evanescent waves that decay in the vicinity of the object [5]. Conventional microscopy based on ray optics and dielectric lenses fails to efficiently refract the evanescent waves. Consequently, the information about subwavelength features is discarded, limiting the ideal spatial resolution to λ0/2. So far, the most reliable technique for high-resolution imaging is near-field scanning optical microscopy [6] (NSOM). NSOM relies on collecting the evanescent information before it decays in the near field of an object. It operates by collecting a point-by-point (pixel-by-pixel) image of the object using a sharp metallic tip of subwavelength dimensions placed in the near field of the object. The image resolution does not have any fundamental limit but depends on the radius of the tip, its distance from the object, and experimental noise. The technique has been successful in providing resolutions of the order of tens of nanometers. However, it is relatively slow paced due to the requirement of mechanical movements with subwavelength steps, making it unsuitable for real-time scenarios.

The requirement of high-resolution imaging devices and the hurdle of scanning is not only limited to visible frequencies. There are some biomedical scenarios where cancers, such as colon cancer [4,7,8] and skin cancer [9], are observed to be detectable only through terahertz radiation. Improving image resolutions at such frequencies is also of immense importance [10,11].

On the other hand, two-dimensional materials, such as grapheme [12,13], transition metal dichalcogenides (TMDCs) [1416], and boron nitrides [17], have been shown to possess metallic properties in the terahertz frequency regime with extensive ability to tune the resonating responses over a broad range of frequencies. The ability to engineer the electronic and optical properties of materials actively using external stimuli has boosted many applications in the field of electro-optics [16,18]. In addition, extreme effort has been made to push the plasmonic resonance of two-dimensional heterostructures to the visible frequency regime [19,20] and also to induce gate-bias tunability to optical properties of existing plasmonic materials at visible and near-IR frequencies [21,22].

In this paper, utilizing the tunable resonance property of graphene, we present the alternate wide-field microscopy technique of graphene-based near-field optical microscopy (GNOM), which is free of mechanical movements based on the principles of diffraction-based microscopy [2333]. In contrast to a moving tip, as in NSOM or other scanning-based imaging devices, we dynamically vary the optical properties (spatial surface conductivity profile) of a planar digitized graphene sheet placed in the near field of an object whose subwavelength features need to be resolved. The scattered light is collected at a far-field distance using conventional microscopy methods. However, as is the case of most diffraction-based imaging techniques, the measurements do not represent a final image of the object and computational postprocessing is required to obtain a high-resolution image. The scanning of optical properties of the sheet, hence the measurements, can be performed dramatically faster in time by varying the electrostatic bias using programmable electronics. Using computational experiments, we show that image resolution of the order of λ0/16 (comparable to NSOM) can be achieved using GNOM. Even though we use properties of a graphene sheet at terahertz frequencies for proof of principle studies, the technique can be adapted to any tunable material at any frequency range.

Diffraction-based imaging techniques, such as structured illumination microscopy (SIM) [24], optical diffraction tomography (ODT) [25,27,33], scanning far-field superlens (FSL) [23,26] and interscale mixing microscopy (IMM) [2830], have come into focus recently, where conventional wide-field microscopy measurements are postprocessed with numerical algorithms to resolve subwavelength features of an object beyond the diffraction limit. SIM relies on Moiré fringes where objects are illuminated with patterned light formed by interference of beams from multiple angles. More images are collected by shifting the fringe pattern, and the data are postprocessed in the Fourier domain to deconvolute the object data from the pattern data. The maximum resolution is limited to λ0/4, corresponding to the periodicity of the fringe pattern. ODT relies on measurements of amplitude and phase of the scattered field from the object by illuminating from multiple directions, where the postprocessing aims to retrieve the refractive index profile of the object, assuming it as a scattering potential. The maximum resolution is also limited to λ0/4 [34]. Even though both techniques are fast paced in comparison to NSOM in terms of measurements, the resolution in these techniques is fundamentally limited to the order of λ0/4 or less since both cases do not capture the information of the evanescent spectrum decaying in the near field.

The latter two techniques (FSL and IMM) utilize a diffraction grating in the near field of the object to transfer the evanescent information fading in the vicinity of the object to the far field. From the principles of diffraction, a diffraction grating reflects and transmits an incident plane wave or evanescent wave with tangential spatial frequency component kx0 into a number of diffraction orders, each characterized by a wave vector component defined by the Bragg-periodic condition

kxm=kx0+m2πΛ,
where m is an integer ranging from to and Λ is the periodicity of the grating. A proper design of the grating (usually with subwavelength periodicity) couples the high-spatial-frequency waves to low-spatial-frequency waves, which fall in the propagation regime and thereby get transferred to the far fields. The measured far-field intensity comprises of information folded from many diffraction orders (of higher spatial frequency). Numerical postprocessing is then applied to unfold the corresponding amplitude spectrum.

FSL relies on the design of the grating where evanescent waves of the object around the surface plasmon polariton (SPP) resonance wave vector [23] of the grating are folded to the propagation regime using only m=±1 diffraction orders. This simplifies the postprocessing of the data; however, the resolutions are limited to the surface plasmon resonance of the grating material (λ0/4).

On the other hand, IMM is designed with no such (resonance) restrictions and includes multiple diffraction orders where the grating can be designed either with metallic or dielectric material [28]. Far-field intensity measurements from multiple incident angles are considered to numerically unfold the evanescent spectrum of the object from multiple diffraction orders. Recently, image resolution of the order of λ0/10 has been demonstrated experimentally using IMM combined with a postprocessing method based on scalar diffraction theory [29]. The resolution limit of the IMM technique does not have any fundamental barrier apart from its dependence on the distance between the diffraction element and the object, and experimental noise. IMM is fast paced in terms of measurement in comparison to pixel-by-pixel scanning. However, it is relatively slow due to the requirement of multiple incident angular measurements to accurately unfold the evanescent spectrum. The multiple incident angle measurements with the rotation of source or object are typically prone to alignment errors [29,35]. Also, the near-field interaction is limited to one diffraction element impelling the limited number of multiple incident angle measurements the responsibility to transfer nonredundant near-field information of the object to the far field. In addition, the resolution is also hampered by the material loss of the diffraction gratings.

GNOM is free from the above imperfections with the ability to actively vary the diffraction properties of graphene without disturbing the system [3639]. Optical properties of graphene are usually described by the real and imaginary parts of its surface conductivity, whose magnitudes are determined by the doping concentration of electrons injected using an electrostatic bias [4042]. In the terahertz regime, the conductivity usually follows the Drude model and is linearly proportional to the square root of the surface doping concentration (ns). For a given doping density, the conductivity of a graphene sheet at terahertz wavelengths can be expressed as [43]

σ=e2Efε0cπ211τiω,
where Ef=vfπns represents the Fermi energy, vf represents the Fermi velocity, and τ represents the relaxation time. The above definition of surface conductivity is unitless (due to the normalization with ε0C) [41]. The doping concentration ns can be varied by controlling the voltage of the gate bias. In practice, Ef can be varied in the range of 0–2 eV [44,45].

