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

Exact dispersion relations for the hybrid plasmon-phonon modes in graphene on dielectric substrates with polar optical phonons

Open Access Open Access

Abstract

Intrinsic optical phonons and extrinsic polar optical phonons (POPs) strongly affect the graphene surface plasmons. Specifically, extraneous POPs present on the surface of an underlying substrate change the behavior of the graphene's surface plasmons sharply due to the plasmon-phonon hybridization. Here, we report modeling of exact dispersion relations for graphene's surface plasmons affected by intrinsic optical phonons and extrinsic POPs of the surface of polar dielectric substrates with one or more vibrational frequencies. In doing so, we have employed random phase approximation with modified two-dimensional polarizability (2D-Π0). The adapted Π0 addresses limitations of the previously derived plasmons dispersion, obtained using classical two-dimensional polarizability. We show the new model overcomes the unsatisfying behavior of the plasmonic dispersion relation obtained by the classical 2D-Π0 at high-wavenumbers and its inability to indicate the starting point of the mode damping. Our new simple model eliminates the complexity of the other presented models in describing the surface plasmons’ behavior, specifically at high wavenumbers. Besides, we use our dispersion model to learn about the plasmon content of the hybrid modes, which is a vital value to compute output current in plasmonic graphene-based devices. The coupled-mode lifetime due to the hybrid nature depends on both plasmon and phonon lifetimes. We capture this value here. There is an excellent agreement between our theoretical results and the experimental data reported earlier. They pave the way for the exact modeling of graphene plasmons on common polar substrates and bring in the closeness of the theoretical approaches and experimental results.

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

1. Introduction

Carrier transport is strongly affected by electron-phonon interactions [14], as inevitable sources of scattering, leading to sharp changes in the behavior of electronic systems, such as plasmons — i.e., collective oscillations of the conduction electrons. Over the years, graphene has attracted ample theoretical and experimental attention due to its unique physical and thermal properties, making it suitable for various applications [57]. In particular, graphene plasmons possess noticeable advantages [811], such as long-lived excitations confining enhanced electric fields in tiny volumes and the ability to control the plasmon peak’s frequency through an electrostatic gating mechanism.

Some experimental [1217] and theoretical [1822] works have focused on the hybridization of plasmon and phonon modes in the past decade. In general, two different types of phonon can affect plasmons (intrinsic and extrinsic). Lattice vibrations and defects in graphene are two inherent scattering sources in the absence of extrinsic phonons, dominating carrier transport and mobility. Intervalley [23] and intravalley acoustic [17,18] and optical [2022] phonon scatterings are the three types of intrinsic phonon scattering mechanisms. Intervalley scattering, induced by high momentum and high energy phonons, becomes significant at high temperatures. The so-called D Raman peak from Raman scattering data for graphene, which is highly sensitive to the fabrication process and experimental measurements setup, shows intervalley scattering is an impurity- and defect-assisted process [2326]. Under particular conditions, however, it is possible to reduce the deteriorating effects of intervalley scattering on plasmons [26]. Low-energy acoustic phonons induce quasielastic intravalley phonon scattering. This mechanism becomes significant at low temperatures where high-energy intravalley optical phonons are absent. The effects of the longitudinal acoustic phonons are more pronounced than the other modes due to their stronger coupling with plasmons [17,19]. For working frequencies higher than those of the acoustic phonons, their deteriorating effect on plasmons is negligible. Although the energy of the intravalley optical phonons is not low enough to be considered a scattering channel, they bring plasmons of higher frequencies into the intraband single-particle excitation (SPEintra) continuum via optical phonon emission (PE) [12,16].

Besides the intrinsic optical phonons, polar optical phonons (POPs) on the surface of substrates, such as SiC [12,13], hBN [14,15], and SiO2 [16], have shown undeniable effects on the graphene surface plasmons. From the first studies on the impact of surface polar optical phonons (S-POPs) on graphene transport [18], compared to the remote phonon scattering in two-dimensional (2D) electron gas in channels of Si metal-oxide-semiconductor field-effect transistors (MOSFETs) [27,28], theoretical and experimental efforts have been made with enthusiasm. When the surface plasmons of graphene, placed on a polar substrate, are excited at frequencies about the S-POPs’, they are strongly affected by the electric dipoles induced via vibrations of the electrically positive and negative atoms on the surface of the substrate. Investigations show that S-POPs can change plasmon dispersion dramatically, especially near the phonon frequencies [1516].

Although the measured results on variations of dispersion in the presence of intrinsic and extrinsic phonons for graphene on nonpolar and polar substrates are available experimentally, the lack of precise theoretical dispersion relations for the plasmonic modes and hybrid plasmon-phonon modes are apparent. So, we seek the exact dispersion relations of the coupled modes in the structured graphene placed on a polar substrate, with one, two, or three vibrational frequencies. Moreover, the existing theoretical predictions [12,29] show graphene surface plasmons fall within the interband single-particle excitation continuum and start to damp (i.e., decay into electron-hole pairs) at a critical wavevector (qcrit), where the plasmon radian frequency exceeds the cutting radian frequency of ωcrit = −vFqcrit + 2μC/ħ. Here, ℏ represents the reduced Planck’s constant, and vF≈106 m·s−1 and μC are graphene Fermi velocity and chemical potential. Nonetheless, based on the experimental data [16], the damping process starts at a smaller critical wavenumber for frequencies higher than the frequency of intrinsic optical phonon. This work aims to address this discrepancy by refining the 2D polarizability, Π0. The new Π0 overcomes the two existing limitations: (i) inability to indication of the starting point for the mode damping and (ii) the complexity of the previous relations for graphene dispersion at high wavenumbers, q [12,29]. Our simple dispersion models bring in two imperative results. That is a method of calculating the plasmonic content in a coupled mode and the corresponding mode lifetime. In most graphene plasmonic-based optoelectronic devices, there is an essential need to measure or obtain the output current [30], experimentally or theoretically. When phonons vibrations are absent, the device output current depends on the entire plasmons content [31]. Otherwise, the scenario for calculating the device output differs [30]. In other words, in calculating the output device current in the presence of plasmons-phonons interactions, knowledge of pure plasmonic content surviving the interactions is vital. We will show later, in the next section, by knowing an exact dispersion relation, one can obtain the pure plasmonic content. To the best of our knowledge, such a dispersion relation is not available to date.

In this paper, we model a precise dispersion relation for the plasmon-phonon modes in graphene placed on nonpolar and polar dielectric substrates having one, two, or three phonons vibrational frequencies, based on the existing absorption experimental data [12,16,30], enabling us to predict the precise output currents of devices like plasmon-enhanced photodetectors, solar cells, and modulators. Moreover, another vital parameter is the coupled-mode lifetime, also captured in this study.

Different research groups have employed various methods like first-principle calculations [2,3], Boltzmann equation approach [20], and effective-mass approximation [32] to investigate the impact of the interaction between graphene intrinsic optical phonons and electrons on its surface plasmons dispersion. The same methods are also applicable to the interactions of the extrinsic phonons (S-POPs) with graphene electrons. Here, we use random phase approximation (RPA) instead.

Through the electromagnetic theory, we know surface plasmons propagate in a medium with a near-zero dielectric constant. Hence, the first step to reach our goal is to obtain the graphene dielectric function in the presence of phonons. In doing so, we need to determine the effective electron-electron potentials mediated by both the intrinsic optical phonons and S-POPs, first. Then, employ these potential values in the RPA expansion of the dielectric function and, finally, equate it to zero. With some algebraic manipulations, we achieve the exact and complete relation for the coupled modes dispersion, for the first time, to the best of our knowledge. In this investigation, we ignore the impact of possible defects in graphene. Nevertheless, we model the edge scattering effect by a broadening factor.

The rest of the paper is organized as follows. In the second section, we review the scheme for describing the electron-electron potentials mediated by intrinsic optical phonons and the extrinsic S-POPs. Then, after a brief introduction of the electron interactions in a many-body system, including phonons and the RPA dielectric expansion, we calculate the desired dispersion function. Besides, we obtain the plasmons lifetime in the single-particle excitation and phonon emission regions while introducing the phonons lifetime. Then, we discuss the results obtained, including the modified dispersion relation for graphene in the absence of extrinsic phonons, beyond 2D classical one, the exact hybrid dispersion relations for graphene on polar substrates with up to three phonons, plasmon content of the coupled modes, and their lifetime. At last, we close the paper with conclusive remarks in the final section.

