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

Ultracompact electro-optic waveguide modulator based on a graphene-covered λ/1000 plasmonic nanogap

Open Access Open Access

Abstract

The extreme field confinement and electro-optic tunability of plasmons in graphene make it an ideal platform for compact waveguide modulators, with device footprints aggressively scaling orders of magnitude below the diffraction limit. The miniaturization of modulators based on graphene plasmon resonances is however inherently constrained by the plasmon wavelength, while their performance is bounded by material loss in graphene. In this report, we propose to overcome these limitations using a graphene-covered λ/1000 plasmonic nanogap waveguide that concentrates light on length scales more than an order of magnitude smaller than the graphene plasmon wavelength. The modulation mechanism relies on interference between the non-resonant background transmission and the transmission mediated by the gate-tunable nanogap mode, enabling modulation depths over 20 dB. Since the operation of the device does not rely on graphene plasmons, the switching behavior is robust against low graphene carrier mobility even under 1000 cm2/Vs, which is desirable for practical applications.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The realization of optical devices unconstrained by the diffraction limit has been a central focus of plasmonics, which capitalizes on the ability of plasmons—coherent charge density oscillations in metals—to concentrate light on length scales well below its wavelength in free space. In particular, metallic 2D materials (graphene [1], black phosphorus [2], etc.) naturally support low-dimensional surface plasmon-polaritons with wavelengths as short as λ0/100, where λ0 denotes the free-space wavelength. Furthermore, large variations in the Fermi level of 2D materials can be electrostatically controlled by gating, which substantially modulates the associated plasmonic response [3]. As a representative example, graphene plasmons (GPs) have been widely employed in a wide range of devices such as active metasurfaces [3,4], molecular sensors [57], and photodetectors [8], along with polaritons in other van der Waals crystals [9,10]. More recently, it has been discovered that the polaritonic modes in 2D materials can be atomically confined also in the transverse direction by sandwiching a thin dielectric layer by a 2D material and a metal substrate [1114].

The tunable optical properties of two-dimensional materials have also been utilized to develop compact electro-optic waveguide modulators [1517], where the extremely confined nature of 2D polaritonic modes can dramatically reduce device footprints down to the extreme sub-diffraction regime [18,19]. An ultracompact device footprint is desirable not only for high-density on-chip integration, but also for low device capacitance, which enables high-speed operation with low switching energy [20]. In such devices, the minimum lateral dimension is fundamentally dictated by the polariton wavelength (λp). For example, two recently proposed graphene plasmon based waveguide modulators have lateral sizes of about λpλ0/50 at mid-infrared frequencies [18,19]. Furthermore, the modulation performance of most graphene plasmonic devices strongly depend on the carrier mobility of graphene, which places practical limits on the large-scale realization of these devices using graphene grown by chemical vapor deposition (CVD).

In this paper, we theoretically demonstrate that an efficient transmission modulation of mid-infrared plasmonic waveguide modes can be achieved even with an active device length of λ0/1000, which is more than an order of magnitude shorter than the graphene plasmon wavelength. Here, the non-resonant background transmission interferes with transmission through the λ0/1000 nanogap mode, where the latter can be electrostatically tuned by altering the carrier density of graphene. The transmission modulation properties can be elaborately tailored by engineering the geometrical parameters of the device, and the resulting modulation depth can exceed 20 dB. In contrast to previously investigated structures [18,19], the operation of this device does not rely on the graphene plasmon resonances. Consequently, its switching behavior is robust against low graphene carrier mobility, which is strongly desirable for practical applications.

2. Device structure and transmission modulation characteristics

The proposed nanogap plasmonic switch consists of a graphene-insulator-metal structure connecting two identical metal-insulator-metal (MIM) waveguides, which serve as the input and output optical channels, as illustrated in Fig. 1(a). The carrier concentration in graphene can be tuned electrostatically by applying a gate bias voltage between the graphene sheet and the underlying metal. The materials of the insulating core and metal claddings of the MIM waveguides are SiO2 and Au, respectively. The target free space wavelength is chosen to be λ0 = 7.5 µm, at which the optical absorption in SiO2 [21] and the optical phonon scattering in graphene [22,23] are negligible. The thickness of the core (d) is chosen to be far below the diffraction limit, which enables an efficient coupling between the waveguide mode and the nanogap cavity mode. With the deep subwavelength thickness, the MIM waveguides only support the fundamental transverse magnetic mode (TM0). The thickness of the cladding (h) is set to be larger than the skin depth of Au to prevent leakage of electromagnetic energy from the waveguides.

 figure: Fig. 1.

Fig. 1. (a) Schematic of the MIM waveguide-integrated plasmonic modulator. Each MIM waveguide supports the TM0 guided mode whose transmission is modulated by the gate voltage. For the carrier mobility of graphene $\mu\; = \; 10,000\; \textrm{c}{\textrm{m}^\textrm{2}}\textrm{/Vs}$ and a fixed ${{\lambda }_\textrm{0}} = \; 7.5\; \mathrm{\mu}\textrm{m}$, the optimized geometric parameters are: core thickness d = 50 nm, cladding thickness h = 30 nm, and gap width g = 5 nm. (b) The transmittance T map as functions of Fermi level EF and the geometric parameter g. (c) The distribution of the normalized Re{Ex} inside the each nanocavity geometries at minimum transmittance when EF = 0.24 eV (left), and EF = 0.26 (right).