A schematic of the setup of GNOM is shown in Fig. 1(a). A planar digitized graphene sheet is attached to a substrate and is placed on top (in close proximity) of an object that needs to be imaged with the graphene side down. The digitization of the sheet can be performed by considering graphene strips each of width d placed side-by-side on a substrate (in the xy-plane) with each strip attached to electrodes along the longitudinal direction (y axis). By controlling the voltage of each strip, various patterns of the surface conductivity profile with a step size of d can be programmed digitally [46]. For example, a surface grating of period 6d and duty cycle 3d can be created by periodically turning on a constant voltage for three consecutive strips and by turning off the next three strips. From the current fabrication technologies, strip width d can be made as small as 50 nm [47]. For terahertz wavelengths (λ010μm), the strip width is around λ0/200, allowing an almost smooth variation of the profile with respect to the wavelength.

 figure: Fig. 1.

Fig. 1. (a) Schematic setup of GNOM. Objects are placed on a substrate (ε=2.25). Graphene sheet slices with electrodes are placed at a distance h on top of the object. Red and black colors of the graphene slices represent different values of surface conductivity. The inset is the schematic for transmission function calculation. (b), (c) The relative variation in field amplitude at angles of 0° and 30°, respectively, for five different objects numbered 1–5 shown as insets. Object 1total size(5λ0), Object 2total size(3λ0), Object 3total size(2λ0), Object 4total size(2λ0) with subwavelength features (λ0/8, λ0/4, λ0/2, λ0/8, λ0/4), and Object 5total size(2λ0) with subwavelength feature sizes (λ0/16, λ0/8, λ0/4, λ0/2, λ0/16, λ0/4). The gaps between the subwavelength features are of relative size. For objects with the smallest subwavelength features (more evanescent spectrum) the variation in relative amplitude is high and irregular (nonredundant) in pattern. |H| and |H0| represent the amplitude of the field with and without a grating, respectively.

Download Full Size | PDF

Light emitted from the object in the form of both propagation and evanescent waves are allowed to interact with various engineered conductivity profiles discussed in later sections. Each profile couples (“mixes”) a part of the evanescent spectrum to the propagation regime that eventually gets transmitted to the far fields. The transmitted light through each profile is measured at a far-field distance through the substrate of graphene. Since the interaction of each evanescent wave with each conductivity profile is unique, each transmitted signal inherently contains nonredundant information about the object’s evanescent spectrum.

To demonstrate the uniqueness in the interaction of evanescent waves with each conductivity profile, we calculated (using COMSOL, RF Module) [48] the relative variation in the amplitude of the transmitted signal with and without the conductivity profile for a set of objects comprised of different subwavelength features. The object that we consider here is a transparent light-emitting source with an overall lateral dimension of wavelength order but is chopped into subwavelength pieces and gaps, as shown in the inset of Figs. 1(b) and 1(c) (numbers 4 and 5). An efficient imaging system is expected to recover the dimensions and locations of both pieces and gaps. In COMSOL, such sources can be defined as surface currents of required lengths by using the current density boundary condition. The surface currents, designed to radiate transverse-magnetic- (TM) polarized light, are then placed below a surface grating of unit cell length Λ made of two components of equal length ratios. The conductivities of the two components are defined as σ1=0.0025+0.2387i(Ef=1eV) and σ2=104(Ef=0eV), respectively (a small real part for σ2 is assumed to represent interband transitions at zero doping concentration [4956]). The light emitted from the sources and transmitted through the grating is first calculated in the near-field zone along a planar boundary in the transmitted region and then a Fourier transform is applied to compute far-field strengths. The relative variation in the far-field amplitudes with and without the grating as a function of periodicity of the grating at 0° and 30° transmitted angles are presented in Figs. 1(b) and 1(c), respectively. Since objects with smaller feature sizes contribute more to the evanescent spectrum, the magnitude of the relative variation of the far field for objects with no subwavelength features (numbers 1–3 in the insets) is less when compared to objects with subwavelength features (numbers 4 and 5 in the insets). Out of the three objects with the same overall size (numbers 3–5), the one with the smallest subwavelength features (number 5) gets maximum variation. The irregularity in the pattern as a function of periodicity represents the uniqueness in the interaction of each grating with each object. The similarity in the relative variation of the fields at both angles [in Figs. 1(b) and 1(c)] demonstrates that the object–grating interaction weakly depends on the measuring direction. In later sections, by defining a suitable transfer function for each grating, we demonstrate recovery of the locations and sizes of pieces and gaps of an object from such far-field measurements.

Note that even though the transparent nature of the utilized virtual object is an ideal assumption, diffraction-based imaging techniques such as IMM started with similar theoretical assumptions have been successful in experimental demonstrations [2830].

2. OPTICAL TRANSFER FUNCTION OF GRAPHENE GRATINGS

The image reconstruction process is equivalent to calculating the electromagnetic field of the object at the object plane. The Fourier spectrum of the magnetic field of the object in the object plane can be expressed as

Hobj(x)=a+(kx)eikxxdkx=wl=al+(kxl)eikxlx,
where w represents the weighting factor. Here, kx spans the whole propagation and evanescent spectrum and the problem of image reconstruction is now reduced to calculating the amplitude spectrum (al+) of waves originating from the object.

Postprocessing of intensity measurements using the GNOM technique requires calculation of the transmission coefficient (or coupling coefficient) of each incident plane wave or evanescent wave, represented by kx, into its diffraction orders. In order to do so, consider a graphene sheet with a one-dimensional surface conductivity profile given by σ(x) between two media (region I and region II) with permittivity εI and εII, respectively, as shown in the inset of Fig. 1(a). Assuming TM-polarized waves with E={Ex,0,Ez} and H={0,Hy,0}, solving Maxwell’s equations in Gaussian units and Cartesian coordinates, the tangential electric and magnetic field components can be expressed as