2. Theory

2.1 Intrinsic optical phonons

The lattice vibrations, responsible for optical phonons in graphene, change the separation between neighboring carbon atoms, modifying the overlap integral between them. Using perturbation theory, let us write the electron-electron potential mediated by optical phonons in the form of [32]

$$V_{\textrm{el - el}}^{\textrm{op}} = \frac{1}{{2A}}\sum\limits_{{{\mathbf k}_1}{\sigma _1}} {\sum\limits_{{{\mathbf k}_2}{\sigma _2}} {\sum\limits_{{\mathbf q}\lambda } {\frac{1}{A}{{|{g_{{\mathbf q}\lambda }^{\textrm{op}}} |}^2}{{\cal G}}_\lambda ^0(\omega )\hat{c}_{{{\mathbf k}_1} + {\mathbf q},{\sigma _1}}^\dagger \hat{c}_{{{\mathbf k}_2} + {\mathbf q},{\sigma _2}}^\dagger {{\hat{c}}_{{{\mathbf k}_2}{\sigma _2}}}{{\hat{c}}_{{{\mathbf k}_1}{\sigma _1}}}} } } \equiv \frac{1}{{2{A^2}}}\sum\limits_{{\mathbf k},\sigma ,{\mathbf q},\lambda } {v_\lambda ^{\textrm{op}}\hat{c}_{{\mathbf k} + {\mathbf q},\sigma }^\dagger {{\hat{c}}_{{\mathbf k}\sigma }}} ,$$
where A is the area of graphene, k and q are the wave vectors of the electrons and phonons,σ represents the spin of the electron, λ denotes phonon identifier, ${g_{{\mathbf q}\lambda }^{\textrm{op}}}$ is the transition matrix elements related to electron-optical phonon interaction, ${{\cal G}}_\lambda ^0$ introduces the phonon Green’s function, and $\hat{c}_{{\mathbf k} + {\mathbf q},\sigma }^\dagger$ and ${{\hat{c}}_{{\mathbf k},\sigma }}$ are the creation and annihilation operators, respectively. For intrinsic optical phonons, the identifier λ is omitted from ${{|{g_{{\mathbf q}\lambda }^{\textrm{op}}} |}^2}$ and ${{\cal G}}_\lambda ^0$ as in [20,32]
$${|{g_{\mathbf q}^{\textrm{op}}} |^2} = \frac{{\hbar D_{\textrm{op}}^2}}{{2{\rho _m}{\omega _{\textrm{op}}}}},$$
and
$${{\cal G}}_{}^0 = \frac{{2{{{\omega _{\textrm{op}}}} / \hbar }}}{{{{({\omega + {i / {{\tau_{\textrm{op}}}}}} )}^2} - \omega _{\textrm{op}}^2}}.$$

Dop ≈ 11 eV·Å−1 is the optical phonon deformation potential, ρm ≈ 7.6×10−7 kg·m−2 is graphene mass density [20], ħωop = 198 meV and τop ≈ 70 fs [16] are the intrinsic optical phonon energy and lifetime, and ω represents the input optical signal frequency. From Eq. (1), one can interpret similarity to the simple electron-electron Coulomb interaction [33] with coupling strength of $(1/A){|{g_{\mathbf q}^{\textrm{op}}} |^2}{{{\cal G}}^0}$. Moreover, we have

$${v^{\textrm{op}}} = {|{g_{\mathbf q}^{\textrm{op}}} |^2}{{{\cal G}}^0}(\omega ),$$
which is the potential between two electrons mediated by an intrinsic optical phonon in graphene, with a definite wave vector.

2.2 Extrinsic surface optical phonons

Due to the optical nature of the surface phonons, an equation similar to Eq. (1) can be used for S-POPs, except for different transition matrix elements and green’s function, as in [20]

$$v_\lambda ^{\textrm{POP}} = \sum\limits_\lambda {{{|{g_{{\mathbf q}\lambda }^{\textrm{POP}}} |}^2}{{\cal G}}_\lambda ^0(\omega ).}$$
in which
$${|{g_{{\mathbf q}\lambda }^{\textrm{POP}}} |^2} = \frac{{\pi {e^2}}}{{{\varepsilon _0}}}{{\cal F}}_\lambda ^2\frac{{\textrm{exp}({ - 2q{z_0}} )}}{q} \approx \frac{{\pi {e^2}}}{{q{\varepsilon _0}}}{{\cal F}}_\lambda ^2,$$
and ${{\cal G}}_\lambda ^0$ is similar to Eq. (2b), in which ωop and τop are replaced by the extrinsic phonons frequency, ωλ, and lifetime, τPOP. In Eq. (5), e is the elementary charge, ε0 is the free-space permittivity, and z0 35 Å is the graphene-substrate van der Waals separation [20]. Notice, since qz0≪1, the term exp(−2qz0) in Eq. (5) becomes almost one. Fröhlich coupling strength, ${{\cal F}}_\lambda ^2$, describes the magnitude of the polarization field of the phonon labeled λ [33]. For substrates with only one dominant phonon (λ = 1), such as SiC, we have
$${{\cal F}}_1^2 = \frac{{\hbar {\omega _1}}}{{2\pi }}\left( {\frac{1}{{\varepsilon_\infty^{} + 1}} - \frac{1}{{\varepsilon_0^{} + 1}}} \right).$$

For those with more than one dominant phonon mode, each phonon mode is associated with one Fröhlich coupling strength — e.g., for AlN, Al2O3, h-BN, and HfO2 with λ= 1 and 2, we have,

$${{\cal F}}_1^2 = \frac{{\hbar {\omega _1}}}{{2\pi }}\left( {\frac{1}{{{\varepsilon_{\textrm{i,1}}} + 1}} - \frac{1}{{{\varepsilon_0} + 1}}} \right)$$
and
$${{\cal F}}_2^2 = \frac{{\hbar {\omega _2}}}{{2\pi }}\left( {\frac{1}{{\varepsilon_\infty^{} + 1}} - \frac{1}{{\varepsilon_{\textrm{i,1}}^{} + 1}}} \right),$$
while for substrates like SiO2 with (λ = 1, 2, and 3), we have
$${{\cal F}}_1^2 = \frac{{\hbar {\omega _1}}}{{2\pi }}\left( {\frac{1}{{\varepsilon_{\textrm{i,1}}^{} + 1}} - \frac{1}{{\varepsilon_0^{} + 1}}} \right),$$
$${{\cal F}}_2^2 = \frac{{\hbar {\omega _2}}}{{2\pi }}\left( {\frac{1}{{\varepsilon_{\textrm{i,2}}^{} + 1}} - \frac{1}{{\varepsilon_{\textrm{i,}1}^{} + 1}}} \right),$$
and
$${{\cal F}}_3^2 = \frac{{\hbar {\omega _3}}}{{2\pi }}\left( {\frac{1}{{\varepsilon_\infty^{} + 1}} - \frac{1}{{\varepsilon_{\textrm{i,2}}^{} + 1}}} \right).$$

In Eqs. (7) and (8) ω1 < ωλ≥2, ε, εi,1(2), and ε0 are the optical, intermediate, and static permittivity of the substrate, εi,1 represents the permittivity for the frequency range of ω1 < ω < ω2, and εi,2 is related to the range ω2 < ω < ω3. The values of the above parameters for different substrates are given in Table 1.

Tables Icon

Table 1. Various permittivities, the dominant phonon energies, and related Fröhlich coupling constants for dielectrics substrates SiC, Al2O3, HfO2, AlN, ZrO2, ZrSiO4, hexagonal BN (hBN), and SiO2

2.3 Dielectric function

The RPA is a powerful tool for finding plasmons dispersion relation, starting with the graphene dielectric function. In this approach, a single electron is exposed to a self-consistent field, including the externally applied electric field plus the one induced by other electrons. For RPA to result in an excellent approximation, we must pay attention to the Wigner-Seitz parameter (rs) [29], which describes the density of an interacting system. For a small value of rs (e.g., ≈ 0.5 for graphene, which is independent of electron concentration [29]), the system interacts weakly, and RPA becomes an excellent approach. Furthermore, changing electron concentration in graphene does not affect the accuracy of the RPA method.

Consider an electron, in the absence of an external field, moves through a weakly interacting system, going from the state k1 to state k2, influenced by two dominant potentials: (i) a Coulomb potential induced by other electrons (VC = e2/20) [16], and (ii) the potential between electrons mediated by phonons (Vph = vop + vλPOP). Feynman diagrams [35] (Table 2) show all possible routes through which the state of an electron can change from k1 to k2. As observed from the table, an electron may propagate freely going from k1 to k2, provided k2 = k1. Moreover, the freely moving electron may undergo a scattering process with a phonon of polarizability Π0 and then continue to move freely again before the second scattering occurs. The latter process may occur many times. Hence, the total possibility is the sum of all possible routes, which may be summarized in an infinite series

$$\Uparrow = \uparrow + \begin{array}{c} \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \end{array} + \begin{array}{c} \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \end{array} + \begin{array}{c} \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \end{array} + \cdots = \uparrow + \left\{ { \uparrow + \begin{array}{c} \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \end{array} + \begin{array}{c} \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \uparrow \end{array} + \cdots } \right\} \times \begin{array}{c} \uparrow \\ {\boxed{{{\Pi _0}}}}\\ {} \end{array} = \uparrow + \begin{array}{c} \uparrow \\ {\boxed{{{\Pi _0}}}}\\ \Uparrow \end{array}$$

