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

Plasmon enhanced Raman scattering effect for an atom near a carbon nanotube

Open Access Open Access

Abstract

Quantum electrodynamics theory of the resonance Raman scattering is developed for an atom in a close proximity to a carbon nanotube. The theory predicts a dramatic enhancement of the Raman intensity in the strong atomic coupling regime to nanotube plasmon near-fields. This resonance scattering is a manifestation of the general electromagnetic surface enhanced Raman scattering effect, and can be used in designing efficient nanotube based optical sensing substrates for single atom detection, precision spontaneous emission control, and manipulation.

© 2015 Optical Society of America

1. Introduction

Surface-enhanced Raman spectroscopy (SERS) has received much of attention recently due to a very broad range of its applications ranging from optics and plasmonics to biochemistry and medicine [13]. High scattering intensities within narrow spectral bands reduce the probability for spectral overlapping to allow for better recognition of multiple markers, making SERS one of the most efficient optical sensing techniques. With the development of advanced nanomaterials, various SERS substrates are demonstrated [410]. However, there is still a need for inexpensive substrates of improved sensitivity and signal reproducibility, which require clear understanding of the underlying scattering mechanisms to be developed.

In general, the SERS effect originates from the resonance increase of the induced transition dipole moment of an atomic or molecular scatterer when positioned in the near-surface zone of a metallic structure. This can be for two reasons. They are: (i) due to quasi-static electric fields associated with resonance plasmon excitation in metallic structures, and/or (ii) due to the electron polarizability increase associated with the charge transfer between the substrate and the scatterer. It is generally agreed to distinguish between the electromagnetic (EM) and chemical SERS effects, accordingly [13].

Most of the applications of carbon nanotubes (CNs) to increase the Raman scattering signal have been to decorate them with metallic nanoparticles [8, 9], in order to obtain local EM field enhancement generated by the spatially confined plasmon modes of the nanoparticles with CNs only used as a network to support the particles. Only very recently, the chemical SERS effect with CNs alone, without any combination with metallic nanoparticles, was first reported by Andrada et al. in [10], where molecules covalently bound to single wall CNs demonstrated the resonantly increased Raman signal that was even stronger than that of the nanotube itself.

In this article, a quantum electrodynamics (QED) theory of the resonance Raman scattering is presented for an atom near a carbon nanotube, to demonstrate that individual CNs are capable of providing the electromagnetic SERS effect as well. Nanotubes offer extraordinary stability, flexibility and precise tunability of their EM properties on-demand by simply varying their diameters and/or chiralities. CNs of different diameters and chiralities feature similar electronic band structure peculiarities, yet shifted in frequency relative to one another [1113]. This yields similar EM properties over a broad range of excitation frequencies both in the far- and in the near-field zone, originating from exciton and plasmon excitations, respectively. Excitons and plasmons are different in their physical nature, but originate from the same circumferentially quantized electronic transitions. Due to the circumferential quantization of the longitudinal electron motion, real axial (along the CN axis) optical conductivities of single wall CNs consist of series of peaks E 11 ,E 22 ,…, representing the 1st, 2nd, etc. excitons, respectively [see Fig. 1(a)]. Imaginary conductivities are linked with the real ones by the Kramers-Kronig relation, and so real inverse conductivities show the resonances P 11 ,P 22 ,… [Fig. 1(a)] next to their excitonic counterparts. These are inter-band plasmons that were theoretically demonstrated quite recently to play the key role in a variety of new interesting surface EM phenomena with CNs [1626], including exciton-plasmon coupling [16, 17] and plasmon generation by excitons [18, 19], exciton Bose-Einstein condensation in individual single wall CNs [20], Casimir attraction in double wall CNs [17, 21, 22], resonance optical absorption [23] and atomic entanglement in hybrid systems of extrinsic atoms/ions doped into CNs [2426], to mention a few, — all of direct relevance to conceptually new tunable optoelectronic device applications with carbon nanotubes [27, 28]. Experimental evidence for these low-energy (~1–2 eV) weakly-dispersive plasmon modes in CNs was first reported by Pichler et al. in [29]. Inter-band plasmons are standing charge density waves due to the periodic opposite-phase displacements of the electron shells with respect to the ion cores in the neighboring elementary cells of the CN [18, 19]. When excited, their (plasmon-induced) quasi-static electric fields can be strong enough to result in the enhanced Raman scattering effect by atomic type species (extrinsic atoms, ions, molecules, or semiconductor quantum dots) in the CN vicinity. This work derives and analyzes the differential cross-section for such scattering.

 figure: Fig. 1

Fig. 1 (a) Fragment of the energy dependence of the dimensionless (normalized by e2/2πh¯) axial surface conductivities σzz for the semiconducting (6,4), (10,0) and (11,0) nanotubes of increasing diameter. Peaks of Reσzz represent excitons (E 11, E 22, …); peaks of Re(1zz) are inter-band plasmons (P 11, …). (b) Photonic DOS functions for the CNs in (a) with the TLS placed at the distance rA=Rcn +2b (see inset). Dimensionless energy is [Energy]/2γ 0. Conductivities are obtained using the (k · p)-scheme developed by Ando [12]. DOS functions are calculated as described by Bondarev and Lambin in [14, 15]. See text for notations.