Download Full Size | PDF

We solve Maxwell’s equations and characterize the waveguide response by using the commercial finite element method (FEM) software COMSOL Multiphysics. The port boundary condition is applied at the input and output terminations of the waveguide in order to calculate the amplitude and phase of the transmission and reflection coefficients for the TM0 mode. The reflection and transmission coefficients are measured at the output ports, and the propagation loss and phase delay in the waveguide region is compensated as described in Appendix 3. Graphene is modeled as a thin layer with an effective relative permittivity $\epsilon = 1 + i\sigma /{\epsilon _0}\omega \tau $, where the layer thickness $\tau $ is assumed to be 0.3 nm. The optical conductivity of graphene $\sigma $ is calculated within the random phase approximation in local limit as a function of frequency $\omega $ and Fermi level EF [24]. Dielectric functions of Au and SiO2 are taken from [21].

Full-wave electromagnetic simulations reveal that the transmittance (T) of the MIM waveguide mode across the gap exhibits a strong dependence on the gap width (g) and the Fermi energy of graphene (EF) as shown in Fig. 1(b). Here, d and h are chosen to be 50 nm and 30 nm, respectively. Interestingly, we find that the resulting resonant transmission exhibits qualitatively different behavior depending on the gap width. For g > 40 nm, where the gap width is comparable to the graphene plasmon wavelength, the transmission across the gap exhibits a highly asymmetric Fano-like profile. The Fano resonance originates from the interference between the resonant transmission (mediated by the graphene plasmon modes) and the non-resonant background transmission, as recently explained by Jang et al. [18]. The n-th order resonance condition for graphene plasmons bounded by metallic walls, which provide a nearly ${\pi }$ phase shift upon GP reflection, can be analytically deduced from the phase matching condition, $\phi _{\textrm{GM}}^\textrm{r}{\; + \; }{{\beta }_\textrm{G}}{g}\; = \; ({{n + 1}} ){\pi }$ [25], where $\phi _{\textrm{GM}}^\textrm{r}$ is the reflection phase shift of graphene plasmons at the graphene-MIM waveguide boundary and ${{\beta }_\textrm{G}}$ is the propagation constant of the graphene plasmon mode that emerges for EF > 0.08 eV. The black dashed curves in Fig. 1(b) indicate the dispersions of zeroth, first, and second order graphene plasmon resonance. Note that, for g > 40 nm, the transmission modulation occurs at the vicinity of the graphene plasmon resonance as explained earlier.

Conversely, as the gap becomes much narrower than the graphene plasmon wavelength, there appears a new, unexpected transmission modulation regime, where the dependence of T on EF and g is significantly different from what was described earlier. In this regime, the transmission line shape becomes much broader and more symmetric. More importantly, the transmission modulation no longer occurs at the Fermi level associated with a graphene plasmon resonance at the operation wavelength. We attribute this discrepancy partly to the inherently broad nature of the zeroth-order graphene plasmon resonance. Since the gap width is much smaller than the graphene plasmon wavelength, the propagation phase of graphene plasmon, ${{\beta }_\textrm{G}}{g}$, becomes marginal while the graphene-metal reflection phase shift $\phi _{\textrm{GM}}^\textrm{r}$ remains nearly ${\pi }$ [18]; consequently, the phase matching condition, $\phi _{\textrm{GM}}^\textrm{r}{\; + \; }{{\beta }_\textrm{G}}{g}\; = \; {\pi }$, is roughly satisfied even though EF deviates significantly from the exact matching condition. In addition, as the gap width is reduced to lengths of a few nm, the nanogap mode, which is confined in the gap between the top metal claddings, begins to play an important role in mode conversion dynamics. The difference between the two regimes is also evident in the electric field distributions near the nanogap region at EF, where transmission modulation occurs. In the wide gap regime (g = 80 nm), the electric field is concentrated on the surface of graphene and decays as it gets further away (Fig. 1(c)). In contrast, the electric field in a narrow gap (g = 5 nm) is confined in the air gap region and does not show vertical intensity variation inside the gap, implying that this phenomenon is closely related to the nanogap mode rather than the graphene plasmon mode.

3. Operation principles

To unravel the operation mechanism of the nanogap transmission modulation, we decompose the net transmission into two channels: the background transmission (tB) and the transmission through the nanogap mode (tNG), as schematically shown at Fig. 2(a). The background transmission tB is inherently independent of EF and can be estimated from the transmission of the nanogap structure without the graphene sheet (see Appendix 3). In contrast, the transmission through the nanogap mode tNG varies as a function of EF. The nanogap mode corresponds to a vertically propagating plasmonic mode squeezed in between the left and right top metal claddings in the transverse direction, as illustrated in Fig. 2(a). Once the input MIM waveguide mode impinges on the nanogap region, a certain fraction of the wave couples to the nanogap mode. The nanogap mode then propagates up and down along the gap and experiences reflection at the top and bottom ends of the gap. Each time the nanogap mode interacts with the graphene-covered bottom end of the gap, some fraction of the wave outcouples to the output MIM waveguide. This process can be analytically described by using a modified Fabry-Perot model for the nanogap transmission,

$$ {{{t}_{\textrm{NG}}}\textrm{ = }\frac{{{{c}_{\textrm{MN}}}{{c}_{\textrm{NM}}}{{e}^{{i}{{\beta }_{\textrm{NG}}}{h}}}{{r}_{\textrm{NF}}}}}{{\textrm{1 - }{{r}_{\textrm{NM}}}{{r}_{\textrm{NF}}}{{e}^{{i2}{{\beta }_{\textrm{NG}}}{h}}}}}} $$
where cij and rij denote the coupling and reflection coefficient with i,j = N (nanogap), M (MIM waveguide), or F (free space), which are individually determined by separate full-field electromagnetic simulations as described in the Appendix 3, and ${{\beta }_{\textrm{NG}}}$ is the propagation constant of the nanogap mode.

 figure: Fig. 2.