For simplicity, we omitted the indices of the diagrams shown in Table 2 from Eq. (9). Here we consider optical phonons as the dominant scatterer neglecting other scatterers, like acoustic phonons, defects, and ionized impurities. In a weakly interacting system, the probability of successive scattering incidences equals the possibilities of all individual scattering incidences multiplied by each other. So, we can simplify Eq. (9) into [22]

$${V_{\textrm{eff}}} = ({{V_\textrm{C}} + {V_{\textrm{ph}}}} )+ {V_{\textrm{eff}}} \times {\Pi _0} \times ({{V_\textrm{C}} + {V_{\textrm{ph}}}} ),$$
where Veff represents the upward double arrow and (VC + Vph) signifies the upward single arrow. Hence, we can write the effective dynamical potential as
$${V_{\textrm{eff}}} = {{({{V_C} + {V_{\textrm{ph}}}} )} / {[{1 - ({{V_C} + {V_{\textrm{ph}}}} ){\Pi _0}} ]}}.$$

Moreover, the dielectric function from RPA, ε(q, ω), can be expressed by the effective potential [22,35],

$${V_{\textrm{eff}}} = {{({{V_\textrm{C}} + {V_{\textrm{ph}}}} )} / {\varepsilon ({q,\omega } )}}.$$

Tables Icon

Table 2. Feynman diagrams and their descriptions

By comparing Eq. (11) and Eq. (12), and using VC, Eq. (3), and Eq. (4) we get

$$\varepsilon ({q,\omega } )= {\varepsilon _{\textrm{env}}} - \left( {\frac{{{e^2}}}{{2q{\varepsilon_0}}} + {{|{g_{\mathbf q}^{\textrm{op}}} |}^2}{{{\cal G}}^0}(\omega )} \right.\left. { + \sum\limits_\lambda {{{|{g_{{\mathbf q}\lambda }^{\textrm{POP}}} |}^2}{{\cal G}}_\lambda^0(\omega )} } \right){\Pi _0},$$
where ε env is the average of the dielectric constants of the upper and lower media surrounding the graphene. In our model, we consider air as that the upper-medium.

As mentioned in the Introduction, we need to modify the existing formula for Π0 used in Ref. [8] to remove the limitations in predicting a precise starting point for the decaying continuum and correct behavior for surface plasmons at high q values to resemble that obtained experimentally [12,15,16].

Comparing the experimental data in Refs. [12,16,24] with classical 2D plasmon dispersion given [8], we see for ℏω > 198 meV, the measured plasmons peak’s frequency is blue-shifted compared to that of the 2D classical model. A further observation from this comparison is the discrepancy in the variations of dispersion for high q values —- i.e., ω $\propto$ q1/2 for classical dispersion and ω $\propto$ q3/2 for the experimental dispersion [12]. To eliminate these deviations, we have added a q-dependent term, γ, in the denominator of the classical Π0, as in

$${\Pi _0} \approx \frac{{{\mu _\textrm{C}}{q^2}}}{{\pi {\hbar ^2}({{{({\omega + \,{i / {{\tau_{\textrm{pl}}}}}} )}^2} - \gamma_{}^2} )}},$$
wherein. μC = ±ℏvF|kF| with kF representing the carriers’ Fermi velocity and wavenumber, and τpl is the plasmons lifetime, which we will introduce in the following subsection. As an appropriate choice for the dependencies of γ on the plasmons wavenumber to eliminate the deviation in the model from the experimental data, we suggest
$${\gamma ^2} = \frac{1}{8}\nu _\textrm{F}^2{q^2}{[{\exp ({{{q^{\prime}} / q}} )- 1} ]^{ - 1}},$$
where
$$q^{\prime} = \frac{{2\pi {\varepsilon _0}{\varepsilon _{\textrm{env}}}}}{{{e^2}{\mu _\textrm{C}}}}{({\hbar {\omega_{\textrm{op}}}} )^2}.$$
is the plasmon’s wavenumber at the frequency ω = ωop obtained from classical 2D graphene dispersion [8]. For qq’, [exp(q'/q)−1]−1→0 and γ→0, and hence Π0 approaches the classical one. As q increases, Π0 and dispersion relation start deviating from those obtained by the classical model. Nonetheless, for large wavenumbers, where ℏω ≫ ℏωop, a Taylor series expansion of the exponential term in Eq. (14b) results in γ $\propto$ q3/2. Using this and substituting Eq. (14) back into Eq. (13) and setting ε(q, ω) = 0 and ignoring the effects of the optical phonons, we see that ω $\propto$ q3/2 at high qs which is consistent with the experimental data [12,15,16]. Hence Eq. (14) is valid for all ranges of plasmons wavenumber and energy.

2.4 Lifetime calculations

2.4.1 Plasmon lifetimes

Plasmons are composed of an ensemble of free electrons. Hence, their lifetimes can be the same, τpl, helping us evaluate Π0 as in Eq. (14) and the plasmon-phonon coupled-mode lifetime, which we will calculate in the next section. Plasmons damp via different processes like scattering from background ionized impurities and the structured graphene edges, optical phonon emission [16], and electron-hole generation within the single-particle excitation region [29]. Hence, we can express τpl as a function of these processes characteristic times,

$$\tau _{\textrm{pl}}^{ - 1} = \tau _0^{ - 1} + \tau _{\textrm{PE}}^{ - 1} + \tau _{\textrm{SPE}}^{ - 1} + \tau _{\textrm{ed}}^{ - 1},$$
where τ≈ 85 fs comes from ionized impurity scattering in graphene [16]. The accuracy of graphene fabrication controls this characteristic time. The more well-constructed the sample is, the less scattering by defects and disorders will be. τPE is the lifetime of the plasmon due to the phonon emission. When plasmon energy exceeds that of the inherent optical phonon, ħωop = 198 meV, they damp through emitting optical phonons and entering the PE continuum [12,16]. This kind of scattering is an unavoidable process and is dominant for frequencies higher than ≈ 48 THz. It is highly dependent on temperature and becomes insignificant at low temperatures, where the number of inherent optical phonons is low, making the rate of the plasmons-phonons scattering negligible. As the temperature increases, the number of phonons increases, and the scattering rate becomes significant [17]. τSPE is the lifetime in the single-particle excitation continuum. It is due to interband transitions [29], which disrupt the balance of electron density with chemical potential μC. As long as the electron density at this level is low, this characteristic time is insignificant. Besides, τed represents the lifetime of electrons due to the edge effects. Scattering by the edges of the structure changes the momentum dramatically and causes high-momentum intervalley scatterings in graphene [2326]. Same as τ0, τed is also highly sensitive to the fabrication process and experimental measurements setup.

The existing theoretical model [29] shows plasmons start decaying via electron-hole generation within the intraband energy range of ħvFq – 2μC < ħω < ħvFq or within the interband transitions continuum. Nonetheless, the experimental results [12,16] suggest the plasmons decay for ħω ≥ ħωop ― i.e., damping by phonon emission and decaying into the intraband continuum before the interband transitions start. The lifetime of the plasmon in the PE continuum is [16]

$$\tau _{\textrm{PE}}^{ - 1} = \frac{{{g_{\textrm{PE}}}}}{\hbar }|{\hbar ({\omega + {\omega_\alpha }} )+ {\mu_\textrm{C}}} |\cdot [{{\mathbb{F}}({\omega ,{\omega_\alpha }} )+ 2} ],$$
where gPE ≈ 0.018,
$${\mathbb{F}}({\omega ,{\omega_\alpha }} )= \textrm{erf}\left( {\frac{{\hbar \omega - \hbar {\omega_\alpha }}}{{2{k_B}T}}} \right) + \textrm{erf}\left( {\frac{{ - \hbar \omega - \hbar {\omega_\alpha }}}{{2{k_B}T}}} \right),$$
represents the thermal broadening effect, with kB and T as the Boltzman constant environment temperature, and ωα=ωop. Moreover, τSPE consists of two components, one related to the intraband, τSPE,intra, and the other with the interband transitions, τSPE,inter
$$\tau _{\textrm{SPE}}^{ - 1} = \tau _{\textrm{SPE,intra}}^{ - 1} + \tau _{\textrm{SPE,inter}}^{ - 1}.$$
in which
$$\tau _{\textrm{SPE,intra}}^{ - 1} ={-} \frac{{{g_{\textrm{SPE,intra}}}}}{\hbar }|{\hbar ({\omega + {\omega_\alpha }} )- {\mu_C} - 3} |\cdot {\mathbb{F}}({\omega ,{\omega_\alpha }} ),$$
with gSPE,intra ≈ 0.009, ωα=vFq (i.e., the critical frequency at which the intraband transition starts), and
$$\tau _{\textrm{SPE,inter}}^{ - 1} = \frac{{{g_{\textrm{SPE,inter}}}}}{\hbar }|{\hbar ({\omega + {\omega_\alpha }} )+ {\mu_C}} |\cdot [{{\mathbb{F}}({\omega ,{\omega_\alpha }} )+ 2} ],$$
with gSPE,inter ≈ 0.02 and ωα = −vFq+2μC/ℏ (i.e., the critical frequency at which the interband transition starts).