Hy{I,II}(x,z)=m=[am+{I,II}eikzm{I,II}z+am{I,II}eikzm{I,II}z]eikxmx,Ex{I,II}(x,z)=m=kzm{I,II}k0ε{I,II}[am+{I,II}eikzm{I,II}zam{I,II}eikzm{I,II}z]eikxmx,Ez{I,II}(x,z)=m=kxmk0ε{I,II}[am+{I,II}eikzm{I,II}z+am{I,II}eikzm{I,II}z]eikxmx.
am±{I,II} represent amplitudes of waves, where (+) represents propagation in forward (+z) direction and () represents propagation in backward (z) direction. Superscripts {I, II} denote regions I and II, respectively. kzm{I,II} are wave vector components of waves along the z axis in regions I and II, calculated from the dispersion relation of each medium as
kzm{I,II}=k02ε{I,II}kxm2,
where kxm, defined by Eq. (1), represents the tangential wave vector component of each diffraction order of the system and is conserved across the interface in regions I and II. The tangential boundary conditions that need to be satisfied at each point along the graphene grating interface (at z=0) can be expressed as
ExI(x,z=0)ExII(x,z=0)=0,HyI(x,z=0)HyII(x,z=0)=σ(x)ExII(x,z=0).
The Fourier series of the function σ(x) can be expressed as σ(x)=m=σmeiqmx; qm=m2πΛ, where σm represent the Fourier coefficients. Applying the Fourier transform to Eq. (6) with eikxnxdx and by using the Fourier convolution theorem on the product σ(x)ExII(x,z=0), the boundary conditions in Eq. (6) can be transformed into linear matrix equations as
1εIKzI(AI+AI)1εIIKzII(AII+AII)=0,(AI+AI)(AII+AII)=1k0εIIΞKzII(AII+AII).
Here, A{I,II}± are column vectors with amplitude coefficients whose elements are given as Am,1±{I,II}=am±{I,II}, Kz{I,II} represent diagonal matrices with diagonal elements defined as Kz;m,m{I,II}=kz;m{I,II}, and Ξ is a Toeplitz matrix of Fourier components of the conductivity profile defined as Ξm,n=σmn. The above matrix equations can be rewritten into a scattering matrix form relating the amplitudes of waves scattered from the interface to the amplitudes of incident waves on the interface as
[AIAII+]=[R+TT+R][AI+AII],
where
R+=[k0(εIIKzI+εIKzII)+KzIIΞKzI]1[k0(εIIKzIεIKzII)+KzIIΞKzI],T=[k0(εIIKzI+εIKzII)+KzIIΞKzI]1[2k0εIKzII],T+=[k0(εIIKzI+εIKzII)+KzIΞKzII]1[2k0εIIKzI],R=[k0(εIIKzI+εIKzII)+KzIΞKzII]1[k0(εIIKzIεIKzII)+KzIΞKzII].
For a simple case with one grating layer and waves incident from one direction, i.e., AII=0, the amplitudes of the reflected and transmitted diffraction orders can be calculated as a function of amplitudes of incident waves as AI=R+AI+ and AII=T+AI+, respectively. If a system has more than one grating layer, the reflection and transmission matrices (R± and T±) can be computed for each layer using Eq. (8) and “multiplied” carefully including the propagation phase factors between the layers to calculate the amplitudes of forward and backward propagating waves in each layer, as shown in [41,57]. The total electric and magnetic fields at any spatial coordinate can be calculated using Eq. (4).

The matrix T+ represents the coupling coefficients (or “mixing” coefficients) of incident plane or evanescent waves with wave vectors represented by kxm to transmitted plane or evanescent waves with the same kxm, which can serve as the optical transfer function. However, note that the wave vector spectrum kxm is discrete with spacing between them defined by the periodicity of the grating kxm+1kxm=2π/Λ. For gratings with periodicities less than or of the order of the wavelength, the discretized kxm spectrum does not represent the object spectrum accurately. Instead, one can define a super cell structure with a large but finite number of repeating unit cells, where Λ now represents the periodicity of the super cell. Since Λ, the kxm set is dense enough to represent the object spectrum. In such case, the matrix T+ itself represents the transfer function (T) of the grating. However, the numerical decomposition in Eq. (3) requires a large number of slices for accuracy making it impractical to invert the matrices [in Eq. (9)] of such high order.

Alternately, the matrix T+ can be computed for each kx0 [in Eq. (1)] spanning from π/Λ to π/Λ, (kx0kx0;l[π/Λ,π/Λ)) to fill the gaps in the kxm spectrum. The optical transfer function T can be obtained by dividing it into number of block matrices as

T=[Tm1,n1Tm1,nTm1,n+1Tm,n1Tm,nTm,n+1Tm+1,n1Tm1,nTm+1,n+1],
where each block is a diagonal matrix whose elements are defined using the transmission matrices T+(l) computed for each kx0;l as (Tm,n)ll=Tmn+(l). In comparison to the inversion of large-order matrices in the former method, here we invert a large number of small-order matrices, increasing the accuracy and decreasing the computational time.

Also, it has been noted that due to the formation of grain boundaries, the surface conductivity of graphene and other two-dimensional materials are not generally uniformly controllable over a wide spatial range [58]. Note that the above equations are also valid in such cases when σ(x) is not periodic. However, precalibration of the graphene sheets to identify and calculate the possible surface conductivity profiles is needed. For such nonuniform cases, the inversion of large matrices is inevitable. In addition, the above formulation is also applicable for thin gratings or metasurfaces (of thickness d0λ0) made of bulk materials. Equation (9) can be computed by converting the bulk permittivity pattern of the materials (εr(x)) to surface conductivity pattern (σ(x)) using σ(x)=i(εr(x)1)k0d0 [59].

3. IMAGE RECONSTRUCTION WITH BINARY GRAPHENE GRATINGS

The transfer of evanescent waves to the propagation regime is usually accompanied by the distribution of light into multiple diffraction orders that eventually decay in the near field of the grating. In order to get more information from the object, here we utilize far-field measurements taken by varying the periodicity of the graphene grating. As discussed before, in practice such variation can be programmed using electronics for graphene without disturbing the system. It has been shown in Figs. 1(b) and 1(c) that different subwavelength features of the object interact differently with each grating and consequently transfer nonredundant information regarding the subwavelength features of the object to the far field. Here we present the numerical image recovery procedure for GNOM utilizing only one far-field measurement in the normal direction through the graphene gratings as a function of its periodicity.

In order to evaluate the performance of GNOM in reconstructing subwavelength features of an object, we performed computational experiments with calculations from COMSOL, considered as “measurements.” A test object is designed with a total length of 2λ0 composed of subwavelength features of dimensions ranging from λ0/16 to λ0/2. The field pattern of the object is shown as the dashed line in Fig. 2(c). The object is placed in the near field of a grating at a distance of 100 nm radiating transverse-magnetic-polarized waves that pass through the grating, as shown in the schematic in Fig. 1(a). Far-field “measurements” are taken using the far-field node in COMSOL by defining the periodicity of the grating as Λ=λ0/f, where f varies from 1 to 16 in 300 steps. The overall grating size is fixed to 10λ0 in all cases. The diffraction gratings are composed of binary units with Ef=0eV for one of the components. To compensate for the loss of phase information, intensity measurements are taken at two Fermi energies of graphene Ef=1eV and Ef=0.9eV for the other component. The intensity “measurements” in the normal direction of the grating as a function of periodicity for the test object are shown in Fig. 2(a). Note that Ef=0eV may represent zero conductivity according to Eq. (2). However, a small nonzero real part (σ=104) is assumed here to represent the interband conductivity of graphene when the doping concentration is zero [4956].

 figure: Fig. 2.

Fig. 2. Image reconstruction using binary graphene gratings. (a) Transmitted far-field intensity “measurements” of the test object (calculated using COMSOL) in the normal direction with grating periodicity reconfigured from Λ=λ0 to Λ=λ0/16 in 300 steps, λ0=10μm. The unit cell has two equal size components with σ1Ef and σ2=104. The object-to-grating distance is h=100nm. (b) The corresponding transmission function matrix (TG) with rows as the rows of the transmission function of each grating in the normal direction calculated using Eq. (10), where color represents the magnitude of the matrix element. Upper and lower sections represent gratings with the two Fermi levels of graphene. (c) The test object (dotted line) and the reconstructed image (solid line) using Eq. (13). The overall size of the object is 2λ0, and the subwavelength feature sizes are (λ0/16, λ0/8, λ0/4, λ0/2, λ0/4) with gaps of relative size between them. The reconstructed image resolves the gaps and features of the object.

