1 Introduction
The remarkable growth of silicon semiconductor industry over the last several decades is driven by Moore's law and the dimensional scaling of on-chip components, which yields consistent improvement in the binary information throughput of microprocessor chips. However, quantum mechanical leakage and energy concerns will be prohibitive to further miniaturisation of devices, thereby slowing down the momentum of Moore's law [1–3]. While transistor scaling improves their performance, the performance of interconnects degrades with scaling [4]. As such, interconnects impose limits on the overall throughput and efficiency of microprocessors. Several potential solutions including the use of ballistic low-dimensional conductors (carbon nanotubes and graphene), less aggressive wire width and thickness scaling, and low-swing signalling have been proposed to overcome on-chip interconnect challenges [5–7]. At the same time, system-level performance is limited by the bandwidth of off-chip interconnects. To address system-level interconnect challenges, three-dimensional (3D) integration and airgap interconnects are being investigated [8, 9]. As the frequency of operation increases, electrical interconnects also suffer from signal distortion resulting from interconnect dispersion, which sets an upper bound on the maximum achievable signal bandwidth [10]. Going forward, terahertz (THz) band communication is envisioned as a key wireless technology to handle bandwidth-greedy applications and delay-critical communication needs for next-generation systems [11, 12].
In this paper, we propose to exploit plasma waves in multilayer (ML) graphene heterostructures for on-chip signalling purposes in the THz band. Graphene, which is an atomically thin sheet of carbon atoms arranged in a honeycomb lattice, displays strong light–matter interaction over a broad frequency range [13–16]. Graphene can be easily combined with other 2D materials such as transition metal dichalcogenides, phosphorene, and hexagonal boron nitride, to develop a fully integrated solution for the generation, detection, modulation, and propagation of plasma waves [14, 17]. At the lower end of the THz frequency spectrum (<10 THz), a sheet of graphene sandwiched between two dielectrics supports the propagation of plasma waves that are polarised in the transverse magnetic (TM) direction [18]. However, a graphene-based parallel-plate waveguide (PPWG) structure is able to support quasi-transverse electromagnetic (TEM) waves in the THz band [19, 20]. These plasma waves can be guided along the waveguide for on-chip local and global communication. The propagation characteristics of graphene plasmons can be configured through electrical or chemical doping. This provides opportunities for reconfigurable plasmonics – something that is not feasible with conventional metal plasmonics [21–24]. For electrostatically tuning the characteristics of plasma waves, a gate voltage is applied through a metal–oxide–semiconductor setup in PPWG geometry.
To quantify the performance of plasmonic interconnects, we analytically model their energy-per-bit and bandwidth density. The energy dissipation includes the energy required for generation, detection, and modulation of the plasmonic signal. We focus on the thermal- and shot-noise limited transmission of plasmons. To derive the bandwidth density, we consider the propagation of a narrowband Gaussian signal centred at THz frequency for both single waveguide (SWG) and PPWG interconnects. The distortion in the pulse width and amplitude reduction due to graphene ohmic losses is obtained by the convolution of the channel impulse response and the input envelope signal. At THz frequencies, only the intra-band dynamical conductivity of ML graphene is considered, while inter-band contribution can be neglected. The conductivity model of graphene includes the effect of carrier scatterings due to acoustic phonons and charged impurities. The effect of finite electrostatic screening between the multiple layers in ML graphene stack is also included. Lossy propagation resulting from energy-dependent carrier scatterings in graphene is studied for both TM and TEM modes as a function of operating frequency and material parameters of ML graphene stack. Owing to their nearly dispersion-less propagation, PPWG plasmonic interconnects offer significantly higher bandwidth density as compared to their electrical counterparts at the 2020 ITRS technology node [25]. Therefore, PPWG interconnects are more suitable for global level or even off-chip signalling purposes. On the other hand, TM plasmons propagating in SWG interconnects suffer much less degradation due to carrier scatterings. As such, they can be used for energy-efficient signalling at the local and semi-global length scales on the chip.