In a periodically structured graphene with the periodicity d and filling factor F, as the physical width of graphene (i.e., w = Fd) shrinks edge scattering becomes significant, changing the momentum of the electrons notably. Hence, the scattered electrons decay via the intervalley phonon emission. The plasmons wavenumber in this periodic structure satisfies a formula like qpl ≈ π / (w – w0), where w0 ≈ 28 nm is the electrically inactive width of graphene [16]. We model τed, representing the lifetime of electrons in the presence of the edge effect, [16]

$${\tau _{\textrm{ed}}} \approx \frac{{2\pi F \times {{10}^{ - 6}}}}{q}\textrm{.}$$

2.4.2 Phonon lifetime

The edge scatterings, resulting in the emission of high-energy intervalley phonon scattering, can have a significant impact on the S-POPs’ lifetime, provided w < 90 nm (i.e., equivalent to q ≈ 4×107 m−1) [16]. Hence, the S-POPs’ lifetime is

$${\tau _{\textrm{POP}}} \approx {\tau _{\textrm{POP,0}}}\textrm{exp}\left( { - \frac{q}{{4 \times {{10}^7}}}} \right),$$
where τPOP,0≈ 200 fs is the lifetime of surface optical phonons in the absence of edge scattering [16,30]. We use Eq. (19) for calculating Green’s function (${{{\cal G}}^0}$) for the extrinsic phonons.

3. Results and discussion

3.1 Plasmon dispersion

To determine the plasmon dispersion, experimentally, the extinction spectrum of a superlattice of graphene is measured at different periodicities (i.e., various wavenumbers). Then, by locating the extinction peaks in the q-ω plane, one can find the dispersion graph [16]. Theoretically, one may obtain the extinction spectrum (i.e., the loss function) from the imaginary part of the inverse dielectric function,

$$L \approx{-} {\mathop{\rm Im}\nolimits} {\{{\varepsilon ({q,\omega } )} \}^{ - 1}}.$$

Although the loss function does not express an exact relation to obtaining the plasmons peaks’ frequencies, it offers valuable information through the graphical description of the q-ω plane, indicating where the incident light is lost by plasmon excitations.

3.1.1 Nonpolar substrates

Concerning the structured graphene placed on a nonpolar substrate (i.e., with no S-POPs), we obtain the plasmon dispersion relation by considering the new Π0 given in Eq. (14),

$$\omega = \sqrt {\frac{{{e^2}{\mu _\textrm{C}}}}{{2\pi {\hbar ^2}{\varepsilon _0}{\varepsilon _{\textrm{env}}}}}q + \gamma _{}^2} .$$

As mentioned in Section II, for high q values where ω ≫ ωop the dependence of ω and γ on q are similar. Using Eq. (20), we obtained the plasmonic mode loss spectrum in the q-ω plane as illustrated in Fig. 1, together with a color bar indicator. The solid white curve in this figure shows how the plasmons peaks’ frequency varies with the wavenumber (Eq. (21)). In these calculations, we considered a graphene layer with a chemical potential of μC = 0.3 eV [16] transferred on a diamond-like-carbon (DLC) substrate with a dielectric constant of εDLC = 5 (i.e., εenv = 3). Moreover, the green dots represent the classical 2D dispersion (i.e.,γ = 0), and the dashed lines indicate the boundaries of regions with allowed damping processes. The data obtained from Eq. (21) is in excellent agreement with those obtained experimentally shown in [16], within the entire q-ω plane, unlike the classical 2D dispersion. Further key features observed from this figure are: (i) As plasmons frequency reaches the optical phonons’ frequency (i.e., the horizontal white dashes), the plasmons start to damp into the PE continuum. (ii) While the plasmonic mode decaying by the phonon emission at high frequencies, the excitation of an interband single-particle enhances the damping process (i.e., within the SPEinter region). (iii) The plasmons’ peak frequency moves away from the SPEitnra region as q increases.

 figure: Fig. 1.

Fig. 1. Plasmons modes loss spectra (Eq. (20)) in graphene with μC = 0.3 eV on a DLC substrate. The white solid curve represents the corresponding peak frequency plasmons wavenumber (Eq. (21)). The dashed lines show the allowed damping boundaries. The green dots represent the classical 2D dispersion (i.e., γ = 0).

Download Full Size | PDF

3.1.2 Polar substrates

The S-POPs of a polar substrate with λ ( = 1, 2, or 3) dominant vibrating modes split the graphene surface plasmons into λ+1 branches of coupled plasmonic modes. To find the exact dispersion relation for every allowed coupled mode, we set Eq. (13) zero, and after some algebraic manipulations, obtain the intended dispersion relations for graphene on a given polar substrate. For a substrate with a single S-POP, Eq. (21) reduces to

$$\begin{aligned} \omega _{1,2}^{\textrm{pl - ph}}({q,{\mu_\textrm{C}}} )&= \Omega \left\{ {1 + \frac{1}{{2{\Omega ^2}}}({\omega_1^2 + {\gamma^2}} )} \right. \mp \left[ {1 + \frac{1}{{4{\Omega ^4}}}{{({\omega_1^2 - {\gamma^2}} )}^2}} \right.\\ &{\left. {{{\left. { + \frac{1}{{{\Omega ^2}}}\left( {\frac{{8\pi {{\cal F}}_1^2}}{\hbar }{\omega_1} - ({\omega_1^2 - {\gamma^2}} )} \right)} \right]}^{1/2}}} \right\}^{1/2}}\end{aligned}$$
where ω1pl-ph and ω2pl-ph(>ω1pl-ph) represent the peak frequencies of the two coupled plasmon-phonon modes and
$${\Omega ^2} = \frac{{q{\mu _\textrm{C}}{e^2}}}{{4\pi {\varepsilon _0}{\varepsilon _{\textrm{env}}}{\hbar ^2}}},$$
which is a function of q and μC.

We substituted Eq. (13) in Eq. (20) and solved it to obtain the loss spectra for the hybrid plasmon-phonon modes on the surface of graphene with μC = 0.57 eV [12] placed on the surface of a SiC substrate with only one dominant S-POPs of energy ħω1 = 116 meV and the dielectric constant of εSiC = 6.5 in the q-ω plane, as illustrated in Fig. 2.

 figure: Fig. 2.

Fig. 2. Hybrid modes loss spectra (Eq. (20)) in graphene with μC = 0.57 eV on a SiC substrate with a single POP-vibrating mode. The two white solid curves represent the corresponding peak frequencies versus the plasmons wavenumber (Eq. (22)). Inset depicts a zoomed-in view of the plots about S-POPs energy (ħω1C ≈ 0.2). The dashed lines show the allowed damping boundaries.

Download Full Size | PDF

The two solid white curves shown in this figure depict the dependencies of the peak frequencies of the hybrid modes on the plasmons wave number (Eq. (22)). The white dashes here represent the boundaries of regions with allowed damping processes. Comparison of these results with similar experimental results presented in Ref. [12] confirms excellent agreements between the two. The inset shows a zoomed-in view of the plots around the splitting energy (ℏω1 / μC ≈ 0.2), assisting visualization. We can easily observe that far from this energy, where S-POPs have negligible effects on the plasmonic modes, and hence the coupled modes peak frequencies (Eq. (22)) and plasmonic modes (Eq. (21)) coincide, showing solely electrons participate in the excitations. As $\hbar \omega _{1,2}^{\textrm{pl - ph}} \to \hbar {\omega _1}$ the plasmon-phonon hybridization prevails. Another distinct key feature observed in this figure is around the S-POPs excitation energy. The lower mode (i.e., the vibrational mode) starts saturating within the SPEintra continuum, and the higher one (i.e., the electronic mode) starts moving away from the SPEintra region. Further observations indicate that as q increases, the weight of the mode intensity transfers from the lower mode to the higher one and enhances until it reaches the onset of the PE continuum, beyond which it damps similar to what has been observed experimentally [12,16]. Moreover, as the plasmon's energy increases beyond that of the graphene intrinsic phonon (i.e., the horizontal white dashes), in addition to damping, the higher mode broadens. When the plasmons wavenumber and energy enter the SPEinter continuum, the corresponding mode damps more rapidly because of single-particle excitations in addition to the phonon emission.

When graphene is on any polar substrate with two fundamental S-POPs modes, we can obtain the peak frequencies for all three hybrid coupled plasmon-phonon modes by equating Eq. (13) to zero for λ = 1, 2,

$$\omega _{1,2}^{\textrm{pl - ph}} = {\textrm{Re}} \sqrt {\alpha - \frac{1}{2}({\beta + {\chi / \beta }} )\pm i\frac{{\sqrt 3 }}{2}({\beta - {\chi / \beta }} )} ,$$
and
$$\omega _3^{\textrm{pl - ph}} = {\textrm{Re}} \sqrt {\alpha + ({\beta + {\chi / \beta }} )} ,$$
where ω1pl-ph < ω2pl-ph < ω3pl-ph and α, β, and χ are as given in Appendix A.