Download Full Size | PDF

To begin the process of image reconstruction from the far-field measurements, the transfer function T (=T(g)) is calculated for a grating with each periodicity using Eq. (10), where g is an index to represent each grating. Since only one far-field measurement per grating is considered in recovery, only one row (central row for normal direction) of each T(g) is required in postprocessing. A composite matrix (TG) is then defined whose rows are filled with a central row of each T(g), i.e., TgmG=T0m(g). The composite transmission matrix TG relates the amplitude spectrum of the object to transmitted far fields by all gratings simultaneously. A section of the elements of matrix TG is shown in Fig. 2(b). The number of nonzero elements in each row indicates the number of diffraction orders coupled into the propagation regime and the magnitude shows the strength of the coupling.

The amplitude spectrum of the object is expanded in the pixel basis, as described in [28], where the object plane is divided into a number of pixels. The problem of image reconstruction is now reduced to the problem of finding the field amplitudes of each pixel (hp) centered at a coordinate xp with width P, where P=λ0/16 in this example. Using the fact that the Fourier transform of each pixel is a sinc function, the transfer function (TF) relating the amplitude of the field (hp) at each pixel (xp) to the transmitted far-field amplitude (ag) by each grating can be defined as

TgpF=m=TgmGexp(ikxmxp)sin(kxmP/2)/(kxmP/2).
The calculated far-field transmitted intensity by each grating can then be expressed using the above equation as
IgC=|ag|2=|pTgpFhp|2.
Each pixel amplitude hp is then obtained using nonlinear numerical optimization methods [60,61] to minimize the deviation between “measured” far-field intensity [Imeas in Fig. 2(a)] and the defined far-field intensity [IgC in Eq. (13)]. We have
|IgCImeas|2min.
The recovered value of amplitude of each pixel (hp) of the test object using the above-described procedure is presented in Fig. 2(c) (solid line). The image clearly demonstrates the smallest features of the order of λ0/16 of the object as expected. It has to be noted that features of the order of λ0/16 are successfully resolved both when nearly isolated (toward the edges of the object) and when placed nearby a strongly scattering object (of width λ0/2). The relatively low magnitude artifacts in the recovered image can be attributed to (i) the discrepancies in the two numerical techniques used (COMSOL for measurements RCWA for recovery) and (ii) the edge effects of the finite size gratings used in COMSOL.

4. IMAGE RECONSTRUCTION WITH OPTIMIZED GRAPHENE GRATINGS

The Imaging performance of GNOM relies on (i) efficiency of the diffraction element to couple (“mix”) the evanescent waves to the propagation regime and (ii) the accuracy of the image reconstruction mechanism to uncouple (“unmix”) the contributions of evanescent and propagation parts originated from the object. However, mixing too many diffraction orders into one far-field measurement complicates the unmixing algorithm, leading to the parasitic artifacts seen in Fig. 2(c). In order to keep the unmixing algorithm accurate and efficient, the diffraction element itself can be designed to share the unmixing process. Using the genetic algorithm [6264], here we designed a number of optimized unit cell patterns with a fixed periodicity Λ, where each pattern enhances the coupling of a specific part of the evanescent spectrum of the object to the propagation regime suppressing the others. The design aim of each unit cell is to maximize the coupling of an evanescent wave with wave vector kxa to its corresponding mth diffraction order in the propagation regime kxb=kxa+m2π/Λ (|kxb|<k0) suppressing others. The design principle is also close to the concept of a metasurface [6569], where an incident plane wave with wave vector kxa is coupled to a plane wave with wave vector kxb at an interface with periodic phase gradient dϕ/dx=kxakxb. However, instead of forcing the phase gradient by engineering each unit of the unit cell, we utilized the genetic algorithm to optimize the unit cell pattern.

The periodicity of the unit cell Λ is fixed equal to the free-space wavelength (Λ=λ0) and then made into 32 pieces of equal length. The free parameter of each unit of the unit cell is only the Fermi level of the graphene Ef, which relates to the surface conductivity by Eq. (2). The magnitude of Ef is limited to be in the range of 02eV [44,45], which corresponds to the feasible range in practice. It can be verified that for gratings with periodicity of the unit cell Λ=λ0, for any chosen kxa there always exists an integer m such that it is the mth and m+1th diffraction orders fall in the propagation regime. To limit the measurement cone to less than 30°, we consider only one out of two diffraction orders that satisfies the condition kxbk0/2 as the output direction. The objective function of the optimization process is defined using the transmission function (T+) in Eq. (9). The far-field transmission into the wave with wave vector kxb is given by one row of the matrix represented as Tbn+ (the subscript b stands for wave vector kxb). To maximize the coupling of the wave with wave vector kxa to the wave with wave vector kxb and suppress others, it is enough to maximize the element Tba+ in the row Tbn+ while suppressing others. Hence, the objective function for optimization can be defined as

n|Tbn+δan|2min.
The resulting unit cell patterns of the diffraction elements from the optimization process are presented in Figs. 3(a) and 3(b). The unit cell patterns are mostly nontrivial to describe; however, it can be observed that Ef(kxa,x)=Ef(kxa,x) and Ef(kxa,x)=0 for kxa<k0/2. The former equation represents the mirror symmetry of the unit cells when coupling waves with kxa<0 and kxa>0. The latter equation represents that when kxa itself is on the propagation regime, no grating is required to maximize the coupling to itself. The nearly smooth unit cell profiles around kxa/k0=10 correspond to the strong surface plasmon resonance region of the graphene sheets [41]. Note that the optimized unit cell patterns are not unique solutions but are the best possible solutions within the limitations of geometrical and optical properties of the graphene sheets. In a realistic setup, the limitations need to be determined from a set of calibration measurements of the surface conductivity to include the effects of all the defects and grain boundaries of the fabricated two-dimensional material. At the implementation level, where defects in the fabrication process are inevitable, optimized grating profiles could also be designed experimentally utilizing a set of known test objects.

 figure: Fig. 3.

Fig. 3. Image reconstruction using gratings with optimized unit cells. (a) Fermi level, consequently the surface conductivity, profile of unit cells optimized to maximize the coupling of each wave vector (kxa) to the corresponding diffraction order in the propagation regime (Λ=λ0, λ0=10μm). Each unit cell has 32 elements. A color bar is displayed for clarity. (b) Line plots of the same quantity for three representative wave vectors. (c) The elements of composite transmission matrix TG, where each row is the row of the transmission function calculated using Eq. (10) for corresponding transmission direction of each optimized grating. (d) The real part of the magnetic field of the object (black dashed line) from COMSOL and the corresponding image (solid and dotted lines) recovered analytically using Eqs. (3) and (15) with no numerical postprocessing. The inset shows the corresponding amplitude spectrum of the object and image. The electric and magnetic fields of the image are computed using Eq. (4) as H=Hz and E=Ex2+Ez2.

Download Full Size | PDF

Even though the optimization maximizes the element Tba+ by suppressing others, the direct coupling of propagation wave to propagation wave Tbb+ is inevitable due to limited range of Ef and fixed geometrical parameters. To perform the imaging recovery process, the composite transmission function of all gratings TG can be defined as a sparse matrix, where each row is filled with elements of the transmission function of each optimized grating carefully aligned to the corresponding kxm values. The elements of the matrix TG are presented in Fig. 3(c). It can be observed that in comparison to the matrix of regular periodic gratings in Fig. 2(b), the number of dominant elements is minimized per each row. The dominant elements along the diagonal represent the maximized element of each row and the transmission band around the central column represents the direct coupling of propagation to propagation orders. It is evident that the transmission measurements taken through each of these gratings consist of dominant contributions only from the propagation part of the object and a specific evanescent wave vector region minimizing the efforts of postprocessing algorithms. Note that there is only one transmission measurement required using each grating similar to the regular periodic case. However, the required measuring angle varies with each grating since kxb (corresponds to measuring angle) depends on kxa.