Download Full Size | PDF

2. The Hamiltonian

In absence of external EM radiation, an atom, modeled here by a two-level system (TLS) positioned at the point r A near an infinitely long single wall CN, interacts with the quantum EM field of the CN via an electric transition dipole moment dz=u|d^z|l between the TLS lower and upper states, |l〉 and |u〉, respectively, with the z-quantization axis being the CN symmetry axis [Fig. 1(b), inset]. Transverse dipole orientations can be neglected due to the strong transverse depolarization effect in individual CNs [3034]. The full QED second quantized Hamiltonian for such a coupled CN–TLS quantum system was earlier formulated by Bondarev and Lambin in [14, 15] to have the following form

H^=H^F+H^A+H^AF=0dωh¯ωdRf^(R,ω)f^(R,ω)+h¯ω˜A2σ^Z+0dωdR[g(+)(rA,R,ω)σ^g()(rA,R,ω)σ^]f^(R,ω)+h.c.,
with the three terms representing the (medium-assisted) quantum EM field of the CN, the TLS, and their interaction, respectively. Here, f^(R,ω) and f^(R,ω) are the scalar bosonic field operators that create and annihilate, respectively, surface EM excitations of frequency ω in the CN field subsystem, R = (Rcn,φ,z) is the radius-vector of a point on the CN surface. Pauli operators, σ^z=|uu||ll|, σ^=|lu| and σ^=|ul|, describe the TLS and its electric dipole transitions between the two states, upper |u〉 and lower |l〉, with the transition frequency ωA modified by the diamagnetic Â2-term (vector potential) to result in the new renormalized transition frequency
ω˜A=ωA[12(h¯ωA)20dωdR|g(rA,R,ω)|2].

The matrix elements of the CN field interaction with the TLS are of the form

g(±)(rA,R,ω)=g(rA,R,ω)±ωωAg(rA,R,ω),g()(rA,R,ω)=i4ωAc2dzπh¯ωReσzz(ω)()Gzz(rA,R,ω),
with σzz(ω) being the CN surface axial conductivity [Fig. 1(a)], and
()Gzz(rA,R,ω)=drδzz()(rAr)Gzz(r,R,ω)
representing the zz-component of the transverse (longitudinal) Green tensor (with respect to the first variable) of the CN assisted quantum field. Here,
δαβ(r)=αβ14π|r|,δαβ(r)=δαβδ(r)δαβ(r)
are the longitudinal and transverse dyadic δ-functions, respectively. The Green tensor components are given by the solutions to the equation
α=r,φ,z(××ω2c2)zαGαz(r,R,ω)=δ(rR),
to fulfill the radiation conditions at infinity and the boundary conditions on the CN surface. This tensor was derived and analyzed by Bondarev and Lambin in [34].

Functions g(±)(r A, R ) have the property as follows

dR|g(±)(rA,R,ω)|2=h¯22πΓ0(ω)[ξ(rA,ω)+ωA2ω2ξ(rA,ω)],
where Γ0(ω)=8πω2dz2ImGzz0(ω)/h¯c2 is the TLS spontaneous decay rate in vacuum with ImGzz0(ω)=ω/6πc being the vacuum imaginary Green tensor zz-component, and
ξ()(rA,ω)=Im()Gzz()(rA,rA,ω)ImGzz0(ω)
representing the transverse (longitudinal) photonic density of states (DOS) relative to vacuum as seen from the TLS location r A,
()Gzz()(rA,rA,ω)=drdrδzz()(rAr)Gzz(r,r,ω)δzz()(rrA)
is the Green tensor zz-component that is transverse (longitudinal) with respect to both variables.