Solving Eq. (20) we obtain the loss spectra for the three coupled plasmon-phonon modes on graphene with μC = 0.3 eV, as illustrated in Fig. 3. As can be seen from this figure, the behavior of the three hybrid coupled plasmon-phonon modes in the allowed damping continua is similar to the behavior of the hybrid modes shown in Fig. 2. For ħωs far away from ħω1, 2, where S-POPs become ineffective, the modes become purely plasmonics. Nonetheless, as ħωħω1, 2, the dispersions change dramatically, as observed in the inset. Moreover, as q increases gradually from zero, the weight of the mode intensity moves from the first mode to the second and then from the second to the third mode, and enhances until it reaches the onset of the PE continuum, beyond which it broadens and damps into SPEintra faster than before.

 figure: Fig. 3.

Fig. 3. Hybrid modes loss spectra (Eq. (20)) in graphene with μC = 0.3 eV on an AlN substrate with two S-POPs vibrating modes. The three solid white curves represent the traces of the corresponding peak frequencies (Eq. (23)). Inset depicts a zoomed-in view of the plots around the two S-POPs excitation energies (ħω1,2C ≈ 0.28 and 0.35). The dashed lines show the allowed damping boundaries.

Download Full Size | PDF

Next, we obtain the peak frequencies for hybrid coupled plasmon-phonon modes induced on the surface of graphene placed on a polar substrate with three S-POPs. In doing so, we equate Eq. (13) to zero for λ = 1, 2, and 3, and after some algebraic manipulations get the formula for the peak frequencies,

$$\omega _{1,2}^{\textrm{pl - ph}} = {\textrm{Re}} \sqrt {\Delta - {{{\Phi ^{{1 / 2}}}({1 \pm {{{\Gamma ^{1/2}}} / {{\Phi ^{3/4}}}}} )} / {6{\Lambda ^{{1 / 6}}}}}} ,$$
$$\omega _{3,4}^{\textrm{pl - ph}} = {\textrm{Re}} \sqrt {\Delta - \Phi {{^{{1 / 2}}({ - 1 \pm {{{{\rm K}^{1/2}}} / {{\Phi ^{3/4}}}}} )} / {6{\Lambda ^{{1 / 6}}}}}} .$$

Parameters Δ, Φ, Γ, Κ, and Λ in Eq. (24) are introduced in Appendix B.

Now, by substituting Eq. (13) into Eq. (20), presuming the polar substrate is SiO2 with three dominant phonons at the energies ħω1, 2, and 3= 100, 134, and 148.8 meV and the dielectric constant of εSiO2 = 2.4, we obtain the loss spectra for the hybrid coupled plasmon-phonon modes on the graphene with μC = 0.5 eV [16]. Figure 4 shows the four dispersive hybrid modes in the q-ω plane. The four solid white curves represent the corresponding modes’ peak frequencies. The results plotted here are in excellent agreement with the experimental data presented in [16]. Moreover, this figure displays all key features mentioned for polar substrates with one and two surface vibrational excitations.

 figure: Fig. 4.

Fig. 4. Hybrid modes loss spectra (Eq. (20)) in graphene with μC = 0.43 eV on a SiO2 substrate with tree S-POPs vibrating modes. The solid white curves represent the traces of the corresponding peak frequencies (Eq. (24)). Inset depicts a zoomed-in view of the plots around the three POPs excitation energies (ħω1,2,3C ≈ 0.233, 0.312, and 0.346). The dashed lines show the allowed damping boundaries.

Download Full Size | PDF

All data shown thus far, except for those shown in Fig. 3, are obtained from the dispersion relation, Eq. (13), we have modeled using the existing experimental data [12,16,30], via modifying the electron polarizability, Π0, as given in Eq. (14). Only for the substrate with two dominant vibrational frequencies, the experimental data are not available. Nonetheless, we presumed our accurate model for the cases with none, one, and three vibrational modes can also be precise for the substrate with two vibrational modes.

The scattering rate of the electrons by phonons is proportional to ωλ. Hence, electrons track slower vibrations. That means the scattering rate decrease as the speed of motion increases. In substrates with four or more phonons ωλ≥4 is high enough for neglecting their scattering effect on the graphene surface plasmons.

3.2 Plasmonic content

For calculating the output current of graphene-based optoelectronic devices, such as plasmonic photodetectors [30], placed on a polar substrate, we need to know the pure plasmonic content of the hybrid coupled plasmon-phonon modes, we have discussed in the earlier subsection. Recall that for a polar substrate with n vibrating modes on the surface, the number of hybrid coupled modes is n + 1 (Eq. (22)-Eq. (24)). The plasmonic content for either of the given dispersion mode is given by [28,30],

$${\Theta _i} = {{\prod\limits_{\lambda = 1}^n {({\omega_{\textrm{pl - ph,}i}^2 - \omega_\lambda^2} )} } / {\prod\limits_{j \ne i = 1}^{n + 1} {({\omega_{\textrm{pl - ph,}i}^2 - \omega_{\textrm{pl - ph,j}}^2} )} }}.$$

Now, we substitute Eqs. (22)-(24) into (25) to obtain the pure plasmonic content related to the dispersion modes in graphene placed on any given polar substrate. The dots, dashes, dotted-dashes, and solid curves in Fig. 5(a), (5(b)), and 5(c) illustrate the numerical data for the plasmonic content for the three substrates (i.e., SiC, AlN, and SiO2), using the same chemical potential for graphene as before. The solid curve and dashes, in each part, represent the behavior of the plasmonic contents for the first and second dispersive modes on the respective polar substrate. The dots-dashes in part (b) or (c) represent the plasmonic content of the third mode on the AlN or SiO2 substrate, whereas the dots only shown in part (c) display the content of the fourth mode on graphene placed on SiO2 substrate. The solid circles, diamonds, and squares in Fig. 5(c) represent the existing experimental data for the plasmonic content of modes ω1pl-ph, ω2pl-ph, and ω4pl-ph when graphene is on SiO2 substrate (extracted from [30]), demonstrating excellent agreements with our developed model. The inset in each part depicts the corresponding dispersion modes zoomed in about the S-POPs excitation energies ħωλ / μC.

 figure: Fig. 5.

Fig. 5. Plasmonic content of the hybrid modes in graphene on (a) SiC (b), AlN, and (c) SiO2 substrates. All conditions are the same as those used for Figs. 23, and 4. Blue circles, magenta diamonds, and green squares in (c) represent the existing experimental data (extracted from [30]) for the plasmonic contents of the first, second, and fourth modes when graphene is on SiO2 substrate. Insets display the corresponding coupled-mode dispersions.

Download Full Size | PDF

As can be observed from this figure, for small qs, (i.e., q/kF ≪ 0.05), where ℏω1pl-ph≪ ℏω1, the corresponding plasmon content for each of the three cases shown in this figure is mainly plasmonic. As q/kF increases and ℏω1pl-ph→ℏω1, the plasmonic content of the first mode starts transferring to the second mode (see Fig. 2), and its content becomes predominantly vibrational for q/kF > 0.1. Contrary to the first mode, the second hybrid mode content, at q/kF ≪ 0.05 for all three given substrates, is dominantly vibrational, and as q increases, the plasmonic content increases. Although this increase for SiC continues until it becomes predominantly plasmonic for large qs, for the other two substrates, the content maximizes as ω2pl-phω2 and then starts to drop fast and hence becomes predominantly vibrational again. The dots-dashes show that the plasmonic content of the third mode for the polar substrate with two vibrational modes behaves like that of the second mode on the substrate with only one S-POPs mode (Fig. 5(b)), whereas the behavior of the one for the polar substrate with three vibrational excitations is somewhat similar to the content of the second hybrid mode for the substrate with two S-POPs modes (5(c)). Similar to those for the second mode on SiC and the third mode on AlN substrates, the plasmonic content of the fourth mode on SiO2 substrate grows as q increases until it becomes predominantly plasmonic (q / kF > 0.2). Moreover, as can be observed from Fig. 5(c), the peak content for the third mode is about an order of magnitude smaller than that of the second mode. That is because of the (ω3 − ω2) << (ω4 − ω3).

3.3 Hybrid mode lifetime

Here, we model the coupled plasmon-phonon mode lifetime, τc, which depends on both electrons’ and phonons’ relaxation times (τpl and τPOP). Hence, to obtain τc, we need to know the plasmonic and phononic contents. Knowing the plasmonic content for each of the three exemplar substrates, Θi(%) from Eq. (25) and the corresponding τpl and τPOP from (15) and (19), we can deduce the coupled-mode lifetime as

$${\tau _{\textrm{c,i}}} = {\Theta _\textrm{i}}{\tau _{\textrm{pl}}} + ({1 - {\Theta _\textrm{i}}} ){\tau _{\textrm{POP}}}.$$

The dots, dashes, dotted-dashes, and solid curves in Fig. 6 illustrates the results of our calculations for the lifetimes of the hybrid coupled modes versus wavenumber q on graphene placed on either of the three exemplar substrates — i.e., SiC (Fig. 6(a)), AlN (Fig. 6(b)), and SiO2 (Figs. 6(c)). Consistent with data illustrated in Fig. 5, for small wavenumbers (i.e., q / kF ≪ 0.05), where the first mode (ℏω1pl-ph) is predominantly plasmonic, the corresponding hybrid coupled-mode lifetime for each of the three examples is τc1τ0 = 85 fs, indicating the dominance of the impurity scattering in that region. As q and ℏω1pl-ph increase, τc1 decreases due to the edge scatterings until the mode’s content becomes almost vibrational when ℏω1pl-ph→ ℏω1. Then, any further increase in q, while ℏω1pl-ph is saturating just below ℏω1, the phononic content gradually increases (Fig. 5), and τc1 starts to rise until the dispersive mode damps into the SPEintra continuum (Figs. 24), beyond which τc1 decreases rapidly. Concerning the plasmonic content of the higher hybrid modes in either of the three examplar cases, they all start at their corresponding phonon excitation frequencies for very small wavenumbers — i.e., they are predominantly vibrational (τci(i≥2)τPOP). As q and energy increase, the hybrid mode lifetimes approach the plasmons lifetimes (τpl, i(≥2)). Further increase in q increases the edge scattering lifetime (τed), τPE, and τSPE, reducing the total hybrid mode lifetime. The solid circles in Fig. 6(c) represent the existing experimental data for the lifetime of the fourth mode (ω4pl-ph) when graphene is on SiO2 substrate (extracted from [16]), demonstrating an excellent agreement with our developed model.

 figure: Fig. 6.

