We study theoretically the in-plane electromagnetic response and the exciton–plasmon interactions for an experimentally feasible carbon nanotube (CN) film system composed of parallel aligned periodic semiconducting CN arrays embedded in an ultrathin finite-thickness dielectric. For homogeneous single-CN films, the intertube coupling and thermal broadening bring the exciton and interband plasmon resonances closer together. They can even overlap due to the inhomogeneous broadening for films composed of array mixtures with a slight CN diameter distribution. In such systems, the real part of the response function is negative for a broad range of energies (negative refraction band), and the CN film behaves as a hyperbolic metamaterial. We also show that for a properly fabricated two-component CN film, by varying the relative weights of the two constituent CN array components, one can tune the optical absorption profile to make the film transmit or absorb light in the neighborhood of an exciton absorption resonance on-demand.

Carbon Nanotubes (CNs) and ultrathin films made of periodic CN arrays provide a variety of useful physical properties that are essential for optoelectronic device applications.1–5 Periodically aligned CN array films, in particular, provide stability and precise tunability of their characteristics by means of the CN diameter, intertube distance, and film thickness variation, which is why they are getting more and more attention of experimental communities.6–17 Self-assembled quasiperiodic finite-thickness single-wall CN (SWCN) films have been recently shown experimentally to exhibit extraordinary optoplasmonic properties such as a tunable negative dielectric response in a broad range of the photon excitation energies.13 Using the Maxwell–Garnett (MG) mixing method18 and the many-particle Green’s function formalism,19 we have lately explained this theoretically to be the case due to the inhomogeneous broadening effect in films composed of mixed SWCN arrays with a narrow nanotube diameter distribution.20 

In this contribution, we focus on the exciton–plasmon coupling in a two-component mixture of periodic SWCN arrays. We start with the analysis of the dielectric response of a homogeneous quasi-2D array of parallel aligned, uniformly spaced, identical SWCNs immersed in a finite-thickness dielectric medium of static relative dielectric permittivity ε, sandwiched between a substrate and a superstrate of relative permittivities ε1 and ε2. A schematic of the array system under study and notations of relevance are presented in the inset of Fig. 1. Next, we present a numerical study on the role of the exciton–plasmon coupling in the optical absorption spectrum of a two-component mixture of the homogeneous SWCN arrays. We discuss how this coupling and thereby the absorption can be controlled by varying the relative weights of the individual components in the mixture.

FIG. 1.

The dielectric response functions of the (10,0) CN film of thickness d=3R, where R=3.91 Å is the radius of (10,0) CN, along the CN alignment direction as functions of qR and the photon energy. All graphs are calculated from Eq. (1) for the energies in the neighborhood of the first exciton resonance of the (10,0) CN. The fraction fCN=π/9 and ε=10, ε1=ε2=1. The red and blue colored surfaces depict the refraction band (Reεyy) and the exciton absorption resonance (Imεyy), respectively. The green colored surface shows the interband plasmon resonance (Im1/εyy). The inset shows a schematic of the geometry of relevance.

FIG. 1.

The dielectric response functions of the (10,0) CN film of thickness d=3R, where R=3.91 Å is the radius of (10,0) CN, along the CN alignment direction as functions of qR and the photon energy. All graphs are calculated from Eq. (1) for the energies in the neighborhood of the first exciton resonance of the (10,0) CN. The fraction fCN=π/9 and ε=10, ε1=ε2=1. The red and blue colored surfaces depict the refraction band (Reεyy) and the exciton absorption resonance (Imεyy), respectively. The green colored surface shows the interband plasmon resonance (Im1/εyy). The inset shows a schematic of the geometry of relevance.

Close modal

The SWCNs are aligned along the y-axis in a dielectric layer of thickness d with the intertube center-to-center distance Δ. The electron charge density is distributed uniformly all over the periodic cylindrical nanotube surfaces, whereby the pairwise electron Coulomb interaction in the system of CNs can be approximated by that of two uniformly charged rings of radius R of the respective nth and th tubules.21 We consider that the dielectric medium embedding a periodic array has much greater dielectric permittivity than the substrate and superstrate permittivities. In this case,22 the Coulomb interaction in the film increases strongly as the film thickness decreases, making the Coulomb interaction independent of the vertical coordinate component.21–25 Although this vertical confinement leads to the reduction of the effective dimensionality from three to two, the thickness d is still a variable parameter there to represent the finite vertical size of the system, which the dielectric response of the film depends on.

We proceed with our studies as follows. First, we determine the optical conductivity of an individual SWCN as a function of photon frequency using the (kp)-band-structure method.26 Then, taking advantage of the vertical electron confinement nature of stronger Keldysh–Rytova interaction potential,21–24 we determine the effective interaction of an ultrathin finite-thickness periodically aligned CN array and derive an expression for the dynamical dielectric response tensor of the CN array using the low-energy plasmonic response calculation technique21 and the many-particle Green’s function formalism.19 We then mix two non-identical CN arrays using the Maxwell-Garnett mixing scheme18 and examine what happens on a CN film’s absorption spectrum if the film has only a few non-identical homogeneous CN arrays with possible exciton–plasmon coupling presence.