Hamiltonian (1) involves only two standard approximations, the electric dipole and two-level approximation [14, 15], while conveniently representing the coupled TLS–CN system in terms of the relative distance dependent DOS functions ξ ⊥(||)(r A). For short TLS–CN separation distances EM retardation effects play no role [15], and so one has ξ = ξ||= ξ(r A) for the DOS functions in Eq. (2). Figure 1(b) shows examples of ξ (rA,x) as functions of the dimensionless energy x=h¯ω/2γ0, where γ 0 =2.7 eV is the C-C overlap integral, calculated for the three semiconducting CNs of increasing diameter with the TLS positioned at rA = Rcn+ 2b, b=1.42 Å being the C-C interatomic distance. Details of these calculations and similar graphs for other CN–TLS separation distances can be found in [14, 15, 34]. We see the sharp single-peak resonances originating from the inter-band plasmons of the respective CNs [cf. Figs. 1(a) and 1(b)]. These are responsible for the CN–TLS coupling in the near-field. The coupling is due to the virtual (vacuum-type) EM energy exchange between the TLS and the CN to create and annihilate plasmon excitations on the CN surface with the TLS de-excited and excited, respectively, as described by the second line in the Hamiltonian (1).

3. Eigen states spectrum

To proceed with the Raman scattering cross-section calculations, it is necessary to determine the spectrum of the eigen states of the Hamiltonian (1). In the linear coupling regime, quite generally, the coupled CN–TLS system can be represented as a four-level system with the eigenvectors of the Hamiltonian (1) of the form

|0=|l|{0},|1,2=Cu(1,2)|u|{0}+0dωdRCl(1,2)(R,ω)|l|{1(R,ω)},|3=|u|{1(R,ω)}.

Here, |{0}〉 and |{1(R )}〉 are, respectively, the vacuum and single-quantum excited states of the CN field subsystem, and Cu,l(1,2) are unknown mixing coefficients for the non-radiative spontaneous decay transition |u〉|{0}〉 →|l〉|{1(R )〉 to excite one plasmon of frequency ω at point R of the CN surface with simultaneous de-excitation of the TLS [15,34]. These mixing coefficients can be found by solving the eigenvalue problem for the Hamiltonian (1) in the basis (3). Similar mixing of the |l〉|{0}〉 and |u〉|{1(R )}〉 states, known to be responsible for the long-range dispersive van der Waals interaction [14, 15], is neglected here for simplicity.

Solving the eigenvalue problem for the Hamiltonian (1) in the basis (3), one obtains the energy eigenvalues

E0=h¯ω˜A2,E3=h¯ω˜A2+h¯ω
for the eigenvectors |0〉 and |3〉, respectively, and the simultaneous equation set for the mixing coefficients as follows
{(h¯ω˜A2E)Cu(1,2)+0dωdRg(+)(rA,R,ω)Cl(1,2)(R,ω)=0,[g(+)(rA,R,ω)]*Cu(1,2)+(h¯ω˜A2+h¯ωE)Cl(1,2)(R,ω)=0.

Here, the second equation gives

Cl(1,2)(R,ω)=[g(+)(rA,R,ω)]*h¯ω˜A/2h¯ω+ECu(1,2)
which, being inserted into the first one, results in the integral equation
E=h¯ω˜A2+0dωdR|g(+)(rA,R,ω)|2h¯ω˜A/2h¯ω+E=h¯ω˜A2+h¯22π0dωΓ0(ω)(1+ωA2/ω2)ξ(rA,ω)h¯ω˜A/2h¯ω+E
to give the energy eigenvalues E 1 , 2 for the eigenvectors |1,2〉. Here, the second line was obtained by using Eq. (2) with ξ = ξ|| = ξ (r A) on assumption of negligible EM retardation effects at short TLS–CN separation distances [15]. Taking advantage of the sharp single peak structure of the DOS function ξ (r A) [cf. Figs. 1(a) and 1(b)] in the vicinity of the plasmon resonance frequency ωp [35], one can use the Lorentzian approximation of the half-width-at-half-maximum Δω 0 of the form ξ(rA,ω)ξ(rA,ωp)Δω02/[(ωωp)2+Δω02] to solve Eq. (7) analytically. One has
0dωΓ0(ω)(1+ωA2/ω2)ξ(rA,ω)h¯ω˜A/2h¯ω+EΓ0(ωp)(1+ωA2/ωp2)ξ(rA,ωp)Δω02h¯ω˜A/2h¯ωp+E0dω(ωωp)2+Δω02,
where the integral calculates to give [arctan(ωpω 0) +π/2]/Δω 0, yielding π/Δω 0 with the arctan function expanded to linear terms in Δω 0 p (≪1, and the stronger this inequality is, the better such a series expansion works). Equation (7) now becomes a simple quadratic equation to bring one, along with Eq. (4), to the complete energy eigenvalue set of the problem as follows
ε0=x˜A2,ε1,2=12(xpδ2+X2iΔxp),ε3=x˜A2+xpiΔxp.

Here, εi = Ei/2γ 0 with i=0,1,2,3 and (x˜A,xp,Δxp)=h¯(ω˜A,ωp,Δωp)/2γ0 are dimensionless energies, δ=x˜Axp, X=(h¯/2γ0)[2Δω0Γ0(ωp)(1+ωA2/ωp2)ξ(rA,ωp)]1/2, and Δxp is added to phenomenologically account for the finite half-width of the plasmon resonance (finite plasmon lifetime), as seen in Fig. 1(a), which is assumed to be much broader than the excited atomic level natural half-width dropped here on this account for simplicity.