Fig. 6. The lifetimes of the hybrid modes in graphene on (a) SiC (b), AlN, and (c) SiO2. All conditions are the same as those used for Figs. 23, and 4. Green squares in (c) represent the existing experimental data (extracted from [16]) for the lifetime of the fourth mode when graphene is on SiO2 substrates. Insets display the corresponding coupled-mode dispersions.

Download Full Size | PDF

As depicted by dashes in Fig. 6(b) and 6(c) and dots-dashes in Fig. 6(c), at intermediate frequencies, the behavior of the corresponding lifetimes is readily comprehensible by considering the vibrational contents. At wavenumbers for which the hybrid modes are out of the SPEintra continuum, an increase in the phononic content (1−Θ) increases the total lifetime. As the hybrid modes enter the intraband damping region, the lifetime decreases even faster to a few fs.

4. Conclusion

Using the random phase approximation method, we have derived an exact dispersion relation for the plasmonic mode on the graphene surface placed on a nonpolar substrate. Our numerical results show that the graphene surface plasmons on a DLC substrate for large wavenumbers vary as ω (q) $\propto$ q3/2, dissipating in the SPEinter continuum and away from the SPEintra continuum consistent with the experimental measurements published by others. On the contrary, the classical 2D dispersion formula, derived by others, shows large wavenumbers ω $\propto$q1/2, entering the SPEintra continuum. Along the same line, using the frequency of each S-POPs mode on the surface of the polar dielectric substrates with λ (=1, 2, or 3) vibrational excitations, we have derived exact dispersion relations for λ+1 hybrid coupled plasmon-phonon modes. Considering graphene monolayers placed on polar substrates like SiC with one, AlN with two, and SiO2 with three S-POPs vibrational frequencies, we have calculated the loss spectra and dispersion relations for all allowed hybrid plasmon-phonon modes in graphene. These results are in excellent agreement with the existing experimental data obtained by others. Moreover, we have found the plasmonic content of each hybrid mode and the related lifetime versus q. Employing these results, one can readily predict the output currents of optoelectronic devices based on graphene plasmonics made on polar and nonpolar dielectric substrates.

Appendix A: parameters of Eq. (23)

Parameter α, β, and χ given in Eq. (23) are

$$\alpha = \frac{2}{3}{\Omega ^2} + \frac{1}{3}\left( {{\gamma^2} + \sum\limits_{\lambda = 1}^2 {\omega_\lambda^2} } \right),$$
$$\beta = \sqrt[3]{{\delta + \sqrt {{\delta ^2} - {\chi ^3}} }},$$
and
$$\chi = {\alpha ^2} - \eta ,$$
wherein
$$\delta = \alpha \left( {\chi - \frac{1}{2}\eta } \right) + \phi ,$$
$$\eta = \frac{2}{3}{\Omega ^2}\left( {\frac{{ - 4\pi }}{\hbar }\sum\limits_{\lambda = 1}^2 {{{\cal F}}_\lambda^2{\omega_\lambda }} + \sum\limits_{\lambda = 1}^2 {\omega_\lambda^2} } \right) + \frac{1}{3}\left( {{\gamma^2}\sum\limits_{\lambda = 1}^2 {\omega_\lambda^2} + \prod\limits_{\lambda = 1}^2 {\omega_\lambda^2} } \right),$$
and
$$\phi = \left[ {{\Omega ^2}\left( {\frac{{ - 4\pi }}{\hbar }\sum\limits_{\scriptstyle\lambda ,\mu = 1\atop \scriptstyle\lambda \ne \mu}^2 {{{\cal F}}_\lambda^2{\omega_\mu }} + \prod\limits_{\lambda = 1}^2 {\omega_\lambda^2} } \right) + \frac{1}{2}{\gamma^2}\prod\limits_{\lambda = 1}^2 {\omega_\lambda^2} } \right]\prod\limits_{\lambda = 1}^2 {\omega _\lambda ^2} .$$

Appendix B: parameters of Eq. (24)

We show the details of parameter A to F given in Eq. (24) here,

$$\Delta = {{\left( {({2{\Omega ^2} + {\gamma^2}} )+ \sum\limits_{\lambda = 1}^3 {\omega_\lambda^2} } \right)} / 4},$$
$$\Gamma ={-} 54{\kappa _2}{\Lambda ^{1/2}} - 18{\kappa _1}{\Lambda ^{{1 / 3}}}{\Phi ^{{1 / 2}}} - {\Phi ^{3/2}},$$
$$\Phi = {\kappa _1}({{\kappa_1} - 6{\Lambda ^{1/3}}} )- 12{\kappa _3} + 9{\Lambda ^{{2 / 3}}},$$
$${\rm K} ={+} 54{\kappa _2}{\Lambda ^{1/2}} - 18{\kappa _1}{\Lambda ^{{1 / 3}}}{\Phi ^{{1 / 2}}} - {\Phi ^{3/2}}.$$
$$\Lambda = {{{{\kappa _1^3} / {27}} + \kappa _2^2} / 2} + {{4{\kappa _1}{\kappa _3}} / 3} + {{{{({3{\kappa_4}} )}^{{1 / 2}}}} / {18}},$$

The parameters κj (j = 1, 2, 3, 4) in (B1)-(B5) are,

$${\kappa _1} ={-} 6{\Delta ^2} + {\xi _1},$$
$${\kappa _2} = 8{\Delta ^3} - 2\Delta {\xi _1} + {\xi _2} + {\xi _3},$$
$${\kappa _3} = 3{\Delta ^4} - {\Delta ^2}{\xi _1} + \Delta ({{\xi_2} + {\xi_3}} )+ {\xi _4},$$
$${\kappa _4} = \kappa _2^2({27\kappa_2^2 + {\kappa_1}({4\kappa_1^2 + 144{\kappa_3}} )} )+ \kappa _1^2{\kappa _3}({16\kappa_1^2 + 128{\kappa_3}} )+ 256\kappa _3^3.$$

Parameters ξj (j = 1, 2, 3, and 4) in (B6)-(B9) are

$${\xi _1} = \sum\limits_{\lambda = 1}^3 {\left[ { - \Omega _\lambda^4 + ({2{\Omega ^2} + {\gamma^2}} )\omega_\lambda^2 + \frac{1}{2}\sum\limits_{\mu \ne \lambda = 1}^3 {\omega_\lambda^2} \omega_\mu^2} \right]} ,$$
$${\xi _2} ={-} \frac{1}{2}\sum\limits_{\scriptstyle\lambda ,\mu ,\nu = 1\atop \scriptstyle\lambda \ne \mu \ne \nu }^3 {\Omega _\lambda ^4({\omega_\mu^2 + \omega_\nu^2} )} ,$$
$${\xi _3} = \frac{1}{2}({2{\Omega ^2} + {\gamma^2}} )\sum\limits_{\scriptstyle\lambda ,\mu = 1\atop \scriptstyle\lambda \ne \mu }^3 {\omega _\lambda ^2} \omega _\mu ^2 + \prod\limits_{\lambda = 1}^3 {\omega _\lambda ^2} ,$$
$${\xi _4} = \frac{1}{2}\sum\limits_{\lambda \ne \mu \ne \nu = 1}^3 {\Omega _\lambda ^4\omega _\mu ^2\omega _\nu ^2} - ({2{\Omega ^2} + {\gamma^2}} )\prod\limits_{\lambda = 1}^3 {\omega _\lambda ^2} ,$$
where
$$\Omega _\lambda ^4 = \frac{{8\pi {\Omega ^2}}}{\hbar }\,{{\cal F}}_\lambda ^2{\omega _\lambda },\quad \textrm{with }\lambda = 1,2,3$$

Funding

Tarbiat Modares University (IG-39703).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. G. D. Mahan and L. M. Woods, “Phonon-modulated electron-electron interactions,” Phys. Rev. B 60(8), 5276–5281 (1999). [CrossRef]  