Fig. 2. (a) Illustration of the nanogap coupling and transmission mechanism. The nanogap transmission is decomposed into independent terms of mode coupling c and reflection r, where the gap-confined MIM mode has the propagation constant βNG. For a device geometry characterized by d = 50 nm, h = 30 nm, g = 5 nm, we present: (b) amplitude (red) and phase (blue) of the coupling cNM = −cMN (dashed) and reflection rNM (solid) coefficients; (c) transmission through the nanogap mode (tNG, dashed) and background transmission (tB, solid); (d) Net transmission amplitude (|tNG + tB|) obtained from the analytical model (dashed) and full-wave simulations (solid). In all simulations, graphene carrier mobility $\mu{\; = \; 10,000\; }{\textrm{cm}^\textrm{2}}\textrm{/Vs}$.

Download Full Size | PDF

For a specific geometry, ${\; }{{\beta }_{\textrm{NG}}}$ and rNF are independent of the graphene properties, while cMN and cNM depend weakly only on EF (Fig. 2(b)), so that the modulation of tNG is mainly achieved by the dependence of rNM on the graphene carrier concentration. We note that the phase factors of rNF and rNM are close to zero as the wave impedances of the free space and the dielectric are sufficiently high, and the phase of cNM and cMN are close to zero and π, respectively, as imposed by the continuity of the electric field at the boundary. As a result, the net phase of tNG roughly varies around π, which is ideal to achieve destructive interference with tB and therefore an efficient modulation of waveguide transmission. Figure 2(c) shows the resulting amplitude and phase of tNG as a function of EF for g = 5 nm, d = 50 nm, and h = 30 nm, which exhibits a broad and symmetric Lorentzian-like resonant feature. The amplitude of tNG has a maximum at the center Fermi level of EF = 0.26 eV, and its phase linearly decreases from 1.24π to 0.67π as EF varies from 0.1 to 0.5 eV.

Figure 2(d) compares the net transmission amplitude |tNG + tB| obtained from full-wave simulations and analytical modeling, and confirms that the analytical model successfully captures the dependence of T on EF. At EF = 0.26 eV, the phase difference between tNG and tB becomes ${\pi }$ while their amplitudes are similar, which induces the transmission minimum by destructive interference (“off” state, Toff). By defining the charge neutrality point as the “on” state, Ton, the on/off transmittance ratio (modulation depth) exceeds 100. The required gate bias difference for the switching operation is calculated to be 11.5 V, a value which can be further reduced by setting the on-state EF closer to the off-state EF. We also note that the active area of the device is unprecedentedly small, since the gap width is more than 1000 times shorter than the free space wavelength, and more than 10 times shorter than the graphene plasmon wavelength. This extreme miniaturization is advantageous not only for high-density integration purposes, but also for high-speed and energy efficient operations owing to the reduced device capacitance.

4. Structural parameter and carrier mobility dependence

The transmission modulation characteristics of the device can be altered by engineering the geometric parameters of the system. As shown in Fig. 1(b), in the narrow gap regime (g < 40 nm), the transmittance dip becomes broader and is shifted to a higher EF with increasing g. The broadening of the transmittance dip can be attributed to the fact that the nanogap mode becomes less confined and thus more efficiently leaks out to free space (i.e., |rNM| and |rNF| decrease) as the gap widens. Figure 3(a) shows that, despite broadening of the tNG, the difference between |tNG| and |tB| stays similar at the destructive interference condition. Consequently, the minimum transmittance value marginally depends on the gap width. The increasing gap width also affects the phase of tNG as shown in Fig. 3(b) (note that the phase of the background transmission tB is nearly zero regardless of the gap size). As the gap widens, more carriers in graphene are required to induce destructive interference between tNG and tB with their phase difference being equal to π, which explains the transmittance dip shift toward higher EF. The dependence of the coupling and reflection coefficients on the gap width is discussed in Appendix 4.

 figure: Fig. 3.

Fig. 3. (a) Amplitude of the transmission coefficients of the nanogap mode, tNG, and the background mode, tB, as a function of Fermi level EF and gap width g. (b) Phase difference of tNG and tB, ${\phi _{\textrm{NG}}} - {\phi _\textrm{B}}$, as a function of EF and g. Black dotted line indicates the destructive interference condition when ${\phi _{\textrm{NG}}} - {\phi _\textrm{B}} = \pi $. Here, d = 50 nm and h = 30 nm.

Download Full Size | PDF

Figure 4(a) illustrates the dependence of the net transmittance T(EF) on the cladding thickness h for a given core thickness d = 50 nm. We note that the minimum transmittance Fermi level is linearly dependent on h when d and g are fixed to 50 and 5 nm, respectively (Fig. 4(b)). This can be understood by recognizing that the increase of the metal cladding thickness $(\Delta {h})$ is equivalent to the addition of $2i{{\beta }_{\textrm{NG}}}\Delta {h}$ phase shift of the propagating nanogap mode in the modified Fabry-Perot model (Eq. (1)), leading to a phase shift of tNG relative to tB. Consequently, the destructive interference between tNG and tB occurs at a shifted Fermi level as shown in Fig. 4(c). On the other hand, the increasing core thickness d reduces the effective index of the guided mode, resulting in the decrease (increase) of coupling (reflection) coefficients, which makes the transmittance dip narrower, as shown in Fig. 4(d). As d increases, the amplitude of the background transmission |tB| increases (Fig. 4(e)), leading to an increased Ton. However, in this case, the EF at which the minimum transmittance occurs stays nearly the same, since the phase difference between tNG and tB is preserved (Fig. 4(f)). At the same time, this also causes Toff to rise by making the destructive interference between tB and tNG less complete due to their enlarged amplitude mismatch. Consequently, the modulation depth (Ton/Toff) of the device decreases with increasing d and h as shown in Fig. 5(a).

 figure: Fig. 4.