The ideal capability of image recovery using optimized gratings can be utilized if both amplitude and phase of the transmitted field can be measured along the required angle. Assuming Ameas(g)(kxb) is the measured complex amplitude of the field at a far-field distance in the corresponding kxb direction with a grating (g) and assuming only dominant elements of the transmission function contribute to the far field, an equation for the amplitude spectrum of the object can be written as

a(kxa)=Ameas(g)(kxb)TabGAmeas(0)(kxb)TaaG,
where TaaG and TabG represent matrix elements corresponding to the diagonal and central band of each grating, respectively, and Ameas(0) represents the complex field measured at the same angle without any grating.

Far-field transmission “measurements” through optimized gratings are taken using COMSOL for the same test object. The amplitude spectrum is calculated from the field measurement of each grating using Eq. (15). The image of the object is then recovered from the calculated amplitude spectrum using Eq. (3) and is presented in Fig. 3(d). The image clearly reconstructs all the subwavelength features of the object to the order of λ0/16 without any complex postprocessing methods. The calculated amplitude spectrum is shown in the inset of Fig. 3(d) in comparison to the actual amplitude spectrum. Relatively large discrepancies in the amplitudes are observed in the range of 0.5|kx/k0|3, where the contribution from the diagonal elements of the TG matrix is weak. These discrepancies resulted in oscillations of the order of λ0 in the image of the object. Note that here the recovery is not based on the pixel basis. Hence, the field of the object can be expressed as a continuous function in space.

The postprocessing of the data can also be performed using the same pixel-basis method described in Eqs. (11) and (13), which can handle intensity measurements. It will be shown in the next section that optimized gratings provide better imaging performance (with fewer artifacts) and high noise tolerance in comparison to periodic gratings.

5. NOISE TOLERANCE

In this section, we test and compare noise tolerance of the imaging process using regular periodic gratings and gratings with optimized unit cells. We performed the image reconstruction process [using Eqs. (11) and (13)] by adding multiple realizations of random noise patterns to the measured data proportional to its peak intensity. The calculated mean and standard deviation of field amplitudes of each pixel are plotted in Fig. 4(a), representing the imaging performance of regular periodic gratings at different noise levels. Clearly, the parasitic artifacts in the objects increase in amplitude, dissolving the sharpness of the edges, specifically in larger objects. At 20% noise, the recovery almost fails in distinguishing the central objects. Figure 4(b) represents image recoveries using gratings with optimized unit cells. Significant improvement in the imaging performance with minimum artifacts is observed for low noise levels in comparison to regular periodic gratings. Both the subwavelength features and features of wavelength order are clearly distinguished, even at higher noise levels. The error bar (standard deviation) in all cases is almost proportional to the noise level.

 figure: Fig. 4.

Fig. 4. Noise tolerance of the image reconstruction. Image reconstruction using (a) binary graphene gratings and (b) gratings with optimized unit cells by adding multiple realizations of random noise to the measured far-field intensity. The dotted line represents the object; the solid line and the shaded region represent the mean image and the standard deviation of all noise patterns, respectively. The addition of noise to the “measured” intensity is performed as Imeas;noise=|Hmeas+Nrmax(|Hmeas|)|2, where I represents intensity, H represents complex field, N is the noise percent level, and r is a random number between 1 and 1. In comparison to (a), (b) has less artifacts surrounding the object and reconstructs all the features of the object at noise levels as high as 20%.

Download Full Size | PDF

6. DISCUSSION

The imaging performance of the proposed GNOM technique depends on the key parameters of the system (i) the smallest achievable width or spacing between graphene strips, (ii) the minimum distance possible between the objects and the gratings, and (iii) the tunability range of the surface conductivity of the gratings via external stimuli. Even though the transmission function calculation in Eq. (10) and the pixel-basis approach described in Eq. (11) do not limit the resolution of the image, the reliable final resolution of the technique is determined by the above [particularly (i) and (iii)]. If one aims on resolving an object with resolution λ0/f (f>1), the maximum width of the smallest strip should be necessarily equal or less than λ0/f. The maximum distance between the object and the gratings should not exceed λ0/f (since evanescent components decay faster than λ0/f). While the above two are the necessary requirements, the tunability range of the surface conductivity determines the coupling efficiency of the evanescent waves to propagating waves. Note that the imaging coupling (imaging performance) weakly depends on the actual magnitude conductivity but strongly depends on the range of tunability of the conductivity. For example, one can design optimized grating profiles similar to Fig. 3(a), restricting the range of the Fermi level between 1 and 1 eV to obtain qualitatively similar imaging performance.

One of the known drawbacks of GNOM (or some diffraction-based techniques) is its strategy in handling imperfect measurements. In spatial scanning techniques, such as NSOM, one imperfect measurement usually affects the quality of only one pixel of the image. In contrast, the computational postprocessing of GNOM may distribute one imperfect measurement into a number of pixels or completely compensate it with the help of other accurate measurements. This property is evident from Fig. 4 where image recovery using data with higher noise levels does not necessarily affect the subwavelength features of the objects but rather increases the overall intensity level of the object (decreases the image contrast). Another known drawback is its inability to image only a specific section of the object where NSOM fits best.

In conclusion, we presented a new graphene-based high-resolution mechanical-scanning-free optical microscopy technique (GNOM) based on the principles of diffraction imaging. In contrast to mechanical scanning of a sharp tip (as in NSOM), we utilize active electronic scanning of diffraction properties of atomic thin graphene gratings placed in the near field of an object. We provided engineered designs of spatial patterns and optical properties of the gratings that can efficiently and systematically out-couple the evanescent information of the object to far fields mixed with diffraction orders. By defining a transmission function of graphene interface gratings, we then provided computational postprocessing methods to “unmix” the evanescent information and consequently recovered the detailed spatial field profile of the object. Using computational experiments, we demonstrated that the technique is capable of imaging objects with an overall size of 2λ0 and a pixel resolution of λ0/16. Numerical optimization based on the genetic algorithm is then used to design an optimum set of diffraction gratings that can potentially eliminate the requirement of postprocessing or at least minimize the numerical artifacts in the image. Noise tolerance and known drawbacks are discussed.

The proposed technique has the capability to transform into a fully static high-resolution imaging setup without the requirement of any linear or angular movements of objects, sources, or tips. The variation of optical properties of graphene gratings using electrostatic bias can be performed dramatically faster by making the technique more suitable for real-time imaging scenarios. Considering the extensive ongoing research on active tunable plasmonic materials at visible and near-IR frequencies [1922], the proposed technique has more avenues to improve and extend to a broad range of frequencies.

Funding

Army Research Office (ARO) (W911NF-14-0247); Air Force Office of Scientific Research (AFOSR) (FA9550-14-1-0349).