2 Guided modes in plasma wave interconnects
The device setup for the propagation of TM-polarised plasma waves consists of an ungated graphene sheet within a dielectric referred to as SWG geometry. To support quasi-TEM modes, we focus on a gated graphene-based PPWG geometry. The two waveguide structures are shown in Fig. 1. In the case of SWG, the electric field has components in both z- and x-directions. However, in the case of PPWG, the electric field is predominantly in the z-direction that is perpendicular to the graphene sheet. In both setups, x-direction is defined along the length of the graphene sheet and corresponds to the direction of propagation of plasma waves. In general, the dielectric and the substrate have different dielectric permittivities. However, in this paper we consider that ungated graphene is embedded in a linear, homogeneous dielectric media, which simplifies analytical calculations. In both device setups, graphene is modelled as a stack of multiple layers with a linear energy-dispersion relationship. In ML graphene grown epitaxially on silicon carbide or via chemical vapour deposition, it is shown that layers in graphene are decoupled. As such, Bernal stacking no longer exists and the layers are rotated relative to each other. Essentially, the decoupling of layers preserves the linear energy dispersion in ML graphene [26, 27].
[IMAGE OMITTED. SEE PDF]
To obtain the plasma wave dispersion relationship for the device setups in Fig. 1, we solve Maxwell's equations and apply impedance boundary condition on graphene–dielectric interface. This is mathematically expressed as . Here, is the complex dynamical conductivity of ML graphene, is the tangential electric field at the graphene interface, and (i = 1, 2) is the magnetic field on either side of the graphene sheet. The TM and quasi-TEM dispersion relationships, under long-wavelength approximation, are given as [28]
2.1 ML graphene conductivity
To study the mode characteristics of plasma waves in graphene, it is important to model the dynamical conductivity of the graphene sheet. In this work, we ignore non-local effects, related to spatial dispersion, in graphene to model the conductivity. Instead we focus on the random-phase approximation in long-wavelength limit for which . Here, is the Fermi velocity of carriers in graphene. If we define the effective plasmon mode index as , to ensure the validity of local conductivity model, . This condition is typically satisfied for both TM and quasi-TEM modes in graphene at THz frequencies. Mathematically, the frequency-dependent intra-band conductivity of monolayer graphene is given as [29]
2.2 Carrier relaxation rate
We consider elastic momentum relaxation processes, including acoustic phonons and charged impurities, in graphene to model the carrier relaxation rate. In graphene, scattering due to longitudinal acoustic phonon modes dominates as coupling of electron–phonon states for other phonon modes is too weak or the energy scales of the optical phonon modes are too high for the frequency range of interest. Carrier scattering time due to acoustic phonons is given as
The scattering time due to long-range Coulomb potential of charged impurities for large doping () is given as [32]
The net relaxation rate is computed using the Mattheissen's sum rule. That is, . As shown in Fig. 2, carrier relaxation rate due to charged impurities decreases as the carrier concentration increases. This is because carriers screen electric field lines, which increases the scattering time. On the other hand, carrier relaxation rate due to acoustic phonons increases with an increase in Fermi level. For typical values of Fermi level () and charged impurity concentration (), the net carrier relaxation rate is around 1–3 THz. Due to electrostatic screening between the layers in ML graphene, the carrier relaxation rate for each layer will be different and will be evaluated accordingly. For all analyses in this paper, and , unless otherwise noted.
[IMAGE OMITTED. SEE PDF]
Using the model of carrier relaxation rate, we obtain the dynamical conductivity of graphene as shown in Fig. 3 for various values of and and . An increase in increases the conductivity as more layers tend to contribute to current conduction. This fact is also evidenced by the increase in the number of effective layers. For lower value of , , but this ratio saturates to unity for , where in ML graphene. Both real and imaginary parts of graphene conductivity increase with an increase in at all values of . While has a non-monotonic dependence on , the factor in (6b) increases with an increase in Fermi level. This leads to a monotonic increase in with .