Fig. 4. (a) The waveguide transmittance T as a function of the graphene Fermi level EF and the cladding thickness h at fixed core thickness d = 50 nm. (b) The amplitude of tNG (solid) and tB (dashed) as a function of EF at different h; d = 50 nm. (c) The phase difference between tNG and tB as a function of EF at different h; d = 50 nm. (d) The waveguide transmittance T as a function of EF and core thickness d for when h = 30 nm. (e,f) Same as in (b,c) for different d; h = 30 nm. Dotted line in (c,f) indicates the destructive interference condition ${\phi _{\textrm{NG}}} - {\phi _\textrm{B}} = \pi $. The gap width g = 5 nm in all cases.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. (a) The maximum modulation depth as a function of ${d}$ and ${h}$. (b) Transmittance as a function of ${{E}_\textrm{F}}$ for different carrier mobility ${\mu}$ of graphene. Geometric parameters are h = 30 nm, d = 50 nm, and g = 5 nm. Inset: transmittance in “on” and “off” states as a function of ${\mu}$.

Download Full Size | PDF

In graphene-based active optical devices, the carrier mobility of graphene is one of the key factors that limits the device performance. The carrier mobility is inversely proportional to the non-radiative loss of the devices, and is thus desired to be as high as possible. The experimentally reported mobility of graphene reaches up to a few millions of cm2/Vs if the sample is prepared by mechanical exfoliation and measured at a low carrier concentration under cryogenic conditions [26]. However, large-scale optical device applications necessitate CVD-grown graphene with high carrier concentration (n > 5×1012 cm-2) operating at room temperature, which tends to have a significantly deteriorated carrier mobility ranging from several hundred to a few thousand cm2/Vs [27,28]. To characterize the tolerance on the carrier mobility of our device, we calculate the transmittance for the various graphene carrier mobility ${\mu}$ spanning 500 to 50,000 cm2/Vs. As shown in Fig. 5(b), the switching behavior of the nanogap device is remarkably stable under carrier mobility variation, and one can see a noticeable degradation when the mobility drops under 1000 cm2/Vs. We attribute this to the operating mechanism of the device, which employs the impedance modulation of the graphene-covered nanogap cavity mode rather than the propagation of graphene plasmons that are highly sensitive to the ohmic loss.

5. Discussion on quantum and non-local effects

Since the length scale of electromagnetic field confinement in the proposed device is comparable to the electron wavelength of its constituent plasmonic materials, a discussion of quantum and non-local effects that may influence the local classical electromagnetic analysis is deemed necessary. Let us first discuss the quantum tunneling and quantization effect. First, the probability of quantum tunneling across a few-nm air gap should be negligible, because the threshold length scale associated with electron tunneling is < 1 nm for a 1 eV work function difference, which is sufficiently smaller than the gap width of our device [29,30]. Second, it is well known that the optical properties of nanostructures, including graphene plasmonic nanoresonators, can be significantly altered by electronic energy quantization when the size of the system is comparable to the electron wavelength (typically ∼10 nm) [31]. We note that the metal on graphene leads to doping due to a difference in work function [32]. Hence, the graphene underneath can be modeled as doping variations across the graphene nanogap, which has minimal effect on electrical conduction due to Klein tunneling [33].

The nonlocal effect originates from the spatial spreading of electron wave function. It has been theoretically predicted that the plasmonic properties of sub-10 nm graphene nanoresonators deviate from that of the local classical electromagnetic model due to non-local effects [34]. We speculate that the non-local effect would also be present in our case, but its influence on the device operation would be smaller compared to the case of a physically patterned graphene plasmonic resonator for the following reasons: First, the operation mechanism of the proposed nanogap device does not rely on the plasmonic properties of graphene, and thus the in-plane spatial field variation in the graphene is almost negligible. Second, due to the conducting boundaries between the graphene and the Au top cladding, the current density distribution in the proposed device is much smoother compared to the graphene plasmonic resonators with abrupt physical terminations. From the perspective of the hydrodynamic model, the reduced gradient of the current density should result in a smaller hydrodynamic pressure for electrons and thus a weaker non-local effect [35]. Although numerical methods for quantitatively estimating the non-local effects have been under active development, it is still difficult to rigorously treat our system, which involves multiple materials and non-nontrivial geometries, using existing computational methods. Therefore, we defer more rigorous numerical investigations of the nanogap non-local effects, which we anticipate will lead to only quantitative corrections, for future studies.

6. Conclusion

In summary, we theoretically investigate an exotic transmission modulation behavior across a sub-10 nm MIM waveguide gap covered by graphene. Here, the interference between the background transmission and the transmission through the tunable nanogap mode enables large on-off transmission modulations of Ton/Toff > 100 with the active device length being λ0/1000, which is more than 10 times shorter than the wavelength of highly confined graphene plasmons. In addition, the device operation is robust under graphene carrier mobility variation, exhibiting a well-preserved switching behavior even at a low mobility of 1000 cm2/Vs. The ultra-compact form factor, high modulation efficiency, and robustness makes the proposed nanogap waveguide architecture a promising platform for future integrated-photonic systems.