REFERENCES

1. M. J. Rust, M. Bates, and X. Zhuang, “Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM),” Nat. Methods 3, 793–796 (2006). [CrossRef]  

2. S. T. Hess, T. P. Girirajan, and M. D. Mason, “Ultra-high resolution imaging by fluorescence photoactivation localization microscopy,” Biophys. J. 91, 4258–4272 (2006). [CrossRef]  

3. E. Betzig, G. H. Patterson, R. Sougrat, O. W. Lindwasser, S. Olenych, J. S. Bonifacino, M. W. Davidson, J. Lippincott-Schwartz, and H. F. Hess, “Imaging intracellular fluorescent proteins at nanometer resolution,” Science 313, 1642–1645 (2006). [CrossRef]  

4. P. Doradla, K. Alavi, C. Joseph, and R. Giles, “Detection of colon cancer by continuous-wave terahertz polarization imaging technique,” J. Biomed. Opt. 18, 090504 (2013). [CrossRef]  

5. E. Abbe, “Beiträge zur Theorie des Mikroskops und der mikroskopischen Wahrnehmung,” Archiv für mikroskopische Anatomie 9, 413–418 (1873). [CrossRef]  

6. U. Dürig, D. Pohl, and F. Rohner, “Near-field optical-scanning microscopy,” J. Appl. Phys. 59, 3318–3327 (1986). [CrossRef]  

7. P. Doradla, K. Alavi, C. Joseph, and R. Giles, “Single-channel prototype terahertz endoscopic system,” J. Biomed. Opt. 19, 080501 (2014). [CrossRef]  

8. P. Doradla, K. Alavi, C. S. Joseph, and R. H. Giles, “Continuous wave terahertz reflection imaging of human colorectal tissue,” Proc. SPIE 8624, 86240O (2013). [CrossRef]  

9. C. S. Joseph, R. Patel, V. A. Neel, R. H. Giles, and A. N. Yaroslavsky, “Imaging of ex vivo nonmelanoma skin cancers in the optical and terahertz spectral regions optical and terahertz skin cancers imaging,” J. Biophoton. 7, 295–303 (2014). [CrossRef]  

10. A. Forouzmand, H. M. Bernety, and A. B. Yakovlev, “Graphene-loaded wire medium for tunable broadband subwavelength imaging,” Phys. Rev. B 92, 085402 (2015). [CrossRef]  

11. A. Forouzmand and A. B. Yakovlev, “Tunable dual-band subwavelength imaging with a wire medium slab loaded with nanostructured graphene metasurfaces,” AIP Adv. 5, 077108 (2015). [CrossRef]  

12. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, “Electric field effect in atomically thin carbon films,” Science 306, 666–669 (2004). [CrossRef]  

13. K. Novoselov, A. K. Geim, S. Morozov, D. Jiang, M. Katsnelson, I. Grigorieva, S. Dubonos, and A. Firsov, “Two-dimensional gas of massless Dirac fermions in graphene,” Nature 438, 197–200 (2005). [CrossRef]  

14. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotechnol. 7, 699–712 (2012). [CrossRef]  

15. A. Ayari, E. Cobas, O. Ogundadegbe, and M. S. Fuhrer, “Realization and electrical characterization of ultrathin crystals of layered transition-metal dichalcogenides,” J. Appl. Phys. 101, 14507 (2007). [CrossRef]  

16. M. Osada and T. Sasaki, “Two-dimensional dielectric nanosheets: novel nanoelectronics from nanocrystal building blocks,” Adv. Mater. 24, 210–228 (2012). [CrossRef]  

17. Z. Jacob, “Nanophotonics: hyperbolic phonon-polaritons,” Nat. Mater. 13, 1081–1083 (2014). [CrossRef]  

18. J. Cheng, W. L. Wang, H. Mosallaei, and E. Kaxiras, “Surface plasmon engineering in graphene functionalized with organic molecules: a multiscale theoretical investigation,” Nano Lett. 14, 50–56 (2013). [CrossRef]  

19. S. N. Shirodkar and E. Kaxiras, “Li intercalation at graphene/hexagonal boron nitride interfaces,” Phys. Rev. B 93, 245438 (2016). [CrossRef]  

20. S. N. Shirodkar, M. Mattheakis, P. Cazeaux, P. Narang, M. Soljačić, and E. Kaxiras, “Visible quantum plasmons in highly-doped few-layer graphene,” arXiv:1703.01558 (2017).

21. H. W. Lee, G. Papadakis, S. P. Burgos, K. Chander, A. Kriesch, R. Pala, U. Peschel, and H. A. Atwater, “Nanoscale conducting oxide PlasMOStor,” Nano Lett. 14, 6463–6468 (2014). [CrossRef]  

22. A. Forouzmand and H. Mosallaei, “Tunable two dimensional optical beam steering with reconfigurable indium tin oxide plasmonic reflectarray metasurface,” J. Opt. 18, 125003 (2016). [CrossRef]  

23. S. Durant, Z. Liu, J. M. Steele, and X. Zhang, “Theory of the transmission properties of an optical far-field superlens for imaging beyond the diffraction limit,” J. Opt. Soc. Am. B 23, 2383–2392 (2006). [CrossRef]  

24. M. G. Gustafsson, “Surpassing the lateral resolution limit by a factor of two using structured illumination microscopy,” J. Microsc. 198, 82–87 (2000). [CrossRef]  

25. A. C. Kak and M. Slaney, Principles of Computerized Tomographic Imaging (IEEE, 1988).

26. Z. Liu, S. Durant, H. Lee, Y. Pikus, N. Fang, Y. Xiong, C. Sun, and X. Zhang, “Far-field optical superlens,” Nano Lett. 7, 403–408 (2007). [CrossRef]  

27. G. Popescu, L. L. Goddard, P. S. Carney, T. Kim, R. Zhou, M. A. Mir, and S. D. Babacan, “White light diffraction tomography of unlabeled live cells,” Nat. Photon. 8, 256–263 (2014). [CrossRef]  

28. S. Inampudi, N. Kuhta, and V. A. Podolskiy, “Interscale mixing microscopy: numerically stable imaging of wavelength-scale objects with sub-wavelength resolution and far field measurements,” Opt. Express 23, 2753–2763 (2015). [CrossRef]  

29. C. M. Roberts, N. Olivier, W. P. Wardley, S. Inampudi, W. Dickson, A. V. Zayats, and V. A. Podolskiy, “Interscale mixing microscopy: far-field imaging beyond the diffraction limit,” Optica 3, 803–808 (2016). [CrossRef]  

30. S. Thongrattanasiri, N. Kuhta, M. Escarra, A. Hoffman, C. Gmachl, and V. Podolskiy, “Analytical technique for subwavelength far field imaging,” Appl. Phys. Lett. 97, 101103 (2010). [CrossRef]  

31. W. Fan, B. Yan, Z. Wang, and L. Wu, “Three-dimensional all-dielectric metamaterial solid immersion lens for subwavelength imaging at visible frequencies,” Sci. Adv. 2, e1600901 (2016). [CrossRef]  

32. C. Park, J.-H. Park, C. Rodriguez, H. Yu, M. Kim, K. Jin, S. Han, J. Shin, S. H. Ko, and K. T. Nam, “Full-field subwavelength imaging using a scattering superlens,” Phys. Rev. Lett. 113, 113901 (2014). [CrossRef]  