Previously, the exciton–plasmon coupling was studied for a variety of nanostructured systems such as organic semiconductors, SiC nanocrystals, gold nanoshells and nanorods, metallic dimers, and hybrid metal–semiconductor nanostructures.27–34 The exciton–plasmon coupling in an individual CN has been investigated theoretically in Refs. 35–40 including the case with a perpendicular electrostatic field applied.38–40 The interactions of excitonic states with interband plasma modes of a semiconducting SWCN were shown to result in a strong exciton–plasmon coupling that splits the exciton absorption line shape (Rabi splitting).38 In recent studies, such a line splitting in CN film spectra has been observed experimentally,8–10 though interpreted slightly differently for this key feature of the exciton–plasmon interaction. We here take an opportunity to demonstrate these effects with rigorous analytical and numerical calculations, explaining how one can control them with a two-component SWCN array film.

We show that for a two-component closely packed mixture of two periodic arrays of identical small-diameter (1 nm) semiconducting SWCNs, in which one SWCN array has the exciton absorption resonance energy to match the other array interband plasmon resonance energy, the near-field exciton–plasmon coupling of the two arrays can provide an extra means to control the light absorption and/or emission by the mixture, thus unveiling a path toward new tunable optoelectronic device applications.36–46 Specifically, by varying the constituent array fractions, it is possible to tune the near-field exciton–plasmon interaction strength, which is useful in applications such as electroplasmonic switch45 as well as quantum information processing, low threshold laser, light-emitting diodes, and solar cells.46 

Absorption of a photon of an external light source excites an electron from the valence to conduction band of a CN, thereby creating an electron–hole pair, an exciton. This makes a longitudinal polarization effect with an induced dipole moment directed predominantly along the nanotube axis, due to a strongly suppressed CN polarizability in the perpendicular direction,47,48 also known as the transverse depolarization.26,49 For a densely packed periodically aligned array of SWCNs, the induced inter-tubule dipole–dipole coupling should then result in the anisotropic collective polarization predominantly along the CN alignment direction of the array (Fig. 1, inset).

The following steps can be taken to determine the dielectric response tensor of the periodic SWCN array.50 The first step is to calculate the longitudinal optical conductivity σyy(ω) (the largest component of the conductivity tensor) of an individual SWCN as a function of photon frequency ω. This can be done by using the (kp) band-structure calculation method for CNs.26 In performing this step, as a matter of convenience, we express the result in the dimensionless form σ¯yy(x) by dividing σyy(ω) by e2/(2π), where x=ω/(2γ0) is the dimensionless energy in units of 2γ0 with γ0=2.7 eV being the carbon–carbon nearest neighbor overlap integral.38 Next, we evaluate the collective polarization of the homogeneous periodic SWCN array that results from the induced dipole–dipole coupling among the individual CNs of the array. We use the low-energy plasmonic response calculation technique21 combined with the many-particle Green’s function formalism19 to relate the CN longitudinal dynamical polarizability per unit length (expressed in terms of the longitudinal conductivity51) to the collective polarization of the periodic CN array. The actual derivation is analytically involved and, therefore, will be presented in full detail separately.50 

The dielectric response tensor of the CN array can be obtained using its standard relationship with the collective polarization tensor. One can see that in this model, the in-plane transverse component of the dielectric response (perpendicular to the CN alignment direction) is nothing but the static permittivity of the host dielectric layer, i.e., εxx=ε. The longitudinal component of the dielectric response is controlled by the CN longitudinal dynamical polarizability (or conductivity) and takes the form as follows (see Ref. 50 for details):

(1)

Here, the parameter fCN is a fraction indicating the relative volume occupied by the CN array in the dielectric layer of the film, i.e., fCN=NVCN/V=πR2/dΔ, where VCN is the volume of an individual SWCN of radius R, N is the total number of SWCNs, V is the volume of the film (dielectric layer embedding the array), d is the film thickness, Δ is the center-to-center distance between two adjacent CNs, I0(qR) and K0(qR) are the zeroth-order modified cylindrical Bessel functions to provide the correct normalization of the electron density distribution over cylindrical surfaces, with q being the longitudinal quasimomentum. The fraction fCN is smaller for a film with a sparse distribution of CNs and larger for the film with the close-packed CNs distribution. More explicitly, fCN satisfies the inequality 0<fCNπ/4 and is always less than unity. Therefore, the screening effect of the dielectric background cannot be eliminated completely no matter how closely the CNs are packed and how thin the film is. Even if the CNs in the array are tightly packed, i.e., Δ=2R, and the film (the dielectric layer) has the minimum possible thickness d=2R, the screening effect of the dielectric background is still present and increases if the thickness d and/or the center-to-center intertube distance Δ increase, thereby reducing the resonance peak intensities in the dielectric response.