The mixing coefficients in the eigenvectors |1〉 and |2〉 in Eq. (3) can be found straightforwardly by using the normalization condition

|Cu(1,2)|2+0dωdR|Cl(1,2)(R,ω)|2=1,
and substituting Eq. (6) in it with E replaced by E 1 and E 2, as given by Eq. (8), for Cl(1) and Cl(2) respectively, followed by the integral evaluation within the same Lorentzian approximation for the DOS function ξ (r A) that was used to evaluate the integral in Eq. (7). This results in
Cu(1,2)=[12(1+11+X2/δ21+X2/δ21+X2/δ2)]1/2,0dωdR|Cl(1,2)(R,ω)|2=1|Cu(1,2)|2=(X2/2δ2)|Cu(1,2)|21+X2/2δ21+X2/δ2.

Equations (3), Eq. (8) and (9) represent the complete solution to the eigenvalue problem for the Hamiltonian (1) of the coupled CN–TLS system in absence of external EM radiation. They are valid both in resonance, where δ ~0 and so X 2 2≫1, and out of resonance where X 2 2≪1, and give different easily derivable asymptotical expressions in these two regimes of relevance to strong and weak CN–TLS coupling, respectively. The eigen energy level structure given by Eq. (8) is sketched in Fig. 2. We see the anti-crossing behavior of the eigen energy levels ε 1 and ε 2, a characteristic of the strong coupling regime, that is controlled by the parameters X and Δxp. When in resonance, the actual coupling regime, strong or weak, depends on the relation between X and Δxp. The CN–TLS system will only be coupled strongly if the parameter X, which plays the role of the vacuum-field Rabi splitting here [15, 23, 34], is much greater than the plasmon resonance broadening Δxp as shown in Fig. 2. When X ≫ Δxp, the Rabi splitting of the levels ε 1 and ε 2 is not hidden by the plasmon resonance broadening, and so the strong CN–TLS coupling regime is realized. Otherwise, if X ≪ Δxp, the levels ε 1 and ε 2 are smeared, showing no clear anti-crossing behavior, and so no strong coupling can be realized.

 figure: Fig. 2

Fig. 2 Schematic of the energy level structure as given by Eq. (8) for the coupled four-level CN–TLS system. Thick red lines show the eigen energy levels as functions of xp. Thin red dashed lines indicate their broadening due to finite Δxp. Horizontal dotted lines are to show Rabi-splitting and the in-resonance strong-coupling solutions given by Eq. (8) with δ = 0.

Download Full Size | PDF

4. Raman scattering cross-section

Under the assumption that the coupled CN–TLS system with the eigen states (3), (8) and (9) is initially in the ground state, the inelastic scattering of external EM radiation by this system only involves transitions between levels |0〉, |1〉 and |2〉, as shown in Fig. 3, top, due to the dipole moment selection rule restrictions. The entire scattering process includes three sequential steps. They are:

  1. excitation of the system by an incident photon of the frequency ωi with the unit polarization vector e i, described by the interaction matrix element
    n|H^R(ωi)|0=ic2πh¯ωidzcosϑiCu(n)*,cosϑi=ei·ez,n=1,2
    (normalized at one photon per unit volume [36]);
  2. plasmon emission (or absorption) on the CN surface, described by the matrix element
    1|H^AF(e)|2=0dωdR[Cl(1)(R,ω)g(+)(rA,R,ω)]*Cu(2)
    [or 2|H^AF(a)|1=1|H^AF(e)|2 for absorption], with H^AF(e) a and H^AF(e) being the ( ~f^) and the absorption term ( ~f^), respectively, of the interaction Hamiltonian H^AF in Eq. (1);
  3. de-excitation of the CN–TLS system by means of the scattered (Raman) photon emission of the frequency ωs with the unit polarization vector e s, described by the interaction matrix element n|H^R(ωi)|0|is in accordance with Eq. (10).

 figure: Fig. 3

Fig. 3 Schematic of the Raman scattering process (top) in terms of the inter-level transitions (levels sketched in Fig. 2) of the coupled CN–TLS system given by Eqs. (3), (8) and (9), and the Feynman diagrams (bottom) for the scattering cross-section calculations.