2. F. Giustino, “Electron-phonon interactions from first principles,” Rev. Mod. Phys. 89(1), 015003 (2017). [CrossRef]  

3. K. M. Borysenko, J. T. Mullen, E. A. Barry, S. Paul, Y. G. Semenov, J. M. Zavada, M. Buongiorno Nardelli, and K. W. Kim, “First-principles analysis of electron-phonon interactions in graphene,” Phys. Rev. B 81(12), 121412 (2010). [CrossRef]  

4. 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]  

5. A. T. Lawal, “Graphene-based nano composites and their applications. A review,” Biosens. Bioelectron. 141, 111384 (2019). [CrossRef]  

6. A. Nag, A. Mitra, and S. C. Mukhopadhyay, “Graphene and its sensor-based applications: A review,” Sens. Actuators, A 270, 177–194 (2018). [CrossRef]  

7. A. Kumar, K. Sharma, and A. R. Dixit, “A review of the mechanical and thermal properties of graphene and its hybrid polymer nanocomposites for structural applications,” J. Mater. Sci. 54(8), 5992–6026 (2019). [CrossRef]  

8. T. Low and P. Avouris, “Graphene plasmonics for terahertz to mid-infrared applications,” ACS Nano 8(2), 1086–1101 (2014). [CrossRef]  

9. L. Ju, B. Geng, J. Horng, C. Girit, M. Martin, Z. Hao, H. A. Bechtel, X. Liang, A. Zettl, Y. R. Shen, and F. Wang, “Graphene plasmonics for tunable terahertz metamaterials,” Nat. Nanotechnol. 6 (10), 630–634 (2011). [CrossRef]  

10. S. Xiao, X. Zhu, B. H. Li, and N. Asger Mortensen, “Graphene-plasmon polaritons: From fundamental properties to potential applications,” Front. Phys. 11(2), 117801 (2016). [CrossRef]  

11. G. F. J. de Abajo, “Graphene plasmonics: challenges and opportunities,” ACS Photonics 1(3), 135–152 (2014). [CrossRef]  

12. Y. Liu, R. F. Willis, K. V. Emtsev, and T. Seyller, “Plasmon dispersion and damping in electrically isolated two-dimensional charge sheets,” Phys. Rev. B 78 (20), 201403 (2008). [CrossRef]  

13. Y. Liu and R. F. Willis, “Plasmon-phonon strongly coupled mode in epitaxial graphene,” Phys. Rev. B 81(8), 081406 (2010). [CrossRef]  

14. V. W. Brar, M. S. Jang, M. Sherrott, S. Kim, J. J. Lopez, L. B. Kim, M. Choi, and H. Atwater, “Hybrid surface-phonon-plasmon polariton modes in graphene/monolayer h-BN heterostructures,” Nano Lett. 14(7), 3876–3880 (2014). [CrossRef]  

15. I. D. Barcelos, A. R. Cadore, L. C. Campos, A. Malachias, K. Watanabe, T. Taniguchi, F. C. B. Maia, R. Freitasd, and C. Deneke, “Graphene/h-BN plasmon–phonon coupling and plasmon delocalization observed by infrared nano-spectroscopy,” Nanoscale 7(27), 11620–11625 (2015). [CrossRef]  

16. 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]  

17. J. H. Chen, C. Jang, S. Xiao, M. Ishigami, and M. S. Fuhrer, “Intrinsic and extrinsic performance limits of graphene devices on SiO2,” Nat. Nanotechnol. 3(4), 206–209 (2008). [CrossRef]  

18. S. Fratini and F. Guinea, “Substrate-limited electron dynamics in graphene,” Phys. Rev. B 77(19), 195415 (2008). [CrossRef]  

19. E. H. Hwang and S. D. Sarma, “Acoustic phonon scattering limited carrier mobility in two-dimensional extrinsic graphene,” Phys. Rev. B 77(11), 115449 (2008). [CrossRef]  

20. T. Low, V. Perebeinos, R. Kim, and M. Freitag Avouris, “Cooling of photoexcited carriers in graphene by internal and substrate phonons,” Phys. Rev. B 86(4), 045413 (2012). [CrossRef]  

21. S. D. Sarma and F. Wu, “Electron–phonon and electron–electron interaction effects in twisted bilayer graphene,” Ann. Phys. 417, 168193 (2020). [CrossRef]  

22. E. H. Hwang, R. Sensarma, and S. D. Sarma, “Plasmon-phonon coupling in graphene,” Phys. Rev. B 82(19), 195406 (2010). [CrossRef]  

23. D. M. Basko, “Theory of resonant multiphonon Raman scattering in graphene,” Phys. Rev. B 78(12), 125418 (2008). [CrossRef]  

24. M. Mohr, J. Maultzsch, E. Dobardžić, S. Reich, I. Milošević, M. Damnjanović, A. Bosak, M. Krisch, and C. Thomsen, “Phonon dispersion of graphite by inelastic x-ray scattering,” Phys. Rev. B 76(3), 035439 (2007). [CrossRef]  

25. V. Nagyte, D. J. Kelly, A. Felten, G. Picardi, Y. Y. Shin, A. Alieva, R. E. Worsley, K. Parvez, S. Dehm, R. Krupke, S. J. Haigh, A. Oikonomou, A. J. Pollard, and C. Casiraghi, “Raman Fingerprints of Graphene Produced by Anodic Electrochemical Exfoliation,” Nano Lett. 20(5), 3411–3419 (2020). [CrossRef]  

26. C. Casiraghi, A. Hartschuh, H. Qian, S. Piscanec, C. Georgi, A. Fasoli, K. S. Novoselov, D. M. Basko, and A. C. Ferrari, “Raman spectroscopy of graphene edges,” Nano Lett. 9(4), 1433–1441 (2009). [CrossRef]  

27. K. Hess and P. Vogl, “Remote polar phonon scattering in silicon inversion layers,” Solid State Commun. 30(12), 797–799 (1979). [CrossRef]  

28. M. V. Fischetti, D. A. Neumayer, and E. A. Cartier, “Effective electron mobility in Si inversion layers in metal–oxide–semiconductor systems with a high-κ insulator: The role of remote phonon scattering,” J. Appl. Phys. 90(9), 4587–4608 (2001). [CrossRef]  

29. E. H. Hwang and S. D. Sarma, “Dielectric function, screening, and plasmons in two-dimensional graphene,” Phys. Rev. B 75(20), 205418 (2007). [CrossRef]  

30. M. Freitag, T. Low, W. Zhu, H. Yan, F. Xia, and P. Avouris, “Photocurrent in graphene harnessed by tunable intrinsic plasmons,” Nat. Commun. 4(1), 1951–1958 (2013). [CrossRef]  

31. S. Cakmakyapan, P. K. Lu, A. Nabavi, and M. Jarrahi, “Gold-patched graphene nano-stripes for high-responsivity and ultrafast photodetection from the visible to infrared regime,” Light: Sci. Appl. 7(1), 1–9 (2018). [CrossRef]  

32. K. Ishikawa and T. Ando, “Optical phonon interacting with electrons in carbon nanotubes,” J. Phys. Soc. Jpn. 75(8), 084713 (2006). [CrossRef]  

33. H. Bruus and K. Flensberg, Many-body quantum theory in condensed matter physics: an introduction, Oxford university press, 2004.

34. B. Scharf, V. Perebeinos, J. Fabian, and P. Avouris, “Effects of optical and surface polar phonons on the optical conductivity of doped graphene,” Phys. Rev. B 87(3), 035414 (2013). [CrossRef]  

35. R. D Mattuck, A guide to Feynman diagrams in the many-body problem. Courier Corporation, 1992.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (6)

Fig. 1.
Fig. 1. Plasmons modes loss spectra (Eq. (20)) in graphene with μC = 0.3 eV on a DLC substrate. The white solid curve represents the corresponding peak frequency plasmons wavenumber (Eq. (21)). The dashed lines show the allowed damping boundaries. The green dots represent the classical 2D dispersion (i.e., γ = 0).
Fig. 2.
Fig. 2. Hybrid modes loss spectra (Eq. (20)) in graphene with μC = 0.57 eV on a SiC substrate with a single POP-vibrating mode. The two white solid curves represent the corresponding peak frequencies versus the plasmons wavenumber (Eq. (22)). Inset depicts a zoomed-in view of the plots about S-POPs energy (ħω1C ≈ 0.2). The dashed lines show the allowed damping boundaries.
Fig. 3.
Fig. 3. Hybrid modes loss spectra (Eq. (20)) in graphene with μC = 0.3 eV on an AlN substrate with two S-POPs vibrating modes. The three solid white curves represent the traces of the corresponding peak frequencies (Eq. (23)). Inset depicts a zoomed-in view of the plots around the two S-POPs excitation energies (ħω1,2C ≈ 0.28 and 0.35). The dashed lines show the allowed damping boundaries.
Fig. 4.
Fig. 4. Hybrid modes loss spectra (Eq. (20)) in graphene with μC = 0.43 eV on a SiO2 substrate with tree S-POPs vibrating modes. The solid white curves represent the traces of the corresponding peak frequencies (Eq. (24)). Inset depicts a zoomed-in view of the plots around the three POPs excitation energies (ħω1,2,3C ≈ 0.233, 0.312, and 0.346). The dashed lines show the allowed damping boundaries.
Fig. 5.
Fig. 5. Plasmonic content of the hybrid modes in graphene on (a) SiC (b), AlN, and (c) SiO2 substrates. All conditions are the same as those used for Figs. 23, and 4. Blue circles, magenta diamonds, and green squares in (c) represent the existing experimental data (extracted from [30]) for the plasmonic contents of the first, second, and fourth modes when graphene is on SiO2 substrate. Insets display the corresponding coupled-mode dispersions.
Fig. 6.
Fig. 6. The lifetimes of the hybrid modes in graphene on (a) SiC (b), AlN, and (c) SiO2. All conditions are the same as those used for Figs. 23, and 4. Green squares in (c) represent the existing experimental data (extracted from [16]) for the lifetime of the fourth mode when graphene is on SiO2 substrates. Insets display the corresponding coupled-mode dispersions.