Equation (1) links the complex-valued longitudinal (axial) conductivity of an individual SWCN, the longitudinal quasimomentum q, and the volume fraction of CNs to the complex-valued dielectric response of the entire CN film along the CN alignment direction. In this direction, the q dependence of εyy(q,x) in Eq. (1) makes the film dielectric response a strongly spatially dispersive non-local function, which can be controlled by adjusting the fraction fCN and the dielectric parameters of the CN film, while it still remains a dielectric of the static permittivity ε in the direction perpendicular to the CN alignment. For our ultrathin CN films, the predominant linear spatial dispersion comes from the finite thickness of the dielectric layer as a qd product, which is much stronger than the quadratic exciton momentum dispersion of individual semiconducting SWCNs. Also interesting, for q=0, it can be seen from Eq. (1) that εyy(0,x)=ε and the longitudinal response component of the CN array dielectric tensor becomes equal to the transverse component εxx, which is always equal to ε if the transverse depolarization is neglected. This behavior of the CN array dielectric tensor for q0 is consistent with that of an array of metallic cylinders discussed in Ref. 21.

Figure 1 shows the dielectric responses for an ultrathin (10,0) CN film of thickness d=3R and intertube distance Δ=3R, calculated from Eq. (1). We consider the host dielectric medium of the static permittivity ε=10 surrounded by air (ε1=ε2=1). For d=3R and Δ=3R, the CN fraction in a homogeneous (single-type) CN array is fCN=π/9, which is less than half of its maximum value, indicating the presence of a significant dielectric screening. The red colored surface shows the refraction band defined by Reεyy, the blue colored surface depicts the exciton absorption resonance given by Imεyy, and the green colored surface is the interband plasmon response defined as Im1/εyy, all plotted in the neighborhood of the single-tube first-subband exciton resonance. One can see the negative refraction band associated with the exciton absorption resonance per Kramers–Kronig relations. Due to the spatial dispersion and screening, the plasmon resonance of the film is much closer in energy to the exciton resonance than it occurs in an individual isolated SWCN (cf. Fig. 2, inset).

FIG. 2.

The thermally averaged response functions for an ultrathin film of the (10,0) CN array as functions of the photon energy. The room-temperature (300 K) spectrum is calculated in the neighborhood of the first exciton resonance of the (10,0) CN for the same array parameters as those used in Fig. 1. Shown in the inset for comparison are the dielectric response functions of the individual (10,0) SWCN in air, calculated by the (kp) method with the 300 K exciton–phonon energy relaxation rate and with the quadratic exciton momentum dispersion neglected.

FIG. 2.

The thermally averaged response functions for an ultrathin film of the (10,0) CN array as functions of the photon energy. The room-temperature (300 K) spectrum is calculated in the neighborhood of the first exciton resonance of the (10,0) CN for the same array parameters as those used in Fig. 1. Shown in the inset for comparison are the dielectric response functions of the individual (10,0) SWCN in air, calculated by the (kp) method with the 300 K exciton–phonon energy relaxation rate and with the quadratic exciton momentum dispersion neglected.

Close modal

To obtain the T-dependence of the CN array longitudinal dielectric response in Eq. (1), the thermal averaging can be done as follows (see Ref. 50 for details):

(2)

where the q-space population distribution function is

(3)

with

(4)

such that qfs(q,T)=1 for proper normalization. The energy ωs(q)=Es2(q)+2Es(q)Vss(q) represents the SWCN array collective eigen-state produced by a single-tube s-subband exciton of the total energy Es(q) to include the excitation energy Eexc(s)=Eg(s)+Eb(s) of the ground-state exciton (Eb and Eg being the binding and bandgap energies, respectively; excited internal states are neglected for simplicity) and its translational kinetic energy 2q2/Mex with the exciton effective mass given by the sum of the subband-dependent electron and hole effective masses. The collective eigen-states of the array originate from the intertube dipole–dipole coupling represented by the interaction matrix element of the form

where d^ind(q)=e2qRI0(qR)K0(qR)β^y is the induced dipole moment operator associated with the (predominant) operator charge displacement β^y along the SWCN axis.

Figure 2 shows the thermally averaged functions Reεyy, Imεyy, and Im1/εyy for the ultrathin film made up of the periodic array of the (10,0) SWCNs. These functions are calculated numerically as given by Eqs. (1)–(4) for T=300 K and the same array parameters as those used in Fig. 1. For comparison, the inset in Fig. 2 shows the longitudinal dynamical dielectric response functions of the individual (10,0) SWCN in air, which was calculated by the (kp) method at 300 K (where the electron Fermi-distribution effect is negligible due to high Fermi temperatures) with the (nominal) 100 fs exciton–phonon energy relaxation time37 and with the (minor) quadratic exciton momentum dispersion neglected.26 The (10,0) CN shows sharp, narrow, and intense resonance peaks, quite different from the thermally broaden resonances of the (10,0) CN array. For a single CN, the plasmon peak is positioned far away from the exciton peak. In the array of nanotubes, they are spaced much closer together. The interaction among the CNs and the thermal broadening effect bring the exciton and plasmon response functions closer to each other.

Figure 3 compares the thermally averaged (300 K) response functions for the ultrathin (10,0) and (11,0) homogeneous SWCN arrays, calculated from Eqs. (1)–(4) with the same array parameters as before. For both arrays, the exciton and plasmon resonance peaks are situated some energy apart. However, an interesting feature can be seen here that the plasmon peak of the (11,0) array is almost exactly in resonance with the exciton absorption peak of the (10,0) array, whereby a strong near-field exciton–plasmon coupling should be possible in an inhomogeneous densely packed CN film of these two arrays.

FIG. 3.