Appendix 1: transmission mechanisms via the waveguide-integrated nanogap

We schematically illustrate the charge distribution of the plasmonic modes in the nanogap region in Fig. 6; both modes contribute to the waveguide transmission. The tightly confined nanogap mode (Fig. 6(a)) is a product of coupling between the graphene and MIM gap plasmons. On the other hand, the background transmission (Fig. 6(b)) is mediated by the continuum of the modes that exist in the gap region excluding the nanogap mode. The two main transmission channels – via the nanogap and via the background modes – have different dependencies on the Fermi level and can be separately calculated as described in Appendix 3.

 figure: Fig. 6.

Fig. 6. The charge distribution schematics of (a) the nanogap mode and (b) background transmission in the MIM waveguide-integrated plasmonic cavity.

Download Full Size | PDF

Appendix 2: derivation of the modified Fabry-Perot cavity model

We have developed a semi-analytical model to calculate the transmission via the Fermi level-dependent nanogap mode by summing up the infinite series of the partial transmissions occurring at every round trip of the nanogap plasmon as schematically shown in Fig. 7. The complex partial transmission for each iteration is given as:

$$\begin{array}{l} {t_\textrm{1}}\textrm{ = }{c_{\textrm{MN}}}{e^{i{\beta _{\textrm{NG}}}h}}{r_{\textrm{NF}}}{e^{i{\beta _{\textrm{NG}}}h}}{c_{\textrm{NM}}}\textrm{ = }{c_{\textrm{MN}}}{c_{\textrm{NM}}}{e^{i2{\beta _{\textrm{NG}}}h}}{r_{\textrm{NF}}}\; \\ {t_\textrm{2}}\textrm{ = }{c_{\textrm{MN}}}{e^{i{\beta _{\textrm{NG}}}h}}{r_{\textrm{NF}}}{e^{i{\beta _{\textrm{NG}}}h}}{r_{\textrm{NM}}}{e^{i{\beta _{\textrm{NG}}}h}}{r_{\textrm{NF}}}{e^{i{\beta _{\textrm{NG}}}h}}{c_{\textrm{NM}}}\textrm{ = }{t_\textrm{1}}\left( {{r_{\textrm{NM}}}{r_{\textrm{NF}}}{e^{2i{\beta _{\textrm{NG}}}h}}} \right)\\ \ldots \\ {t_\textrm{N}}\textrm{ = }{t_\textrm{1}}{\left( {{r_{\textrm{NM}}}{r_{\textrm{NF}}}{e^{i2{\beta _{\textrm{NG}}}h}}} \right)^{\textrm{N - 1}}}. \end{array}$$

 figure: Fig. 7.

Fig. 7. Conceptual illustration of the nanogap plasmon mode propagation and coupling.

Download Full Size | PDF

Therefore, the cumulative transmission via the nanogap mode tNG is

$${{{t}_{\textrm{NG}}} = \mathop \sum \limits_{\textrm{k = 1}}^\infty {{t}_\textrm{k}} = \frac{{{{t}_\textrm{1}}}}{{\textrm{1 - }{{r}_{\textrm{NM}}}{{r}_{\textrm{NF}}}{{e}^{i{{\beta }_{\textrm{NG}}}{h}}}}}\textrm{ = }\frac{{{{c}_{\textrm{MN}}}{{c}_{\textrm{NM}}}{e^{i{\beta _{\textrm{NG}}}h}}r_{\textrm{NF}}}}{{\textrm{1 - }{{r}_{\textrm{NM}}}{{r}_{\textrm{NF}}}{e^{i2{\beta _{\textrm{NG}}}h}}}}.} $$

Here, ${{\beta }_{\textrm{NG}}}$ is the vertical propagation constant of the nanogap plasmon. cij and rij are the coupling and reflection coefficients, respectively, for different interfaces where i, j = N (nanogap), M (MIM waveguide), and F (free space), and are obtained from full-wave simulations as described below.

Appendix 3: calculation of ${{c}_{\textrm{ij}}}$, ${{r}_{\textrm{ij}}}$ and ${{t}_\textrm{B}}$

The reflection coefficients (rij), coupling coefficients (cij), and the background transmission coefficient (tB) can be obtained by performing electromagnetic simulations of the structures illustrated in Fig. 8(a) and (b). Figure 8(a) is a schematic of a plasmonic waveguide T-junction connecting Au-SiO2-Au (ASA) and Au-Air-Au (AAA) waveguides. The cladding thickness of the AAA waveguide is set to 100 nm, which is sufficiently thicker than the skin depth of Au. Figure 8(b) shows a schematic of a terminated AAA waveguide with an open end.

 figure: Fig. 8.

Fig. 8. (a) T-junction, and (b) the open-ended MIM waveguides. Numbers indicate the incident and reflection ports, and colored arrows show the propagation direction of the waveguide mode.

Download Full Size | PDF

cMN, cNM, rNM, and tB are determined from the electric fields of the transmitted and reflected modes at each port in the T-junction: cMN = E5exp(inAAAk0L5)/E1exp(–inASAk0L1), cNM = E6exp(inASAk0L6)/E2exp(–inAAAk0L2), rNM = E5exp(inAAAk0L5)/E2exp(–inAAAk0L2), and tB = E6exp(inAAAk0L6)/E1exp(–inASAk0L1). Here, Ei is the complex amplitude of the transverse electric field component of TM0 mode measured at the port number i; n, k0, and Li are the effective index of the TM0 mode, free space wavevector, and distance between port i and the nanogap waveguide junction. In tB calculation, the Fermi level of graphene is assumed to be 0.01 eV to exclude graphene plasmon transmission. Coefficient rNF = E8exp(inAAAk0L8)/ E7exp(–inAAAk0L7) is calculated in the open-ended waveguide structure shown in Fig. 8(b).