[IMAGE OMITTED. SEE PDF]
2.3 Electrostatic doping
In the case of PPWG geometry, electrostatic gating can be used to adjust the Fermi level defined in (7). In the absence of interface and substrate traps, the Fermi level in the graphene sheet as a function of gate voltage () is given as [24]
[IMAGE OMITTED. SEE PDF]
The complex conductivity of ML graphene as a function of gate voltage evaluated at a frequency of 1 THz is shown in Fig. 5. There exists a minimum conductivity around the Dirac point due to the formation of electron–hole puddles. The imaginary part of conductivity increases more rapidly than the real part as gate voltage increases. The real part of conductivity is related to in-phase current which produces Joule heating, while the imaginary part represents the out-of-phase current or the inductive response of electrons. It can be seen from Fig. 5 that inductive effects become more important at higher end of the THz frequencies. In fact, in the low-loss scenario (), per-unit-length resistance () and kinetic inductance () of the graphene sheet supporting plasmons can be defined as
[IMAGE OMITTED. SEE PDF]
3 Performance modelling
Using the conductivity model derived in prior sections, we obtain the characteristics of propagating plasmon modes in ML graphene using (1b). The advantage of graphene plasmons is that they have smaller mode areas, i.e. higher value of compared to noble metals. In Fig. 6, we plot the wave vector of both TM and TEM plasmons in graphene as a function of frequency. Material parameters chosen for this analysis are indicated in the figure caption. The real part of the wave vector is related to the plasmon wavelength, group velocity, and wave-front velocity, while its imaginary part determines the propagation length of plasmons owing to finite ohmic losses in graphene. Propagation length is defined as the length scale at which the plasmon field amplitude falls to 1/e of its initial value.
[IMAGE OMITTED. SEE PDF]
The most obvious difference in the propagation characteristics of TM and TEM plasmons is related to their frequency dependence. While the real part of the wave vector of TM plasmons displays behaviour, TEM plasmons have a nearly dispersion-less propagation. That is, increases linearly with frequency for TEM modes. Unfortunately, at the lower end of the THz spectrum, TEM plasmons display high imaginary part of the wave vector. This means that TEM plasmon modes are extremely lossy and may not be suitable for information propagation for low THz frequencies.
To compare the performance of TM and TEM plasmon modes, we consider two metrics: propagation length () and localisation length (LL). Mathematically
[IMAGE OMITTED. SEE PDF]
We also define a figure-of-merit (FOM) as the ratio , which indicates the number of wavelengths () that the plasma wave could travel on the waveguide before losing most of its energy. The FOM incorporates the impact of Fermi level and other material parameters on the performance of the waveguide (Fig. 8). The FOM improves for TEM waves as a function of frequency. This means that TEM plasmons are more advantageous for information propagation at higher frequencies. On the other hand, the FOM corresponding to TM plasmons exhibits a non-monotonic dependence on frequency. In general, TM modes are preferred at the lower end of the THz spectrum for frequencies on the order of a few 100s of GHz. Beyond a frequency of 1–2 THz, the FOM of TEM modes exceeds that of TM modes. If the FOM is less than unity, it indicates that the plasmon modes are too lossy and unsuitable for waveguiding application. For the chosen material parameters, TM modes always have a FOM higher than unity. However, TEM modes display huge ohmic losses for frequency <1 THz. We also note the distinct behaviour of TM and TEM modes on the Fermi level in the graphene sheet. The FOM of TEM modes degrade with an increase in Fermi level at all frequencies, while the FOM of TM modes improve with an increase in Fermi level for frequencies <7 THz. For higher frequencies, the FOM of TM modes has the same dependence on Fermi level as that of TEM modes. TEM modes tend to be more confined for lower Fermi level and their confinement degrades rapidly with an increase in Fermi level.
[IMAGE OMITTED. SEE PDF]
3.1 Energy-per-bit
The propagation length plays an important role in determining the energy dissipation of plasmonic interconnects as it limits the maximum distance between transceivers communicating using plasma waves. The energy dissipation of the plasmonic interconnect is given as [36, 37]
A lower bound on the energy dissipation of plasmonic interconnects can be obtained by considering the transmission of plasmons to be thermal- and shot-noise limited. That is, the bit-error rate (BER) requirement supersedes the requirement for the received optical power and extinction ratio. In this case, the energy dissipation for a plasmonic interconnect of length L is given as [47]
The product as a function of the modulation depth for various values of the detector capacitance is shown in Fig. 9. As modulation depth increases, the mean number of plasmons needed for detection reduces. In the high modulation depth regime, the detector capacitance significantly impacts the value of mean number of plasmons. The specific choice of does not significantly impact the value of as shown in the inset plot of Fig. 9.
[IMAGE OMITTED. SEE PDF]
A lower bound on the energy dissipation of the modulator is given as [37, 48, 49]
The energy dissipation of SWG and PPWG interconnects of length 10 μm is shown in Fig. 10 as a function of frequency of operation. At the lower end of the THz spectrum, PPWG interconnects consume approximately three orders of magnitude higher energy compared to SWG interconnects. At shorter interconnect length scales, the overall energy consumption of SWG plasmonic interconnects tends to be dominated by the modulator energy dissipation; therefore, for a fixed value of , the energy dissipation of SWG interconnects becomes nearly independent of frequency. The excessive energy dissipation of PPWG interconnects is a result of the short propagation length, which could limit the use of PPWG interconnects for energy-efficient signalling.