Comparison of the thermally averaged (300 K) dielectric response functions for the ultrathin arrays of the (10,0) and (11,0) SWCNs along the CN alignment direction. The tubules in each array are 3R distance apart, and the films are 3R thick with R being the respective CN radius. Other material parameters are the same as in Figs. 1 and 2. Only the first exciton–plasmon resonances are shown.

FIG. 3.

Comparison of the thermally averaged (300 K) dielectric response functions for the ultrathin arrays of the (10,0) and (11,0) SWCNs along the CN alignment direction. The tubules in each array are 3R distance apart, and the films are 3R thick with R being the respective CN radius. Other material parameters are the same as in Figs. 1 and 2. Only the first exciton–plasmon resonances are shown.

Close modal

To obtain the dielectric response functions for an inhomogeneous CN film composed of non-identical homogeneous SWCN arrays, the Maxwell-Garnett (MG) mixing method can be used18 to give the effective dielectric response ε¯yy(T,x) of the n-component mixture in the form

(5)

Here, εyy(T,x)i is the thermally averaged dielectric response of the ith homogeneous CN array component and wi is its relative weight such that i=1nwi=1 for the n-component mixture. The MG mixing of the homogeneous SWCN arrays leads to an additional inhomogeneous broadening of the exciton and plasmon resonances in the CN array mixture.50 We believe that the broad CN film dielectric response spectra reported in Ref. 13 come mostly from the inhomogeneous broadening. For a properly fabricated closely packed SWCN array mixture, broadened exciton and plasmon resonances can overlap like in the (10,0)/(11,0) CN array case shown in Fig. 3, which makes the near-field exciton–plasmon coupling strongly affect the optical properties of a composite film.

Previously, the exciton absorption line shape under the exciton–plasmon coupling was analyzed for individual CNs and composite CN materials in Refs. 9, 38, 52, and 53. We here use the approach of Ref. 38 for a properly tuned mixture of the two ultrathin homogeneous arrays of the (10,0) and (11,0) SWCNs as an example. We mix these arrays using Eq. (5), in which we have w(10,0)+w(11,0)=1, where w(10,0) and w(11,0) are the relative weights of the thermally averaged (10,0) and (11,0) CN arrays, respectively. We focus on the energy range in the neighborhood of the first exciton absorption resonance of the (10,0) CN array. As was mentioned above and is shown in Fig. 3, the interband plasmon resonance of the (11,0) CN array is almost exactly in resonance with the exciton absorption resonance of the (10,0) CN array, whereas the (10,0) CN array plasmon peak is positioned pretty much aside at higher energy and out of resonance with its own exciton absorption peak. Under these conditions, in the tightly packed mixture of the two arrays, the lineshape profile of the first exciton absorption resonance of the (10,0) CN array takes the form38 

(6)

Here, all quantities are dimensionless, i.e., normalized by 2γ0 to match Eq. (1) so that εe=E1(10,0)(q)/2γ0, xp=ωp(11,0)/2γ0, and the condition εexp is assumed to hold,38 which is indeed the case for the MG mixture we consider. Other notations are I0=Γ(εe)/2π, Γ is the exciton spontaneous decay rate into plasmons, X=4πΔxpI0 is the Rabi-splitting parameter, Δxp is the half-width-at-half-maximum of the plasmon resonance with the energy xp, and Δεe is an additional exciton energy broadening (normally attributed to the exciton–phonon relaxation for which we use τr=100fs, a nominal scattering time).

We mix these CN arrays at different weight combinations. From the room-temperature response functions Imε¯yy(x) and Im1/ε¯yy(x) of each mixture, we determine the exact exciton and interband plasmon resonance positions, εe and xp, respectively, as well as their respective intensities and the half-width-at-half-maxima Δεe and Δxp to be plugged in Eq. (6) to obtain the lineshape profile for the respective mixture of the (10,0) and (11,0) CN arrays forming the composite film. Furthermore, since only excitons are produced in optical dipole transitions stimulated by the transversely polarized electromagnetic (light) radiation, while plasmons require longitudinally polarized electromagnetic waves (electron beam) to be excited, the optical absorption is due to excitons only even though they might be mixed with plasmons to result in an absorption lineshape profile (Rabi) splitting. In view of this, the equal intensities of the split resonances in Eq. (6) should be corrected by the respective exciton and plasmon fractions, which are different from 1/2 if the exciton εe and plasmon xp resonance positions do not coincide exactly. More specifically, the exciton fraction is given by54 

(7)

where δ=εexp, and the plasmon fraction is Cp2=1Ce2, respectively. The split resonances of Eq. (6) should be multiplied by their respective fractions according to Eq. (7).

Figure 4 shows the absorption lineshape profiles of the first exciton absorption resonance of the (10,0) CN array calculated using Eqs. (6) and (7) for four different relative weight combinations of the (10,0) and (11,0) CN arrays in the MG mixture of the same material and geometry parameters as before. All profiles exhibit the line splitting (aka Rabi-splitting), which is a signature of the strong exciton–plasmon coupling.38 The line shapes are normalized by the largest peak intensity. As the relative weight of the (11,0) CN array increases, the split peak intensities tend to become even. The larger weight of the (11,0) CN array also results in an increased broadening and a greater splitting of the entire absorption profile. This splitting quickly decreases with the reduction of the (11,0) CN array relative weight to eventually turn into a single-peak resonance absorption profile for zero relative weight of the (11,0) CN array in the mixture.