Download Full Size | PDF

There are four Feynman diagrams, shown in Fig. 3, bottom, to contribute to this three-step process. They are two for plasmon emission (bottom left) and two for plasmon absorption (bottom right), to represent two indistinguishable ways for emission and absorption to occur. Two types of the emission (absorption) diagrams should be summed up and squared, followed by adding the emission and absorption contributions together [37], to result in the Fermi Golden Rule transition rate

(2πh¯)|0|H^R(ωs)|11|H^AF(e)|22|H^R(ωi)|0[h¯ωih¯ωp(E1E0)][h¯ωi(E2E0)]+0|H^R(ωi)|11|H^AF(e)|22|H^R(ωs)|0[h¯ωsh¯ωp(E1E0)][h¯ωs(E2E0)]|2δ(h¯ωih¯ωph¯ωs)+|0|H^R(ωs)|22|H^AF(a)|11|H^R(ωi)|0[h¯ωi+h¯ωp(E2E0)][h¯ωi(E1E0)]+0|H^R(ωi)|22|H^AF(a)|11|H^R(ωs)|0[h¯ωs+h¯ωp(E2E0)][h¯ωs(E1E0)]|2δ(h¯ωi+h¯ωph¯ωs).

Matrix elements in here can be consistently evaluated within the Lorentzian approximation for the DOS function ξ (r A). Substituting Cl(1) out of Eq. (6), with E replaced by E 1 per Eq. (8), into Eq. (11) and performing exactly the same integral evaluation as was done in Eq. (7), one has

1|H^AF(e)|2=2γ0Cu(1)*Cu(2)(X2/4)x˜A/2xp+ε1=2|H^AF(a)|1,
yielding in view of Eq. (10)
0|H^R|11|H^AF(e)|22|H^R|0=(0|H^R|22|H^AF(a)|11|H^R|0)=2πh¯ωiωsc2dz2cosϑicosϑs2γ0|Cu(1)Cu(2)|2(X2/4)x˜A/2xp+ε1
with
|Cu(1)Cu(2)|2=X2/4δ2+X2,
according to Eq. (9).

To obtain the differential scattering cross-section, one has to multiply Eq. (12) by the density of final states (h¯ωs)2d(h¯ωs)dΩs/(2πh¯)3 for photons scattered within the solid angle dΩs, and then integrate it over h¯ωs

, the scattered radiation energy [36]. With Eqs. (8), (13) and (14), this eventually results in the differential Raman scattering cross-section as follows

dσdΩs=(2γ0)2|dz|4h¯4c4cos2ϑicos2ϑsP(xi,xs),
with the dimensionless (angle-free) scattering probability function
P(xi,xs)=xixs3A(δ,X,Δxp){1[(xixpδ+/2)2+Δxp2][(xsxpδ/2)2+Δxp2]+1[(xixpδ/2)2+Δxp2][(xsxpδ+/2)2+Δxp2]},
where xi,s=h¯ωi,s/2γ0,
A(δ,X,Δxp)=X826(δ2+X2)2(δ2+Δxp2),

δ±=δ±δ2+X2, and only the resonant terms are left to represent the contributions from plasmon emission and absorption, respectively, while (insignificant, quasi-constant back ground) non-resonant terms of the transition rate (12) are dropped for brevity.

5. Discussion

Each term in Eq. (16) has the product of two resonance energy denominators that include xi and xs, incident (incoming) and scattered (outgoing) photon energies. This is what makes the Raman scattering cross-section (15) resonant. In addition to that, there is an important pre-factor there, A(δ, X, Δxp), given by Eq. (17). This comes from Eq. (13), which can be viewed as [dzEz(ωi)][dzEz(ωs)][dzEz(loc)(rA)]2 with dzEz(loc)(rA)~X[Γ0(ωp)ξ(rA,ωp)]1/2, thus bringing the local-field enhancement factor [dzEz(loc)(rA)]4ξ2(rA,ωp) into the Raman cross-section due to plasmon generated quasi-static electric fields (see [18,19] for more details) at the TLS location r A when in the CN near-surface zone rA~RCN [Fig. 1(b)]. This factor is significant when δ ~0 and X ≫ Δxp simultaneously, a regime whereby the CN–TLS system couples strongly by means of the virtual EM energy exchange between the TLS and the nanotube, corresponding to non-exponential spontaneous decay dynamics with Rabi oscillations of the TLS excited state [15, 34], and Rabi splitting of the TLS optical absorption line profile [23].