[IMAGE OMITTED. SEE PDF]
3.2 Bandwidth density
The bandwidth of a single channel in nanophotonic interconnects is typically limited by the operation speed of the transceiver circuits at the ends of the interconnect. However, in this paper, we will obtain an upper bound on the bandwidth of plasmonic interconnects by considering only the interconnect-dominated limits. To evaluate the bandwidth of a plasmonic interconnect, we consider the propagation of a narrow-band Gaussian signal centred at the THz frequency through SWG and PPWG geometries. As the Gaussian pulse propagates, it suffers from attenuation due to ohmic losses in graphene and the pulse shape distorts due to waveguide dispersion. The amplitude of the Gaussian signal, , at any location z on the waveguide measured at time t can be obtained by the convolution of the channel impulse response and the input signal. Details can be found in prior works [50, 51]. Mathematically, is given as
[IMAGE OMITTED. SEE PDF]
Due to pulse shape distortion, successive pulses launched into the waveguide may overlap if inter-pulse timing is not enough. The maximum achievable signal bandwidth or bitrate, , for plasmonic interconnects is given as
[IMAGE OMITTED. SEE PDF]
Fig. 13 shows the bandwidth density versus frequency of both SWG and PPWG plasmonic interconnects. While the bandwidth density of SWG interconnects supporting TM plasmon modes is nearly independent of frequency, for PPWG interconnects supporting TEM modes bandwidth density increases with an increase in frequency. Due to the dispersion-less propagation of TEM modes, the bandwidth density of PPWG interconnects is significantly higher than that of SWG interconnects. At a frequency of 2 THz, the bandwidth density of PPWG interconnects is for a channel length of 100 μm. However, for the same parameters, the bandwidth density of SWG interconnects is 117 Gbit/s/μm. As such, PPWG interconnects are more appropriate for global or off-chip signalling purposes, where higher bandwidth density is desirable.
[IMAGE OMITTED. SEE PDF]
4 Comparison with electrical interconnects
The performance of plasmonic and electrical interconnects is compared with respect to energy-per-bit and bandwidth density. We also quantify interconnect length scales for which plasmonic transmission is more advantageous compared to electrical transmission on the chip.
4.1 Energy dissipation
The energy-per-bit of plasmonic interconnects is discussed in Section 3.1. For electrical interconnects, the energy dissipation is determined by the charging and discharging of interconnect and transistor capacitances. As such, the energy-per-bit of electrical interconnects can be stated as
In Fig. 14, the energy-per-bit of electrical and plasmonic interconnects is shown as a function of interconnect length. For plasmonic interconnects, the frequency of operation is selected as 1 THz to achieve a reasonable trade-off between energy-per-bit and bandwidth density. While the energy dissipation of electrical interconnects increases linearly with interconnect length, the energy dissipation of plasmonic interconnects grows exponentially with interconnect length. If the energy dissipation associated with modulation is not included, SWG interconnects are more energy efficient compared to electrical interconnects up to a few millimetre length scale. For an energy dissipation of 5 fJ/bit of the modulator, SWG interconnects have lower energy than electrical interconnects only up to 100 μm as indicated by point ‘a’ in the figure. PPWG plasmonic interconnects have a lower propagation length; therefore, their energy dissipation increases much more rapidly with interconnect length. As such, PPWG plasmonic interconnects have lower energy dissipation compared to electrical interconnects only up to 20 μm when as shown by point ‘b’ in the figure. In the presence of modulator energy dissipation, PPWG plasmonic interconnects consume more energy than their electrical counterparts for lengths >1 μm. For energy-efficient communication using plasmonic interconnects, it is better to exploit TM plasmons in SWG interconnect geometries owing to the larger propagation length. Further, for competitive advantage against electrical interconnects, it is important to bring the energy dissipation of modulators to below 1 fJ/bit.