FIG. 4.

Normalized lineshape profiles of the first exciton absorption resonance of the (10,0) CN array in the MG mixture with varied relative weights of the (10,0) and (11,0) CN arrays (top). The line shapes are first calculated using Eq. (6), followed by the rescaling of the two split resonance peak intensities per Eq. (7). See the text for more details.

FIG. 4.

Normalized lineshape profiles of the first exciton absorption resonance of the (10,0) CN array in the MG mixture with varied relative weights of the (10,0) and (11,0) CN arrays (top). The line shapes are first calculated using Eq. (6), followed by the rescaling of the two split resonance peak intensities per Eq. (7). See the text for more details.

Close modal

Figure 5 shows the Rabi-splitting itself of Fig. 4, computed and plotted as a function of the (11,0) CN array component relative weight. The absorption profile Rabi-splitting of the composite film can be seen to increase with the relative weight of the (11,0) CN array, thus reflecting the increasing role of the plasmon contribution in the MG mixture. The dependence is not a linear one though; for w(11,0)>0.75, it goes abruptly up. One can see the variation of the characteristic exciton absorption line splitting in the range of 0.10.5 eV in our case, which is as large as the typical exciton binding energies (0.30.8 eV55–58) of individual small-diameter semiconducting CNs. The larger Rabi-splitting indicates the stronger exciton–plasmon coupling and the decreased light absorption, accordingly, in the energy window in-between the split resonance absorption peaks. Therefore, by varying the relative weights of the two array components, one can tune the exciton–plasmon coupling and thereby the optical absorption profile of the composite film to make the film transmit or absorb light in the neighborhood of an exciton absorption resonance on-demand. A very similar effect was recently demonstrated theoretically for double-wall CNs where the near-field exciton–plasmon coupling was proposed to control the exciton Bose condensation in properly selected double-wall CN systems.43 

FIG. 5.

Rabi-splitting of the spectra shown in Fig. 4 as a function of the (11,0) CN array component relative weight in the MG mixture.

FIG. 5.

Rabi-splitting of the spectra shown in Fig. 4 as a function of the (11,0) CN array component relative weight in the MG mixture.

Close modal

In this contribution, we study theoretically the electromagnetic response for an experimentally feasible CN film system composed of periodic arrays of parallel aligned semiconducting SWCNs embedded in an ultrathin finite-thickness dielectric layer. Our main focus is the near-field exciton–plasmon interactions. We show how the exciton–plasmon coupling can be controlled by adjusting the intrinsic parameters of the system. We evaluate the in-plane dynamical dielectric response functions along the CN alignment direction using the low-energy plasmonic response calculation technique21 combined with the many-particle Green’s function formalism.19 The expression we obtain accounts for the intertube dipole–dipole interaction and links the response of the CN film to the complex optical conductivity of an individual constituent SWCN, an active component of the composite ultrathin film. The individual SWCN conductivity can be calculated using the (kp) method of the SWCN band-structure theory26 or by other methods. The overall response of the system can be controlled by its intrinsic collective parameters such as the CN volume fraction and the dielectric permittivities of materials involved in addition to the individual constituent SWCN conductivity. The CN fraction depends not only on the CN density but also on the thickness of the dielectric layer.

We also study the thermal broadening and inhomogeneity effects for the CN film dielectric response functions to understand the optical properties of realistic experimental systems at room temperature. For homogeneous single-type CN films, the intertube coupling and thermal broadening bring the exciton and plasmon resonances closer together. They can even overlap due to the inhomogeneous broadening for films composed of (quasi)periodic array mixtures with a slight SWCN diameter distribution. In such systems, the real part of the dielectric response function is negative for a sufficiently broad range of the incident photon energy (negative refraction band), and the CN film behaves as a hyperbolic metamaterial.50 This explains the experimental observations reported recently for horizontally aligned finite-thickness quasi-homogeneous SWCN films.13 Using the MG mixing method,18 we show that for a properly fabricated closely packed two-component inhomogeneous SWCN film, the broadened exciton and plasmon resonances can strongly overlap or coincide in their energies in the way it occurs for the (10,0)/(11,0) CN array mixture, which we here simulate as an example. In such systems, the strong near-field exciton–plasmon coupling and associated hybridization can result in the Rabi-splitting of the exciton absorption lineshape profile, thereby strongly affecting the optical response of the two-component composite film. We show that by varying the relative weights of the two array components, one can tune the optical absorption profile to make the film transmit or absorb light in the neighborhood of an exciton absorption resonance on demand. Additionally, the electrostatic doping could also help control the effect to a certain extent due to the electron density variation and a concurrent scattering process (the Coulomb scattering) in addition to the electron–phonon scattering to result in the line-profile shift and broadening, respectively. As a function of the CN doping level, the peaks in Fig. 4 are expected to shift and broaden with their relative intensity redistributed and the Rabi-splitting in Fig. 5 to increase due to the plasmon resonance width increase, just as suggested by Eqs. (6) and (7).

This research was supported by the U.S. National Science Foundation (NSF) under Condensed Matter Theory Program Award No. DMR-1830874 (I.V.B.)