Maximum Raman intensity is controlled by the ratio A(δ,X,Δxp)/Δxp4 as can be seen from Eq. (16). This is a slightly asymmetric function of δ peaked at δX/3 when X ≫ Δxp, not at δ =0 as one would expect, which can be easily shown by testing it for maximum analytically. Bondarev and Vlahovic have shown earlier in [23] by analyzing the absorption line shape profile that X, the Rabi splitting, may typically be as large as ~0.01–0.1, as, whereas Δxp ~0.005–0.01 can be seen from Fig. 1(a). Figure 4, left and right, shows the enhancement factor A/Δxp4 as a function of δ and X at Δxp = 0.005, and as a function of δ and Δxp at X = 0.05, respectively. We see that as long as X ≫ Δxp, whereby the CN–TLS system is in the strong coupling regime, the maximum intensity decreases by a factor ~ 1.5 within the detuning window ~0.08, thus providing the spectral band as large as 0.08×2γ 0 .~ 0.43 eV for a significant plasmon enhanced Raman scattering effect to occur. This spectral band can be shifted both to the red and to the blue by shifting the inter-band plasmon resonance energy, as can be seen in Fig. 1, which can merely be achieved by using CNs with different diameters and/or chiralities, thereby offering the flexibility and precise tunability in designing CN based SERS substrates with parameters required on-demand.

 figure: Fig. 4

Fig. 4 Enhancement factor A(δ,X,Δxp)/Δxp4 as given by Eq. (17) for Δxp = 0.005 (left) and for X =0.05 (right) to show the influence of the detuning δ(=x˜Axp) on the maximum intensity of the plasmon enhanced Raman scattering effect.

Download Full Size | PDF

Returning back to Eq. (16) we see that for each xi = xp +δ±/2 only one term contributes, resulting in either Stokes scattering with xs=xp+δ/2<xi=xp+δ +/2 and a plasmon created in the CN, or in anti-Stokes scattering with xs=xp+δ +/2 > xi=xp+δ/2 and a plasmon absorbed from the CN. The absolute value of the Raman shift is (δ+δ)/2=δ2+X2, yielding a quantity ∼X in resonance, where δ ∼ 0 whereby X 2 2≫1, and that ∼δ out of resonance with X 2 2≪1. In the latter case, A(δ≫X, Δxp) ∝X 4 4, according to Eq. (17), totally ruling out the probability P(xi,xs) of the scattering process. When in resonance, on the other hand, the P(xi,xs) maximum value goes as A(0,X,Δxp)/Δxp4(X4/Δxp4)/(X2+Δxp2), being strongly suppressed under the weak CN–TLS coupling where X2/Δxp21, and being dramatically enhanced in the case where X2/Δxp21 so that the CN–TLS coupling is strong. The scattering enhancement factor is about square of that for resonance absorption by atomically doped CNs [23], the way it should be for scattering as a two-step process of absorption followed by emission (viewed as ”reversed absorption”).

Figure 5 shows an example of the numerical calculations for the scattering probability P(xi,xs) as given by Eq. (16) for xp=0.35[P11(6,4) plasmon in Fig. 1(a), corresponding to the energy 0.35×2γ 0 =1.89 eV (red spectral line)] with X and Δxp being varied independently [rows (a) and (b), respectively], to see the role of the DOS resonance variation due to the local field enhancement/dehancement effect as the TLS–CN separation distance changes and the plasmon decoherence effect, respectively. As discussed above, Raman scattering is seen to be very sensitive to the strong CN–TLS coupling, blowing up by a factor of over 103 for X/Δxp ~10 and totally vanishing when X/Δxp ~1. Raising X increase both the Raman shift and the intensity, while greater Δxp quench the intensity with no Raman shift change.

 figure: Fig. 5

Fig. 5 Raman scattering probability as a function of the incident xi and scattered xs photon energies as given by Eq. (16) with δ=X/3 (maximum of A/Δxp4 in Fig. 4, see text) for xp=0.35[P11(6,4) plasmon in Fig. 1(a)], with X and Δxp being varied independently [rows (a) and (b)]. TLS–plasmon coupling strength is represented by the ratio X/Δxp being greater or less than unity for strong and weak coupling, respectively. Raman scattering is seen to be manifestly indicative of the strong TLS–plasmon coupling, dramatically increasing as X/Δxp goes much greater than unity and disappearing when it is comparable with unity.

Download Full Size | PDF

The QED theory of the plasmon enhanced Raman scattering developed here applies to chemically inactive atoms, ions, or even organic molecules and semiconductor quantum dots that are physisorbed on the nanotube walls, whereby there is no local electronic orbitals hybridization between the nanotube and the atomic transition levels involved. The theory can be tested experimentally by using rear-earth ion complexes, Eu3+ ions, in particular [3840]. These are known to be excellent probes for near-field effects in spatially confined systems, owing to the dominant narrow, easily detectible 5D 0 7F 2 electric dipole transition of the wavelength ~614 nm between two deep-lying electronic levels (f-shell) of europium that essentially create an ideal TLS. Corresponding transition energy is 2.02 eV, falls within the 0.43 eV spectral band of the first inter-band plasmon resonance P11(6,4) of the (6, 4) nanotube (see Fig. 1, xp=0.35 corresponding to Ep =1.89 eV) whose calculated Raman spectrum is shown in Fig. 5 and was discussed above. Pre-alignment of europium doped CNs is desirable to facilitate the excitation efficiency, but is not crucially important.