[IMAGE OMITTED. SEE PDF]
4.2 Bandwidth density
For digital electrical interconnects in which signals are sent as ‘on’ and ‘off’, the bitrate capacity depends on the rise time of the signal. The rise time of the electrical interconnects depends on whether the interconnect is operating in the RC or LC regime. In general, the bitrate of electrical interconnects with a cross-sectional area of and length L is given as
[IMAGE OMITTED. SEE PDF]
4.3 Partition length
We obtain a partition length, corresponding to energy dissipation of plasmonic and electrical interconnects. is the interconnect length beyond which plasmonic interconnects consume more energy than their electrical counterparts. Therefore, a higher value of is preferred for energy-efficient plasmonic signalling. As shown in Table 1, modulator energy dissipation strongly affects the value of for both TEM and TM modes. In general, for a fixed value of , for TM modes is higher than for TEM modes. When dominates the total energy dissipation in the plasmonic domain, it is found that TEM modes will always consume more energy than their electrical counterparts even for the shortest length scale. This is indicated using ‘ ’ in the table for TEM modes when . Interestingly, TM modes becomes more energy efficient at longer length scales when . This is contrary to the behaviour observed when dominates the overall energy dissipation. Our results indicate that SWG plasmonic interconnects are more suited for energy-efficient signalling purposes for intermediate and semi-global length scales on the chip.
Table 1 Partition length, , corresponding to energy dissipation
TEM modes | TM modes | |||
0 | 13.8 μm | 16.5 μm | 1.63 mm | 5.4 mm |
1 fJ/bit | 9.4 μm | 11.4 μm | 19.2 μm | |
5 fJ/bit | >100 μm |
We also quantify an interconnect length scale, , beyond which the bandwidth density of plasmonic interconnects exceeds that of electrical interconnects. A lower value of indicates superior performance of plasmonic interconnects. Results are given in Table 2. The wiring pitch of electrical interconnects operating in the RC regime is chosen between the minimum and maximum allowed values for global interconnects per the 2020 ITRS technology roadmap. As the results show, TEM modes offer very high bandwidth density and are more advantageous than electrical interconnects for length scales as short as a few micrometres. By up-sizing the wiring pitch of electrical interconnects, their bandwidth density improves, which means values would increase. The dependence of on is different for TM and TEM modes. While an increase in reduces the value of for TM modes, the corresponding value for TEM modes increases. This is due to the different dispersive features of TM and TEM modes as discussed in Section 3.
Table 2 Partition length, , corresponding to bandwidth density
TEM modes | TM modes | |||
Pitch | ||||
32 nm | 0.21 μm | 0.31 μm | 13.7 μm | 6.0 μm |
200 nm | 4.8 μm | 7.1 μm | 308.0 μm | 135.2 μm |
2000 nm | 80.9 μm | 118.6 μm | 5.2 mm | 2.27 mm |
5 Conclusion
In this paper, graphene-based plasmonic interconnects are investigated for low-energy and high-bandwidth on-chip signalling purposes for next-generation systems. Two geometries of the plasmonic interconnects, namely SWG and PPWG, are investigated. SWG interconnects support the propagation of TM modes, while PPWG interconnects support nearly dispersion-less TEM modes. The energy dissipation and bandwidth density of both modes is analytically modelled as a function of material (Fermi level, scattering rate, electrostatic screening) and geometrical parameters of the waveguides. It is shown that due to their parallel-plate structure, PPWG interconnects confine plasmons better yielding a superior localisation length. The improvement in localisation reduces the propagation length significantly for PPWG interconnects. The energy dissipation of plasmonic interconnects is modelled by incorporating the energy consumed by source, detection, and modulation circuitry within the regime of noise-limited plasmon transmission. We show that when the source/detect circuit is the dominant component of energy dissipation, SWG interconnects have lower energy dissipation compared to their electrical counterparts up to length scales of few millimetres. This energy advantage drops rapidly with growing modulator energy dissipation. On the other hand, PPWG interconnects have limited propagation length. Therefore, their energy advantage ceases beyond length scales of . To model bandwidth density, we focus on the propagation of a narrowband Gaussian signal centred at THz frequency through both waveguide geometries. Our results clearly show that PPWG interconnects exceed the bandwidth density of RC-limited electrical interconnects for interconnects as short as few micrometres at the 2020 ITRS technology node. Hence, PPWG interconnects are more suitable for global chip-level signalling, while SWG interconnects can achieve low-energy operation at the local and semi-global length scales.