The data that support the findings of this study are available within the article.

1.
I. V.
Bondarev
, “
Surface electromagnetic phenomena in pristine and atomically doped carbon nanotubes
,”
J. Comput. Theor. Nanosci.
7
,
1673
(
2009
).
2.
A.
Martinez
and
S.
Yamashita
, “Carbon nanotube-based photonic devices: Applications in nonlinear optics,” in Carbon Nanotubes, edited by J. M. Marulanda (IntechOpen, Rijeka, 2011).
3.
M. F. D.
Volder
,
S. H.
Tawfick
,
R. H.
Baughman
, and
A.
Hart
, “
Carbon nanotubes: Present and future commercial applications
,”
Science
339
,
535
(
2013
).
4.
P.
Avouris
,
M.
Freitag
, and
V.
Perebeinos
, “
Carbon-nanotube photonics and optoelectronics
,”
Nat. Photon.
2
,
341
(
2008
).
5.
M.-Y.
Wu
,
J.
Zhao
,
N. J.
Curley
,
T.-H.
Chang
,
Z.
Ma
, and
M. S.
Arnold
, “
Biaxially stretchable carbon nanotube transistors
,”
J. Appl. Phys.
122
,
124901
(
2017
).
6.
T.
Hertel
and
I. V.
Bondarev
, “
Photophysics of carbon nanotubes and nanotube composites
,”
Chem. Phys.
413
,
1
(
2013
).
7.
X.
He
,
W.
Gao
,
L.
Xie
,
B.
Li
,
Q.
Zhang
,
S.
Lei
,
J. M.
Robinson
,
E. H.
Hároz
,
S. K.
Doorn
,
W.
Wang
,
R.
Vajtai
,
P. M.
Ajayan
,
W. W.
Adams
,
R. H.
Hauge
, and
J.
Kono
, “
Wafer-scale monodomain films of spontaneously aligned single-walled carbon nanotubes
,”
Nat. Nanotechnol.
11
,
633
(
2016
).
8.
A. L.
Falk
,
K.-C.
Chiu
,
D. B.
Farmer
,
Q.
Cao
,
J.
Tersoff
,
Y.-H.
Lee
,
P.
Avouris
, and
S.-J.
Han
, “
Coherent plasmon and phonon-plasmon resonances in carbon nanotubes
,”
Phys. Rev. Lett.
118
,
257401
(
2017
).
9.
K.-C.
Chiu
,
A. L.
Falk
,
P.-H.
Ho
,
D. B.
Farmer
,
G.
Tulevski
,
Y.-H.
Lee
,
P.
Avouris
, and
S.-J.
Han
, “
Strong and broadly tunable plasmon resonances in thick films of aligned carbon nanotubes
,”
Nano Lett.
17
,
5641
(
2017
).
10.
P.-H.
Ho
,
D. B.
Farmer
,
G. S.
Tulevski
,
S.-J.
Han
,
D. M.
Bishop
,
L. M.
Gignac
,
J.
Bucchignano
,
P.
Avouris
, and
A. L.
Falk
, “
Intrinsically ultrastrong plasmon-exciton interactions in crystallized films of carbon nanotubes
,”
Proc. Natl. Acad. Sci. U.S.A.
115
,
12662
(
2018
).
11.
M. E.
Green
,
D. A.
Bas
,
H.-Y.
Yao
,
J. J.
Gengler
,
R. J.
Headrick
,
T. C.
Back
,
A. M.
Urbas
,
M.
Pasquali
,
J.
Kono
, and
T.-H.
Her
, “
Bright and ultrafast photoelectron emission from aligned single-wall carbon nanotubes through multiphoton exciton resonance
,”
Nano Lett.
19
,
158
(
2019
).
12.
W.
Gao
,
C. F.
Doiron
,
X.
Li
,
J.
Kono
, and
G. V.
Naik
, “
Macroscopically aligned carbon nanotubes as a refractory platform for hyperbolic thermal emitters
,”
ACS Photonics
6
,
1602
(
2019
).
13.
J. A.
Roberts
,
S. J.
Yu
,
P. H.
Ho
,
S.
Schoeche
,
A. L.
Falk
, and
J. A.
Fan
, “
Tunable hyperbolic metamaterials based on self-assembled carbon nanotubes
,”
Nano Lett.
19
,
3131
(
2019
).
14.
S.
Schoc̈he
,
P.-H.
Ho
,
J. A.
Roberts
,
S. J.
Yu
,
J. A.
Fan
, and
A. L.
Falk
, “
Mid-IR and UV-Vis-NIR Mueller matrix ellipsometry characterization of tunable hyperbolic metamaterials based on self-assembled carbon nanotubes
,”
J. Vac. Sci. Technol. B
38
,
014015
(
2020
).
15.
J. A.
Roberts
,
P.-H.
Ho
,
S.-J.
Yu
,
X.
Wu
,
Y.
Luo
,
W. L.
Wilson
,
A. L.
Falk
, and
J. A.
Fan
, “
Multiple tunable hyperbolic resonances in broadband infrared carbon-nanotube metamaterials
,”
Phys. Rev. Appl.
14
,
044006
(
2020
).
16.
S.
Zhu
,
J.
Ni
, and
Y.
Li
, “
Carbon nanotube-based electrodes for flexible supercapacitors
,”
Nano Res.
13
,
1815
(
2020
).
17.
N.
Komatsu
,
M.
Nakamura
,
S.
Ghosh
,
D.
Kim
,
H.
Chen
,