33. A. Sentenac, P. C. Chaumet, and K. Belkebir, “Beyond the Rayleigh criterion: grating assisted far-field optical diffraction tomography,” Phys. Rev. Lett. 97, 243901 (2006). [CrossRef]  

34. V. Lauer, “New approach to optical diffraction tomography yielding a vector equation of diffraction tomography and a novel tomographic microscope,” J. Microsc. 205, 165–176 (2002). [CrossRef]  

35. R. M. Silver, B. M. Barnes, R. Attota, J. Jun, M. Stocker, E. Marx, and H. J. Patrick, “Scatterfield microscopy for extending the limits of image-based optical metrology,” Appl. Opt. 46, 4248–4257 (2007). [CrossRef]  

36. J. Chen, M. Badioli, P. Alonso-González, S. Thongrattanasiri, F. Huth, J. Osmond, M. Spasenović, A. Centeno, A. Pesquera, and P. Godignon, “Optical nano-imaging of gate-tunable graphene plasmons,” Nature 487, 77–81 (2012).

37. L. Ju, B. Geng, J. Horng, C. Girit, M. Martin, Z. Hao, H. A. Bechtel, X. Liang, A. Zettl, and Y. R. Shen, “Graphene plasmonics for tunable terahertz metamaterials,” Nat. Nanotechnol. 6, 630–634 (2011). [CrossRef]  

38. C. H. Lui, Z. Li, K. F. Mak, E. Cappelluti, and T. F. Heinz, “Observation of an electrically tunable band gap in trilayer graphene,” Nat. Phys. 7, 944–947 (2011). [CrossRef]  

39. Y. Zhang, T.-T. Tang, C. Girit, Z. Hao, M. C. Martin, A. Zettl, M. F. Crommie, Y. R. Shen, and F. Wang, “Direct observation of a widely tunable bandgap in bilayer graphene,” Nature 459, 820–823 (2009). [CrossRef]  

40. G. W. Hanson, “Dyadic Green’s functions and guided surface waves for a surface conductivity model of graphene,” J. Appl. Phys. 103, 064302 (2008). [CrossRef]  

41. S. Inampudi, M. Nazari, A. Forouzmand, and H. Mosallaei, “Manipulation of surface plasmon polariton propagation on isotropic and anisotropic two-dimensional materials coupled to boron nitride heterostructures,” J. Appl. Phys. 119, 025301 (2016). [CrossRef]  

42. A. Woessner, M. B. Lundeberg, Y. Gao, A. Principi, P. Alonso-González, M. Carrega, K. Watanabe, T. Taniguchi, G. Vignale, and M. Polini, “Highly confined low-loss plasmons in graphene-boron nitride heterostructures,” Nat. Mater. 14, 421–425 (2015). [CrossRef]  

43. F. H. Koppens, D. E. Chang, and F. J. García de Abajo, “Graphene plasmonics: a platform for strong light-matter interactions,” Nano Lett. 11, 3370–3377 (2011). [CrossRef]  

44. C.-F. Chen, C.-H. Park, B. W. Boudouris, J. Horng, B. Geng, C. Girit, A. Zettl, M. F. Crommie, R. A. Segalman, and S. G. Louie, “Controlling inelastic light scattering quantum pathways in graphene,” Nature 471, 617–620 (2011). [CrossRef]  

45. D. K. Efetov and P. Kim, “Controlling electron-phonon interactions in graphene at ultrahigh carrier densities,” Phys. Rev. Lett. 105, 256805 (2010). [CrossRef]  

46. J. Cheng, S. Jafar-Zanjani, and H. Mosallaei, “Real-time two-dimensional beam steering with gate-tunable materials: a theoretical investigation,” Appl. Opt. 55, 6137–6144 (2016). [CrossRef]  

47. M. Zeng, Y. Xiao, J. Liu, W. Lu, and L. Fu, “Controllable fabrication of nanostructured graphene towards electronics,” Adv. Electron. Mater. 2, 1500456 (2016). [CrossRef]  

48. COMSOL Multiphysics v. 5.2, “COMSOL AB,” 2015, http://www.comsol.com.

49. L. Falkovsky, “Optical properties of graphene,” in Journal of Physics: Conference Series (IOP Publishing, 2008), 012004.

50. L. Falkovsky and A. Varlamov, “Space–time dispersion of graphene conductivity,” Eur. Phys. J. B 56, 281–284 (2007). [CrossRef]  

51. L. Falkovsky and S. Pershoguba, “Optical far-infrared properties of a graphene monolayer and multilayer,” Phys. Rev. B 76, 153410 (2007). [CrossRef]  

52. T. Stauber, N. Peres, and A. Geim, “Optical conductivity of graphene in the visible region of the spectrum,” Phys. Rev. B 78, 085432 (2008). [CrossRef]  

53. K. F. Mak, M. Y. Sfeir, Y. Wu, C. H. Lui, J. A. Misewich, and T. F. Heinz, “Measurement of the optical conductivity of graphene,” Phys. Rev. Lett. 101, 196405 (2008). [CrossRef]  

54. A. Kuzmenko, E. Van Heumen, F. Carbone, and D. Van Der Marel, “Universal optical conductance of graphite,” Phys. Rev. Lett. 100, 117401 (2008). [CrossRef]  

55. J. M. Dawlaty, S. Shivaraman, J. Strait, P. George, M. Chandrashekhar, F. Rana, M. G. Spencer, D. Veksler, and Y. Chen, “Measurement of the optical absorption spectra of epitaxial graphene from terahertz to visible,” Appl. Phys. Lett. 93, 131905 (2008). [CrossRef]  

56. Y. Yang, G. Yu, J. J. Cha, H. Wu, M. Vosgueritchian, Y. Yao, Z. Bao, and Y. Cui, “Improving the performance of lithium–sulfur batteries by conductive polymer coating,” ACS Nano 5, 9187–9193 (2011). [CrossRef]  

57. S. Inampudi, D. Adams, T. Ribaudo, D. Slocum, S. Vangala, W. Goodhue, D. Wasserman, and V. Podolskiy, “ϵ-near-zero enhanced light transmission through a subwavelength slit,” Phys. Rev. B 89, 125119 (2014). [CrossRef]  

58. Z. Fei, A. Rodin, W. Gannett, S. Dai, W. Regan, M. Wagner, M. Liu, A. McLeod, G. Dominguez, and M. Thiemens, “Electronic and plasmonic phenomena at graphene grain boundaries,” Nat. Nanotechnol. 8, 821–825 (2013). [CrossRef]  

59. E. Kaxiras, Atomic and Electronic Structure of Solids (Cambridge University, 2003).

60. R. H. Byrd, M. E. Hribar, and J. Nocedal, “An interior point algorithm for large-scale nonlinear programming,” SIAM J. Optim. 9, 877–900 (1999). [CrossRef]  

61. M. J. Powell, “A fast algorithm for nonlinearly constrained optimization calculations,” in Numerical Analysis (Springer, 1978), pp. 144–157.