Appendix 4: dependence of ${{c}_{\textrm{ij}}}$, ${{r}_{\textrm{ij}}}$and${{t}_\textrm{B}}$ on geometrical parameters

The dependence of the coupling, and reflection coefficients on the nanogap width g are presented in Fig. 9. The coupling coefficients cMN and cNM have identical amplitude with $\pi $ phase difference owing to the nature of the E-type T-junction. rNF stays near-unity and exhibits a marginal dependence on g due to the huge mode mismatch between the nanogap mode and the free space radiation. On the other hand, the amplitude of rNM decreases with a considerable amount of phase variation as the nanogap mode becomes less confined with increasing g.

 figure: Fig. 9.

Fig. 9. Amplitude and phase of (a) the coupling coefficient cNM and the reflection coefficients (b) rNF and (c) rNM as functions of the gap width g. (d) dependence of the effective index of nanogap mode nNG on g. The other device parameters are set as h = 30 nm, d = 50 nm, $\mu\; = \; 10,000\; \textrm{c}{\textrm{m}^\textrm{2}}\textrm{/Vs}$, and EF = 0.26 eV, respectively.

Download Full Size | PDF

Funding

National Research Foundation of Korea (2017R1E1A1A01074323, 2019K1A3A1A14064929, BK21 FOUR Program); Villum Fonden (16498); Syddansk Universitet; Det Frie Forskningsråd (0165-00051B).

Disclosures

The authors declare no conflicts of interest.

References

1. M. Jablan, H. Buljan, and M. Soljačić, “Plasmonics in graphene at infrared frequencies,” Phys. Rev. B 80(24), 245435 (2009). [CrossRef]  

2. T. Low, R. Roldán, H. Wang, F. Xia, P. Avouris, L. M. Moreno, and F. Guinea, “Plasmons and screening in monolayer and multilayer black phosphorus,” Phys. Rev. Lett. 113(10), 106802 (2014). [CrossRef]  

3. S. Kim, M. S. Jang, V. W. Brar, Y. Tolstova, K. W. Mauser, and H. A. Atwater, “Electronically tunable extraordinary optical transmission in graphene plasmonic ribbons coupled to subwavelength metallic slit arrays,” Nat. Commun. 7(1), 12323 (2016). [CrossRef]  

4. S. Kim, M. S. Jang, V. W. Brar, K. W. Mauser, L. Kim, and H. A. Atwater, “Electronically Tunable Perfect Absorption in Graphene,” Nano Lett. 18(2), 971–979 (2018). [CrossRef]  

5. H. Hu, X. Yang, F. Zhai, D. Hu, R. Liu, K. Liu, Z. Sun, and Q. Dai, “Far-field nanoscale infrared spectroscopy of vibrational fingerprints of molecules with graphene plasmons,” Nat. Commun. 7(1), 12334 (2016). [CrossRef]  

6. H. Hu, X. Yang, X. Guo, K. Khaliji, S. R. Biswas, F. J. García de Abajo, T. Low, Z. Sun, and Q. Dai, “Gas identification with graphene plasmons,” Nat. Commun. 10(1), 1131 (2019). [CrossRef]  

7. I. H. Lee, D. Yoo, P. Avouris, T. Low, and S. H. Oh, “Graphene acoustic plasmon resonator for ultrasensitive infrared spectroscopy,” Nat. Nanotechnol. 14(4), 313–319 (2019). [CrossRef]  

8. I. J. Luxmoore, P. Q. Liu, P. Li, J. Faist, and G. R. Nash, “Graphene-Metamaterial Photodetectors for Integrated Infrared Sensing,” ACS Photonics 3(6), 936–941 (2016). [CrossRef]  

9. T. Low, A. Chaves, J. D. Caldwell, A. Kumar, N. X. Fang, P. Avouris, T. F. Heinz, F. Guinea, L. Martin-Moreno, and F. Koppens, “Polaritons in layered two-dimensional materials,” Nat. Mater. 16(2), 182–194 (2017). [CrossRef]  

10. S. Kim, S. G. Menabde, V. W. Brar, and M. S. Jang, “Functional Mid-Infrared Polaritonics in van der Waals Crystals,” Adv. Opt. Mater. 8(5), 1901194 (2020). [CrossRef]  

11. D. A. Iranzo, S. Nanot, E. J. C. Dias, I. Epstein, C. Peng, D. K. Efetov, M. B. Lundeberg, R. Parret, J. Osmond, J. Y. Hong, J. Kong, D. R. Englund, N. M. R. Peres, and F. H. L. Koppens, “Probing the ultimate plasmon confinement limits with a van der Waals heterostructure,” Science 360(6386), 291–295 (2018). [CrossRef]  

12. E. H. Hwang and S. Das Sarma, “Plasmon modes of spatially separated double-layer graphene,” Phys. Rev. B 80(20), 205405 (2009). [CrossRef]  

13. I. H. Lee, M. He, X. Zhang, Y. Luo, S. Liu, J. H. Edgar, K. Wang, P. Avouris, T. Low, J. D. Caldwell, and S. H. Oh, “Image polaritons in boron nitride for extreme polariton confinement with low losses,” Nat. Commun. 11(1), 3649 (2020). [CrossRef]  