A.
Katagiri
,
Y.
Yomogida
,
W.
Gao
,
K.
Yanagi
, and
J.
Kono
, “
Groove-assisted global spontaneous alignment of carbon nanotubes in vacuum filtration
,”
Nano Lett.
20
,
2332
(
2020
).
18.
V. A.
Markel
, “
Introduction to the Maxwell Garnett approximation: Tutorial
,”
J. Opt. Soc. Am. A
33
,
1244
(
2016
).
19.
G. D.
Mahan
,
Many-Particle Physics
, 3rd ed. (
Kluwer Academic
,
New York
,
2000
).
20.
C. M.
Adhikari
and
I. V.
Bondarev
, “
Optical response of ultrathin periodically aligned single-wall carbon nanotube films
,”
MRS Adv.
5
,
2685
(
2020
).
21.
I. V.
Bondarev
, “
Finite-thickness effects in plasmonic films with periodic cylindrical anisotropy [Invited]
,”
Opt. Mater. Express
9
,
285
(
2019
).
22.
L. V.
Keldysh
, “
Coulomb interaction in thin semiconductor and semimetal films
,”
JETP Lett.
29
,
658
(
1979
), available at http://www.jetpletters.ac.ru/ps/1458/article_22207.shtml.
23.
I. V.
Bondarev
and
V. M.
Shalaev
, “
Universal features of the optical properties of ultrathin plasmonic films
,”
Opt. Mater. Express
7
,
3731
(
2017
).
24.
I. V.
Bondarev
,
H.
Mousavi
, and
V. M.
Shalaev
, “
Optical response of finite-thickness ultrathin plasmonic films
,”
MRS Commun.
8
,
1092
(
2018
).
25.
I. V.
Bondarev
,
H.
Mousavi
, and
V. M.
Shalaev
, “
Transdimentional epsilon-near-zero modes in planar plasmonic nanostructures
,”
Phys. Rev. Res.
2
,
013070
(
2020
).
26.
T.
Ando
, “
Theory of electronic states and transport in carbon nanotubes
,”
J. Phys. Soc. Jpn.
74
,
777
(
2005
).
27.
J.
Bellessa
,
C.
Bonnand
,
J. C.
Plenet
, and
J.
Mugnier
, “
Strong coupling between surface plasmons and excitons in an organic semiconductor
,”
Phys. Rev. Lett.
93
,
036404
(
2004
).
28.
A. E.
Schlather
,
N.
Large
,
A. S.
Urban
,
P.
Nordlander
, and
N. J.
Halas
, “
Near-field mediated plexcitonic coupling and giant Rabi splitting in individual metallic dimers
,”
Nano Lett.
13
,
3281
(
2013
).
29.
A.
Manjavacas
,
F. J.
Garcia de Abajo
, and
P.
Nordlander
, “
Quantum plexcitonics: Strongly interacting plasmons and excitons
,”
Nano Lett.
11
,
2318
(
2011
).
30.
Y.
Fedutik
,
V. V.
Temnov
,
O.
Schöps
,
U.
Woggon
, and
M. V.
Artemyev
, “
Exciton-plasmon-photon conversion in plasmonic nanostructures
,”
Phys. Rev. Lett.
99
,
136802
(
2007
).
31.
D.
Dai
,
Z.
Dong
, and
J.
Fan
, “
Giant photoluminescence enhancement in SiC nanocrystals by resonant semiconductor exciton–metal surface plasmon coupling
,”
Nanotechnology
24
,
025201
(
2012
).
32.
P.
Vasa
,
R.
Pomraenke
,
S.
Schwieger
,
Y. I.
Mazur
,
V.
Kunets
,
P.
Srinivasan
,
E.
Johnson
,
J. E.
Kihm
,
D. S.
Kim
,
E.
Runge
,
G.
Salamo
, and
C.
Lienau
, “
Coherent exciton–surface-plasmon-polariton interaction in hybrid metal-semiconductor nanostructures
,”
Phys. Rev. Lett.
101
,
116801
(
2008
).
33.
G. A.
Wurtz
,
P. R.
Evans
,
W.
Hendren
,
R.
Atkinson
,
W.
Dickson
,
R. J.
Pollard
,
A. V.
Zayats
,
W.
Harrison
, and
C.
Bower
, “
Molecular plasmonics with tunable exciton–plasmon coupling strength in J-aggregate hybridized Au nanorod assemblies
,”
Nano Lett.
7
,
1297
(
2007
).
34.
N.
Asgari
and
S. M.
Hamidi
, “
Exciton-plasmon coupling in 2D plexitonic nanograting
,”
Opt. Mater.
81
,
45
(
2018
).
35.
I. V.
Bondarev
,
L. M.
Woods
, and
A.
Popescu
,
Exciton-Plasmon Interactions in Individual Carbon Nanotubes
(
Nova Science
,
New York
,
2011
).
36.
I. V.
Bondarev
,
K.
Tatur
, and
L. M.
Woods
, “
Surface exciton-plasmons and optical response of small-diameter carbon nanotubes
,”
Opt. Spectrosc.
108
,
376
(