Tables (2)

Tables Icon

Table 1. Various permittivities, the dominant phonon energies, and related Fröhlich coupling constants for dielectrics substrates SiC, Al2O3, HfO2, AlN, ZrO2, ZrSiO4, hexagonal BN (hBN), and SiO2

Tables Icon

Table 2. Feynman diagrams and their descriptions

Equations (58)

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

V el - el op = 1 2 A k 1 σ 1 k 2 σ 2 q λ 1 A | g q λ op | 2 G λ 0 ( ω ) c ^ k 1 + q , σ 1 c ^ k 2 + q , σ 2 c ^ k 2 σ 2 c ^ k 1 σ 1 1 2 A 2 k , σ , q , λ v λ op c ^ k + q , σ c ^ k σ ,
| g q op | 2 = D op 2 2 ρ m ω op ,
G 0 = 2 ω op / ( ω + i / τ op ) 2 ω op 2 .
v op = | g q op | 2 G 0 ( ω ) ,
v λ POP = λ | g q λ POP | 2 G λ 0 ( ω ) .
| g q λ POP | 2 = π e 2 ε 0 F λ 2 exp ( 2 q z 0 ) q π e 2 q ε 0 F λ 2 ,
F 1 2 = ω 1 2 π ( 1 ε + 1 1 ε 0 + 1 ) .
F 1 2 = ω 1 2 π ( 1 ε i,1 + 1 1 ε 0 + 1 )
F 2 2 = ω 2 2 π ( 1 ε + 1 1 ε i,1 + 1 ) ,
F 1 2 = ω 1 2 π ( 1 ε i,1 + 1 1 ε 0 + 1 ) ,
F 2 2 = ω 2 2 π ( 1 ε i,2 + 1 1 ε i, 1 + 1 ) ,
F 3 2 = ω 3 2 π ( 1 ε + 1 1 ε i,2 + 1 ) .
⇑=↑ + Π 0 + Π 0 Π 0 + Π 0 Π 0 Π 0 + =↑ + { + Π 0 + Π 0 Π 0 + } × Π 0 =↑ + Π 0
V eff = ( V C + V ph ) + V eff × Π 0 × ( V C + V ph ) ,
V eff = ( V C + V ph ) / [ 1 ( V C + V ph ) Π 0 ] .
V eff = ( V C + V ph ) / ε ( q , ω ) .
ε ( q , ω ) = ε env ( e 2 2 q ε 0 + | g q op | 2 G 0 ( ω ) + λ | g q λ POP | 2 G λ 0 ( ω ) ) Π 0 ,
Π 0 μ C q 2 π 2 ( ( ω + i / τ pl ) 2 γ 2 ) ,
γ 2 = 1 8 ν F 2 q 2 [ exp ( q / q ) 1 ] 1 ,
q = 2 π ε 0 ε env e 2 μ C ( ω op ) 2 .
τ pl 1 = τ 0 1 + τ PE 1 + τ SPE 1 + τ ed 1 ,
τ PE 1 = g PE | ( ω + ω α ) + μ C | [ F ( ω , ω α ) + 2 ] ,
F ( ω , ω α ) = erf ( ω ω α 2 k B T ) + erf ( ω ω α 2 k B T ) ,
τ SPE 1 = τ SPE,intra 1 + τ SPE,inter 1 .
τ SPE,intra 1 = g SPE,intra | ( ω + ω α ) μ C 3 | F ( ω , ω α ) ,
τ SPE,inter 1 = g SPE,inter | ( ω + ω α ) + μ C | [ F ( ω , ω α ) + 2 ] ,
τ ed 2 π F × 10 6 q .
τ POP τ POP,0 exp ( q 4 × 10 7 ) ,
L Im { ε ( q , ω ) } 1 .
ω = e 2 μ C 2 π 2 ε 0 ε env q + γ 2 .
ω 1 , 2 pl - ph ( q , μ C ) = Ω { 1 + 1 2 Ω 2 ( ω 1 2 + γ 2 ) [ 1 + 1 4 Ω 4 ( ω 1 2 γ 2 ) 2 + 1 Ω 2 ( 8 π F 1 2 ω 1 ( ω 1 2 γ 2 ) ) ] 1 / 2 } 1 / 2
Ω 2 = q μ C e 2 4 π ε 0 ε env 2 ,
ω 1 , 2 pl - ph = Re α 1 2 ( β + χ / β ) ± i 3 2 ( β χ / β ) ,
ω 3 pl - ph = Re α + ( β + χ / β ) ,
ω 1 , 2 pl - ph = Re Δ Φ 1 / 2 ( 1 ± Γ 1 / 2 / Φ 3 / 4 ) / 6 Λ 1 / 6 ,
ω 3 , 4 pl - ph = Re Δ Φ 1 / 2 ( 1 ± K 1 / 2 / Φ 3 / 4 ) / 6 Λ 1 / 6 .
Θ i = λ = 1 n ( ω pl - ph, i 2 ω λ 2 ) / j i = 1 n + 1 ( ω pl - ph, i 2 ω pl - ph,j 2 ) .
τ c,i = Θ i τ pl + ( 1 Θ i ) τ POP .
α = 2 3 Ω 2 + 1 3 ( γ 2 + λ = 1 2 ω λ 2 ) ,
β = δ + δ 2 χ 3 3 ,
χ = α 2 η ,
δ = α ( χ 1 2 η ) + ϕ ,
η = 2 3 Ω 2 ( 4 π λ = 1 2 F λ 2 ω λ + λ = 1 2 ω λ 2 ) + 1 3 ( γ 2 λ = 1 2 ω λ 2 + λ = 1 2 ω λ 2 ) ,
ϕ = [ Ω 2 ( 4 π λ , μ = 1 λ μ 2 F λ 2 ω μ + λ = 1 2 ω λ 2 ) + 1 2 γ 2 λ = 1 2 ω λ 2 ] λ = 1 2 ω λ 2 .
Δ = ( ( 2 Ω 2 + γ 2 ) + λ = 1 3 ω λ 2 ) / 4 ,
Γ = 54 κ 2 Λ 1 / 2 18 κ 1 Λ 1 / 3 Φ 1 / 2 Φ 3 / 2 ,
Φ = κ 1 ( κ 1 6 Λ 1 / 3 ) 12 κ 3 + 9 Λ 2 / 3 ,
K = + 54 κ 2 Λ 1 / 2 18 κ 1 Λ 1 / 3 Φ 1 / 2 Φ 3 / 2 .
Λ = κ 1 3 / 27 + κ 2 2 / 2 + 4 κ 1 κ 3 / 3 + ( 3 κ 4 ) 1 / 2 / 18 ,
κ 1 = 6 Δ 2 + ξ 1 ,
κ 2 = 8 Δ 3 2 Δ ξ 1 + ξ 2 + ξ 3 ,
κ 3 = 3 Δ 4 Δ 2 ξ 1 + Δ ( ξ 2 + ξ 3 ) + ξ 4 ,
κ 4 = κ 2 2 ( 27 κ 2 2 + κ 1 ( 4 κ 1 2 + 144 κ 3 ) ) + κ 1 2 κ 3 ( 16 κ 1 2 + 128 κ 3 ) + 256 κ 3 3 .
ξ 1 = λ = 1 3 [ Ω λ 4 + ( 2 Ω 2 + γ 2 ) ω λ 2 + 1 2 μ λ = 1 3 ω λ 2 ω μ 2 ] ,
ξ 2 = 1 2 λ , μ , ν = 1 λ μ ν 3 Ω λ 4 ( ω μ 2 + ω ν 2 ) ,
ξ 3 = 1 2 ( 2 Ω 2 + γ 2 ) λ , μ = 1 λ μ 3 ω λ 2 ω μ 2 + λ = 1 3 ω λ 2 ,
ξ 4 = 1 2 λ μ ν = 1 3 Ω λ 4 ω μ 2 ω ν 2 ( 2 Ω 2 + γ 2 ) λ = 1 3 ω λ 2 ,
Ω λ 4 = 8 π Ω 2 F λ 2 ω λ , with  λ = 1 , 2 , 3
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.