6. Conclusion

In this article, the QED theory of the resonance Raman scattering is developed for a two-level dipole emitter — TLS coupled to a weakly-dispersive low-energy (~1–2 eV) inter-band plasmon resonance of a carbon nanotube. The theory applies to atomic type species such as atoms, ions, molecules, or semiconductor quantum dots that are physisorbed on the nanotube walls. The analytical expression derived for the Raman cross-section covers both weak and strong TLS–plasmon coupling, and shows dramatic enhancement in the strong coupling regime. Such resonance scattering is a manifestation of the general electromagnetic SERS effect, in which the enhancement is due to the plasmon-induced near-fields that affect the TLS in a close proximity to the CN surface, given that the TLS transition energy is within the spectral band of ~0.43 eV of the corresponding inter-band plasmon resonance energy of the nanotube.

The QED scattering model used here belongs to a broad class of driven four-level quantum systems analyzed by Delgado and Gomez Llorente in [41]. There is an important distinction though, that instead of being driven by the strength of the coupling to an external periodic field, the scattering properties of the four-level CN–TLS system considered here are controlled by the strength of the (vacuum-type, intrinsic) coupling of the TLS to the plasmon-induced quasi-static electric fields in the near-surface zone of a carbon nanotube. Similar models are used to describe the peculiarities of EM interaction processes in a variety of quantum systems. These include both cavity-type systems such as atoms in optical traps [42], matter excitations in one-dimensional [4345] or zero-dimensional [46, 47] nanostructures embedded in solid-state nanocavities, and general molecular [48] and semiconductor systems [37]. In the latter case, the phenomenon of the resonance Raman scattering originates from lambda-type transitions, such as vibrational transitions of a molecule or electron band transitions in a semiconductor, within a subset of states of the same system. In the case considered here, the resonance scattering effect becomes possible due to the formation of the hybridized (”dressed”) states when two different systems, the TLS and the CN, enter the strong coupling regime.

This theoretical work provides a unified description of the near-field plasmon enhancement effects. The work will help establish new design concepts for future generation CN based nanophotonics platforms for single molecule/atom/ion detection, precision spontaneous emission control, and optical manipulation, which will benefit from the extraordinary stability, flexibility and precise tunability of the electromagnetic properties of carbon nanotubes by means of their diameter and chirality variation.

Acknowledgments

This work is supported by DOE ( DE-SC0007117). I. V. B. acknowledges hospitality of Munich Advanced Photonics Center at TU-Minuch, Germany, where this work was started, as well as fruitful discussions with its staff members, Prof. W. Domcke and Dr. M. Gelin.

References and links

1. K. Kneipp, M. Moscovits, and H. Kneipp, Surface-Enhanced Raman Scattering: Physics and Applications (Springer-Verlag, 2006). [CrossRef]  

2. A. Otto, “The chemical (electronic) contribution to surface enhanced Raman scattering,” J. Raman Spectrosc. 36, 497–509 (2005). [CrossRef]  

3. R. Zhang, Y. Zhang, Z. C. Dong, S. Jiang, C. Zhang, L. G. Chen, L. Zhang, Y. Liao, J. Aizpurua, Y. Luo, J. L. Yang, and J. G. Hou, “Chemical mapping of a single molecule by plasmon-enhanced Raman scattering,” Nature 498, 82–86 (2013). [CrossRef]   [PubMed]  

4. M. Peng, H. Xu, and M. Shao, “Ultrasensitive surface-enhanced Raman scattering based gold deposited silicon nanowires,” Appl. Phys. Lett. 104, 193103 (2014). [CrossRef]  

5. Q. Hao, S. M. Morton, B. Wang, Y. Zhao, L. Jensen, and T. J. Huang, “Tuning surface-enhanced Raman scattering from graphene substrates using the electric field effect and chemical doping,” Appl. Phys. Lett. 102, 011102 (2013). [CrossRef]  

6. R. Lv, Q. Li, A. R. Botello-Mendez, T. Hayashi, B. Wang, A. Berkdemir, Q. Hao, A. L. Elias, R. Cruz-Silva, H. R. Gutierrez, Y. A. Kim, H. Muramatsu, J. Zhu, M. Endo, H. Terrones, J.-C. Charlie, M. Pan, and M. Terrones, “Nitrogen-doped graphene: beyond single substitution and enhanced molecular sensing,” Sci. Rep. 2, 586 (2012). [CrossRef]   [PubMed]  