14. I. H. Lee, L. Martin-Moreno, D. A. Mohr, K. Khaliji, T. Low, and S. H. Oh, “Anisotropic Acoustic Plasmons in Black Phosphorus,” ACS Photonics 5(6), 2208–2216 (2018). [CrossRef]  

15. I. Datta, S. H. Chae, G. R. Bhatt, M. A. Tadayon, B. Li, Y. Yu, C. Park, J. Park, L. Cao, D. N. Basov, J. Hone, and M. Lipson, “Low-loss composite photonic platform based on 2D semiconductor monolayers,” Nat. Photonics 14(4), 256–262 (2020). [CrossRef]  

16. V. Sorianello, M. Midrio, G. Contestabile, I. Asselberghs, J. Van Campenhout, C. Huyghebaert, I. Goykhman, A. K. Ott, A. C. Ferrari, and M. Romagnoli, “Graphene-silicon phase modulators with gigahertz bandwidth,” Nat. Photonics 12(1), 40–44 (2018). [CrossRef]  

17. M. Liu, X. Yin, E. Ulin-Avila, B. Geng, T. Zentgraf, L. Ju, F. Wang, and X. Zhang, “A graphene-based broadband optical modulator,” Nature 474(7349), 64–67 (2011). [CrossRef]  

18. M. S. Jang, S. Kim, V. W. Brar, S. G. Menabde, and H. A. Atwater, “Modulated Resonant Transmission of Graphene Plasmons Across a λ /50 Plasmonic Waveguide Gap,” Phys. Rev. Appl. 10(5), 054053 (2018). [CrossRef]  

19. M. Maier, A. Nemilentsau, T. Low, and M. Luskin, “Ultracompact Amplitude Modulator by Coupling Hyperbolic Polaritons over a Graphene-Covered Gap,” ACS Photonics 5(2), 544–551 (2018). [CrossRef]  

20. G. E. Moore, “Cramming more components onto integrated circuits,” IEEE Solid-State Circuits Soc. Newsl. 11(3), 33–35 (2006). [CrossRef]  

21. E. D. Palik, Handbook of Optical Constants of Solids (Academic, 1998).

22. C. H. Park, F. Giustino, M. L. Cohen, and S. G. Louie, “Velocity renormalization and carrier lifetime in graphene from the electron-phonon interaction,” Phys. Rev. Lett. 99(8), 086804 (2007). [CrossRef]  

23. H. Yan, T. Low, W. Zhu, Y. Wu, M. Freitag, X. Li, F. Guinea, P. Avouris, and F. Xia, “Damping pathways of mid-infrared plasmons in graphene nanostructures,” Nat. Photonics 7(5), 394–399 (2013). [CrossRef]  

24. L. A. Falkovsky and A. A. Varlamov, “Space-Time Dispersion of Graphene Conductivity,” Eur. Phys. J. B 56(4), 281–284 (2007). [CrossRef]  

25. B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics, Wiley Series in Pure and Applied Optics (Wiley, 2007).

26. L. Wang, I. Meric, P. Y. Huang, Q. Gao, Y. Gao, H. Tran, T. Taniguchi, K. Watanabe, L. M. Campos, D. A. Muller, J. Guo, P. Kim, J. Hone, K. L. Shepard, and C. R. Dean, “One-dimensional electrical contact to a two-dimensional material,” Science 342(6158), 614–617 (2013). [CrossRef]  

27. M. C. Sherrott, P. W. C. Hon, K. T. Fountaine, J. C. Garcia, S. M. Ponti, V. W. Brar, L. A. Sweatlock, and H. A. Atwater, “Experimental Demonstration of >230° Phase Modulation in Gate-Tunable Graphene-Gold Reconfigurable Mid-Infrared Metasurfaces,” Nano Lett. 17(5), 3027–3034 (2017). [CrossRef]  

28. W. S. Leong, H. Wang, J. Yeo, F. J. Martin-Martinez, A. Zubair, P. C. Shen, Y. Mao, T. Palacios, M. J. Buehler, J. Y. Hong, and J. Kong, “Paraffin-enabled graphene transfer,” Nat. Commun. 10(1), 867 (2019). [CrossRef]  

29. K. J. Savage, M. M. Hawkeye, R. Esteban, A. G. Borisov, J. Aizpurua, and J. J. Baumberg, “Revealing the quantum regime in tunnelling plasmonics,” Nature 491(7425), 574–577 (2012). [CrossRef]  

30. K. J. Lee, S. Kim, W. Hong, H. Park, M. S. Jang, K. Yu, and S. Y. Choi, “Observation of Wavelength-Dependent Quantum Plasmon Tunneling with Varying the Thickness of Graphene Spacer,” Sci. Rep. 9(1), 1199 (2019). [CrossRef]  

31. S. Thongrattanasiri, A. Manjavacas, and F. J. García de Abajo, “Quantum finite-size effects in graphene plasmons,” ACS Nano 6(2), 1766–1775 (2012). [CrossRef]  

32. G. Giovannetti, P. A. Khomyakov, G. Brocks, V. M. Karpan, J. Van Den Brink, and P. J. Kelly, “Doping graphene with metal contacts,” Phys. Rev. Lett. 101(2), 026803 (2008). [CrossRef]  

33. T. Low, V. Perebeinos, J. Tersoff, and P. Avouris, “Deformation and scattering in graphene over substrate steps,” Phys. Rev. Lett. 108(9), 096601 (2012). [CrossRef]  