Cavin, R.K., Lugli, P., Zhirnov, V.V.: ‘Science and engineering beyond moore's law’, Proc. IEEE, 2012, 100, pp. 1720–1749 (Special Centennial Issue)
Cumming, D.R.S., Furber, S.B., Paul, D.J.: ‘Beyond moore's law’, Philos. Trans. R. Soc. Lond. A, Math. Phys. Eng. Sci., 2014, 372, (2012), p. 20130376
Markov, I.L.: ‘Limits on fundamental limits to computation’, Nature, 2014, 512, (7513), pp. 147–154
Meindl, J.D., Chen, Q., Davis, J.A.: ‘Limits on silicon nanoelectronics for terascale integration’, Science, 2001, 293, (5537), pp. 2044–2049
Chen, J.: ‘Self‐calibrating on‐chip interconnects’. PhD thesis, Stanford University, 2012
Ceyhan, A., Naeemi, A.: ‘Cu interconnect limitations and opportunities for swnt interconnects at the end of the roadmap’, IEEE Trans. Electron Devices, 2013, 60, (1), pp. 374–382
Prasad, D., Ceyhan, A., Pan, C. et al.: ‘Adapting interconnect technology to multigate transistors for optimum performance’, IEEE Trans. Electron Devices, 2015, 62, (12), pp. 3938–3944
Kumar, V., Bashirullah, R., Naeemi, A.: ‘Modeling, optimization and benchmarking of chip‐to‐chip electrical interconnects with low loss air‐clad dielectrics’. IEEE 61st Electronic Components and Technology Conf. (ECTC), 2011, 2011, pp. 2084–2090
Kumar, V.: ‘Modeling and optimization approaches for benchmarking emerging on‐chip and off‐chip interconnect technologies’. PhD thesis, Georgia Institute of Technology, 2014
Miller, D.A.B., Ozaktas, H.M.: ‘Limit to the bit‐rate capacity of electrical interconnects from the aspect ratio of the system architecture’, J. Parallel Distrib. Comput., 1997, 41, (1), pp. 42–52
Jornet, J.M., Akyildiz, I.F.: ‘Fundamentals of electromagnetic nanonetworks in the terahertz band’, Found. Trends Netw., 2013, 7, (2‐3), pp. 77–233, ISSN 1554‐057X, doi: 10.1561/1300000045, available at [DOI: https://dx.doi.org/10.1561/1300000045] , 2004
Rakheja, S.: ‘On the Gaussian pulse propagation through multilayer graphene plasmonic waveguides – impact of electrostatic screening and frequency dispersion on group velocity and pulse distortion’, IEEE Trans. Nanotechnol., 2016, 15, (6), pp. 936–946
Ceyhan, A., Naeemi, A.: ‘Overview of the interconnect problem’, in Todri‐Sanial, Aida, Dijon, Jean, Maffuci, Antonio (Ed.): ‘Carbon nanotubes for interconnects’ (Springer, Switzerland, 2017), pp. 3–36
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2018. This work is published under http://creativecommons.org/licenses/by-nc-nd/3.0/ (the "License"). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Graphene‐based heterostructures provide a viable platform to implement optoelectronic devices that can operate in the terahertz (THz) band. In this study, the authors focus on multilayer (ML) graphene as the building block to implement high‐frequency and low‐energy plasmonic interconnects for on‐chip signalling in next‐generation systems. Two specific plasmonic interconnect geometries are analysed: single waveguide (SWG) and parallel‐plate waveguide (PPWG). While SWG interconnects support propagating surface plasmons that are polarised in the transverse magnetic direction, in PPWG interconnects, nearly dispersion‐less quasi‐transverse electromagnetic modes are supported. The dispersion characteristics are derived by solving Maxwell's equations in the device setup in which ML graphene presents an impedance boundary condition. The effects of number of layers, electrostatic screening, and Fermi level are included in the model of intra‐band dynamical surface conductivity of ML graphene. The authors also develop analytical models of energy‐per‐bit and bandwidth density for both SWG and PPWG interconnects. The energy dissipation includes the effect of plasmon generation, detection, and modulation circuitry within a thermal‐ and shot‐noise‐limited transmission of information. They quantify optimal interconnect length scales for which plasmonic interconnects provide lower energy and higher bandwidth when compared against their electrical (copper/low‐κ) counterparts at the 2020 ITRS technology node.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details
1 Electrical and Computer Engineering, New York University, Brooklyn, NY, USA