62. A. Conn, N. Gould, and P. Toint, “A globally convergent Lagrangian barrier algorithm for optimization with general inequality constraints and simple bounds,” Math. Comput. Am. Math. Soc. 66, 261–289 (1997). [CrossRef]  

63. D. E. Goldberg and J. H. Holland, “Genetic algorithms and machine learning,” Mach. Learn. 3, 95–99 (1988). [CrossRef]  

64. C. M. Roberts, S. Inampudi, and V. A. Podolskiy, “Diffractive interface theory: nonlocal susceptibility approach to the optics of metasurfaces,” Opt. Express 23, 2764–2776 (2015). [CrossRef]  

65. J. Cheng, D. Ansari-Oghol-Beig, and H. Mosallaei, “Wave manipulation with designer dielectric metasurfaces,” Opt. Lett. 39, 6285–6288 (2014). [CrossRef]  

66. J. Cheng and H. Mosallaei, “Optical metasurfaces for beam scanning in space,” Opt. Lett. 39, 2719–2722 (2014). [CrossRef]  

67. M. Farmahini-Farahani, J. Cheng, and H. Mosallaei, “Metasurfaces nanoantennas for light processing,” J. Opt. Soc. Am. B 30, 2365–2370 (2013). [CrossRef]  

68. Z. Li, K. Yao, F. Xia, S. Shen, J. Tian, and Y. Liu, “Graphene plasmonic metasurfaces to steer infrared light,” Sci. Rep. 5, 12423 (2015). [CrossRef]  

69. N. Yu and F. Capasso, “Flat optics with designer metasurfaces,” Nat. Mater. 13, 139–150 (2014). [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 (4)

Fig. 1.
Fig. 1. (a) Schematic setup of GNOM. Objects are placed on a substrate (ε=2.25). Graphene sheet slices with electrodes are placed at a distance h on top of the object. Red and black colors of the graphene slices represent different values of surface conductivity. The inset is the schematic for transmission function calculation. (b), (c) The relative variation in field amplitude at angles of 0° and 30°, respectively, for five different objects numbered 1–5 shown as insets. Object 1total size(5λ0), Object 2total size(3λ0), Object 3total size(2λ0), Object 4total size(2λ0) with subwavelength features (λ0/8, λ0/4, λ0/2, λ0/8, λ0/4), and Object 5total size(2λ0) with subwavelength feature sizes (λ0/16, λ0/8, λ0/4, λ0/2, λ0/16, λ0/4). The gaps between the subwavelength features are of relative size. For objects with the smallest subwavelength features (more evanescent spectrum) the variation in relative amplitude is high and irregular (nonredundant) in pattern. |H| and |H0| represent the amplitude of the field with and without a grating, respectively.
Fig. 2.
Fig. 2. Image reconstruction using binary graphene gratings. (a) Transmitted far-field intensity “measurements” of the test object (calculated using COMSOL) in the normal direction with grating periodicity reconfigured from Λ=λ0 to Λ=λ0/16 in 300 steps, λ0=10μm. The unit cell has two equal size components with σ1Ef and σ2=104. The object-to-grating distance is h=100nm. (b) The corresponding transmission function matrix (TG) with rows as the rows of the transmission function of each grating in the normal direction calculated using Eq. (10), where color represents the magnitude of the matrix element. Upper and lower sections represent gratings with the two Fermi levels of graphene. (c) The test object (dotted line) and the reconstructed image (solid line) using Eq. (13). The overall size of the object is 2λ0, and the subwavelength feature sizes are (λ0/16, λ0/8, λ0/4, λ0/2, λ0/4) with gaps of relative size between them. The reconstructed image resolves the gaps and features of the object.
Fig. 3.
Fig. 3. Image reconstruction using gratings with optimized unit cells. (a) Fermi level, consequently the surface conductivity, profile of unit cells optimized to maximize the coupling of each wave vector (kxa) to the corresponding diffraction order in the propagation regime (Λ=λ0, λ0=10μm). Each unit cell has 32 elements. A color bar is displayed for clarity. (b) Line plots of the same quantity for three representative wave vectors. (c) The elements of composite transmission matrix TG, where each row is the row of the transmission function calculated using Eq. (10) for corresponding transmission direction of each optimized grating. (d) The real part of the magnetic field of the object (black dashed line) from COMSOL and the corresponding image (solid and dotted lines) recovered analytically using Eqs. (3) and (15) with no numerical postprocessing. The inset shows the corresponding amplitude spectrum of the object and image. The electric and magnetic fields of the image are computed using Eq. (4) as H=Hz and E=Ex2+Ez2.
Fig. 4.
Fig. 4. Noise tolerance of the image reconstruction. Image reconstruction using (a) binary graphene gratings and (b) gratings with optimized unit cells by adding multiple realizations of random noise to the measured far-field intensity. The dotted line represents the object; the solid line and the shaded region represent the mean image and the standard deviation of all noise patterns, respectively. The addition of noise to the “measured” intensity is performed as Imeas;noise=|Hmeas+Nrmax(|Hmeas|)|2, where I represents intensity, H represents complex field, N is the noise percent level, and r is a random number between 1 and 1. In comparison to (a), (b) has less artifacts surrounding the object and reconstructs all the features of the object at noise levels as high as 20%.

Equations (15)

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

kxm=kx0+m2πΛ,
σ=e2Efε0cπ211τiω,
Hobj(x)=a+(kx)eikxxdkx=wl=al+(kxl)eikxlx,
Hy{I,II}(x,z)=m=[am+{I,II}eikzm{I,II}z+am{I,II}eikzm{I,II}z]eikxmx,Ex{I,II}(x,z)=m=kzm{I,II}k0ε{I,II}[am+{I,II}eikzm{I,II}zam{I,II}eikzm{I,II}z]eikxmx,Ez{I,II}(x,z)=m=kxmk0ε{I,II}[am+{I,II}eikzm{I,II}z+am{I,II}eikzm{I,II}z]eikxmx.
kzm{I,II}=k02ε{I,II}kxm2,
ExI(x,z=0)ExII(x,z=0)=0,HyI(x,z=0)HyII(x,z=0)=σ(x)ExII(x,z=0).
1εIKzI(AI+AI)1εIIKzII(AII+AII)=0,(AI+AI)(AII+AII)=1k0εIIΞKzII(AII+AII).
[AIAII+]=[R+TT+R][AI+AII],
R+=[k0(εIIKzI+εIKzII)+KzIIΞKzI]1[k0(εIIKzIεIKzII)+KzIIΞKzI],T=[k0(εIIKzI+εIKzII)+KzIIΞKzI]1[2k0εIKzII],T+=[k0(εIIKzI+εIKzII)+KzIΞKzII]1[2k0εIIKzI],R=[k0(εIIKzI+εIKzII)+KzIΞKzII]1[k0(εIIKzIεIKzII)+KzIΞKzII].
T=[Tm1,n1Tm1,nTm1,n+1Tm,n1Tm,nTm,n+1Tm+1,n1Tm1,nTm+1,n+1],
TgpF=m=TgmGexp(ikxmxp)sin(kxmP/2)/(kxmP/2).
IgC=|ag|2=|pTgpFhp|2.
|IgCImeas|2min.
n|Tbn+δan|2min.
a(kxa)=Ameas(g)(kxb)TabGAmeas(0)(kxb)TaaG,
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.