34. A. Fallahi, T. Low, M. Tamagnone, and J. Perruisseau-Carrier, “Nonlocal electromagnetic response of graphene nanostructures,” Phys. Rev. B 91(12), 121405 (2015). [CrossRef]  

35. C. Ciracì, R. T. Hill, J. J. Mock, Y. Urzhumov, A. I. Fernández-Domínguez, S. A. Maier, J. B. Pendry, A. Chilkoti, and D. R. Smith, “Probing the ultimate limits of plasmonic enhancement,” Science 337(6098), 1072–1074 (2012). [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 (9)

Fig. 1.
Fig. 1. (a) Schematic of the MIM waveguide-integrated plasmonic modulator. Each MIM waveguide supports the TM0 guided mode whose transmission is modulated by the gate voltage. For the carrier mobility of graphene $\mu\; = \; 10,000\; \textrm{c}{\textrm{m}^\textrm{2}}\textrm{/Vs}$ and a fixed ${{\lambda }_\textrm{0}} = \; 7.5\; \mathrm{\mu}\textrm{m}$ , the optimized geometric parameters are: core thickness d = 50 nm, cladding thickness h = 30 nm, and gap width g = 5 nm. (b) The transmittance T map as functions of Fermi level EF and the geometric parameter g. (c) The distribution of the normalized Re{Ex} inside the each nanocavity geometries at minimum transmittance when EF = 0.24 eV (left), and EF = 0.26 (right).
Fig. 2.
Fig. 2. (a) Illustration of the nanogap coupling and transmission mechanism. The nanogap transmission is decomposed into independent terms of mode coupling c and reflection r, where the gap-confined MIM mode has the propagation constant βNG. For a device geometry characterized by d = 50 nm, h = 30 nm, g = 5 nm, we present: (b) amplitude (red) and phase (blue) of the coupling cNM = −cMN (dashed) and reflection rNM (solid) coefficients; (c) transmission through the nanogap mode (tNG, dashed) and background transmission (tB, solid); (d) Net transmission amplitude (|tNG + tB|) obtained from the analytical model (dashed) and full-wave simulations (solid). In all simulations, graphene carrier mobility $\mu{\; = \; 10,000\; }{\textrm{cm}^\textrm{2}}\textrm{/Vs}$ .
Fig. 3.
Fig. 3. (a) Amplitude of the transmission coefficients of the nanogap mode, tNG, and the background mode, tB, as a function of Fermi level EF and gap width g. (b) Phase difference of tNG and tB, ${\phi _{\textrm{NG}}} - {\phi _\textrm{B}}$ , as a function of EF and g. Black dotted line indicates the destructive interference condition when ${\phi _{\textrm{NG}}} - {\phi _\textrm{B}} = \pi $ . Here, d = 50 nm and h = 30 nm.
Fig. 4.
Fig. 4. (a) The waveguide transmittance T as a function of the graphene Fermi level EF and the cladding thickness h at fixed core thickness d = 50 nm. (b) The amplitude of tNG (solid) and tB (dashed) as a function of EF at different h; d = 50 nm. (c) The phase difference between tNG and tB as a function of EF at different h; d = 50 nm. (d) The waveguide transmittance T as a function of EF and core thickness d for when h = 30 nm. (e,f) Same as in (b,c) for different d; h = 30 nm. Dotted line in (c,f) indicates the destructive interference condition ${\phi _{\textrm{NG}}} - {\phi _\textrm{B}} = \pi $ . The gap width g = 5 nm in all cases.
Fig. 5.
Fig. 5. (a) The maximum modulation depth as a function of ${d}$ and ${h}$ . (b) Transmittance as a function of ${{E}_\textrm{F}}$ for different carrier mobility ${\mu}$ of graphene. Geometric parameters are h = 30 nm, d = 50 nm, and g = 5 nm. Inset: transmittance in “on” and “off” states as a function of ${\mu}$ .
Fig. 6.
Fig. 6. The charge distribution schematics of (a) the nanogap mode and (b) background transmission in the MIM waveguide-integrated plasmonic cavity.
Fig. 7.
Fig. 7. Conceptual illustration of the nanogap plasmon mode propagation and coupling.
Fig. 8.
Fig. 8. (a) T-junction, and (b) the open-ended MIM waveguides. Numbers indicate the incident and reflection ports, and colored arrows show the propagation direction of the waveguide mode.
Fig. 9.
Fig. 9. Amplitude and phase of (a) the coupling coefficient cNM and the reflection coefficients (b) rNF and (c) rNM as functions of the gap width g. (d) dependence of the effective index of nanogap mode nNG on g. The other device parameters are set as h = 30 nm, d = 50 nm, $\mu\; = \; 10,000\; \textrm{c}{\textrm{m}^\textrm{2}}\textrm{/Vs}$ , and EF = 0.26 eV, respectively.

Equations (3)

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

t NG  =  c MN c NM e i β NG h r NF 1 -  r NM r NF e i 2 β NG h
t 1  =  c MN e i β NG h r NF e i β NG h c NM  =  c MN c NM e i 2 β NG h r NF t 2  =  c MN e i β NG h r NF e i β NG h r NM e i β NG h r NF e i β NG h c NM  =  t 1 ( r NM r NF e 2 i β NG h ) t N  =  t 1 ( r NM r NF e i 2 β NG h ) N - 1 .
t NG = k = 1 t k = t 1 1 -  r NM r NF e i β NG h  =  c MN c NM e i β NG h r NF 1 -  r NM r NF e i 2 β NG h .
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.