2010
).
37.
I. V.
Bondarev
,
K.
Tatur
, and
L. M.
Woods
, “
Optical response of small-diameter semiconducting carbon nanotubes under exciton–surface-plasmon coupling
,”
Opt. Commun.
282
,
661
(
2009
).
38.
I. V.
Bondarev
,
L. M.
Woods
, and
K.
Tatur
, “
Strong exciton-plasmon coupling in semiconducting carbon nanotubes
,”
Phys. Rev. B
80
,
085407
(
2009
).
39.
I. V.
Bondarev
and
T.
Antonijevic
, “
Surface plasmon amplification under controlled exciton-plasmon coupling in individual carbon nanotubes
,”
Phys. Stat. Solidi C
9
,
1259
(
2012
).
40.
I. V.
Bondarev
, “
Single-wall carbon nanotubes as coherent plasmon generators
,”
Phys. Rev. B
85
,
035448
(
2012
).
41.
I. V.
Bondarev
and
P.
Lambin
, “
Vacuum-field Rabi oscillations in atomically doped carbon nanotubes
,”
Phys. Lett. A
328
,
235
(
2004
).
42.
I. V.
Bondarev
and
A. V.
Meliksetyan
, “
Possibility for exciton Bose-Einstein condensation in carbon nanotubes
,”
Phys. Rev. B
89
,
045414
(
2014
).
43.
I. V.
Bondarev
and
A.
Popescu
, “
Exciton Bose-Einstein condensation in double walled carbon nanotubes
,”
MRS Adv.
2
,
2401
(
2017
).
44.
W.
Zhang
,
A. O.
Govorov
, and
G. W.
Bryant
, “
Semiconductor-metal nanoparticle molecules: Hybrid excitons and the nonlinear Fano effect
,”
Phys. Rev. Lett.
97
,
146804
(
2006
).
45.
A. M.
Dibos
,
Y.
Zhou
,
L. A.
Jauregui
,
G.
Scuri
,
D. S.
Wild
,
A. A.
High
,
T.
Taniguchi
,
K.
Watanabe
,
M. D.
Lukin
,
P.
Kim
, and
H.
Park
, “
Electrically tunable exciton–plasmon coupling in a WSe2 monolayer embedded in a plasmonic crystal cavity
,”
Nano Lett.
19
,
3543
(
2019
).
46.
E.
Cao
,
W.
Lin
,
M.
Sun
,
W.
Liang
, and
Y.
Song
, “
Exciton-plasmon coupling interactions: From principle to applications
,”
Nanophotonics
7
,
145
(
2018
).
47.
L. X.
Benedict
,
S. G.
Louie
, and
M. L.
Cohen
, “
Static polarizabilities of single-wall carbon nanotubes
,”
Phys. Rev. B
52
,
8541
(
1995
).
48.
B.
Kozinsky
and
N.
Marzari
, “
Static dielectric properties of carbon nanotubes from first principles
,”
Phys. Rev. Lett.
96
,
166801
(
2006
).
49.
S.
Tasaki
,
K.
Maekawa
, and
T.
Yamabe
, “
π-band contribution to the optical properties of carbon nanotubes: Effects of chirality
,”
Phys. Rev. B
57
,
9301
(
1998
).
50.
I. V.
Bondarev
and
C. M.
Adhikari
, “Collective excitations and optical response of ultrathin carbon nanotube films,” arXiv:2011.11216 (2020).
51.
I. V.
Bondarev
and
P.
Lambin
, “
van der Waals coupling in atomically doped carbon nanotubes
,”
Phys. Rev. B
72
,
035451
(
2005
).
52.
I. V.
Bondarev
, “
Asymptotic exchange coupling of quasi-one-dimensional excitons in carbon nanotubes
,”
Phys. Rev. B
83
,
153409
(
2011
).
53.
A. E.
Nikolaenko
,
F.
De Angelis
,
S. A.
Boden
,
N.
Papasimakis
,
P.
Ashburn
,
E.
Di Fabrizio
, and
N. I.
Zheludev
, “
Carbon nanotubes in a photonic metamaterial
,”
Phys. Rev. Lett.
104
,
153902
(
2010
).
54.
I. V.
Bondarev
, “
Plasmon enhanced Raman scattering effect for an atom near a carbon nanotube
,”
Opt. Express
23
,
3971
(
2015
).
55.
T. G.
Pedersen
, “
Exciton effects in carbon nanotubes
,”
Carbon
42
,
1007
(
2004
).
56.
T. G.
Pedersen
, “
Variational approach to excitons in carbon nanotubes
,”
Phys. Rev. B
67
,
073401
(
2003
).
57.
R. B.
Capaz
,
C. D.
Spataru
,
S.
Ismail-Beigi
, and
S. G.
Louie
, “
Diameter and chirality dependence of exciton properties in carbon nanotubes
,”
Phys. Rev. B
74
,
121401
(
2006
).
58.
F.
Wang
, “
The optical resonances in carbon nanotubes arise from excitons
,”
Science
308
,
838
(
2005
).