7. D. Z. Lin, Y. P. Chen, P. J. Jhuang, J. Y. Chu, J. T. Yeh, and J.-K. Wang, “Optimizing electromagnetic enhancement of flexible nano-imprinted hexagonally patterned surface-enhanced Raman scattering substrates,” Opt. Express 19, 4337–4345 (2011). [CrossRef]   [PubMed]  

8. Y.-C. Chen, R. J. Young, J. V. Macpherson, and N. R. Wilson, “Silver-decorated carbon nanotube networks as SERS substrates,” J. Raman Spectrosc. 42, 1255–1256 (2011). [CrossRef]  

9. Y. Sun, K. Liu, J. Miao, Z. Wang, B. Tian, and L. Zhang, “Highly sensitive surface-enhanced Raman scattering substrate made from superaligned carbon nanotubes,” NanoLett. 10, 1747–1753 (2010). [CrossRef]  

10. D. M. Andrada, H. S. Vieira, M. M. Oliveira, A. P. Santos, L. Yin, R. Saito, M. A. Pimenta, C. Fantini, and C. A. Furtado, “Dramatic increase in the Raman signal of functional groups on carbon nanotube surfaces,” Carbon 56, 235–242 (2013). [CrossRef]  

11. R. Saito, G. Dresselhaus, and M. S. Dresselhaus, Science of Fullerens and Carbon Nanotubes (Imperial College, 1998).

12. T. Ando, “Theory of electronic states and transport in carbon nanotubes,” J. Phys. Soc. Jpn. 74, 777–817 (2005). [CrossRef]  

13. T. Hertel and I. V. Bondarev, eds., Photophysics of Carbon Nanotubes and Nanotube Composites (Special Issue), Chem. Phys. 413, 1–131 (2013). [CrossRef]  

14. I. V. Bondarev and Ph. Lambin, “van der Waals coupling in atomically doped carbon nanotubes,” Phys. Rev. B 72, 035451 (2005). [CrossRef]  

15. I. V. Bondarev and Ph. Lambin, “Near-field electrodynamics of atomically doped carbon nanotubes,” in Trends in Nanotubes Research, D. A. Martin, ed. (Nova Science, 2006), Ch. 6, pp. 139–183.

16. I. V. Bondarev, L. M. Woods, and K. Tatur, “Strong exciton-plasmon coupling in semiconducting carbon nanotubes,” Phys. Rev. B 80, 085407 (2009). [CrossRef]  

17. I. V. Bondarev, L. M. Woods, and A. Popescu, “Exciton-plasmon interactions in individual carbon nanotubes,” in Plasmons: Theory and Applications, K. N. Helsey, ed. (Nova Science, 2011), Ch. 16, pp. 381–435.

18. I. V. Bondarev, “Single-wall carbon nanotubes as coherent plasmon generators,” Phys. Rev. B 85, 035448 (2012). [CrossRef]  

19. I. V. Bondarev and T. Antonijevic, “Surface plasmon amplification under controlled exciton-plasmon coupling in individual carbon nanotubes,” Phys. Stat. Sol. C 9, 1259–1264 (2012). [CrossRef]  

20. I. V. Bondarev and A. V. Meliksetyan, “Possibility for exciton Bose-Einstein condensation in carbon nanotubes,” Phys. Rev. B 89, 045414 (2014). [CrossRef]  

21. A. Popescu, L. M. Woods, and I. V. Bondarev, “Chirality dependent carbon nanotube interactions,” Phys. Rev. B 83, 081406 (2011). [CrossRef]  

22. L. M. Woods, A. Popescu, D. Drosdoff, and I. V. Bondarev, “Dispersive interactions in graphitic nanostructures,” Chem. Phys. 413, 116–122 (2013). [CrossRef]  

23. I. V. Bondarev and B. Vlahovic, “Optical absorption by atomically doped carbon nanotubes,” Phys. Rev. B 74, 073401 (2006). [CrossRef]  

24. I. V. Bondarev and B. Vlahovic, “Entanglement of a pair of atomic qubits near a carbon nanotube,” Phys. Rev. B 75, 033402 (2007). [CrossRef]  

25. M. F. Gelin, I. V. Bondarev, and A. V. Meliksetyan, “Monitoring bipartite entanglement in hybrid carbon nanotube systems via optical 2D photon-echo spectroscopy,” Chem. Phys. 413, 123–131 (2013). [CrossRef]  

26. M. F. Gelin, I. V. Bondarev, and A. V. Meliksetyan, “Optically promoted bipartite atomic entanglement in hybrid metallic carbon nanotube systems,” J. Chem. Phys. 140, 064301 (2014). [CrossRef]   [PubMed]  

27. I. V. Bondarev, “Surface electromagnetic phenomena in pristine and atomically doped carbon nanotubes,” J. Comp. Theor. Nanoscience 7, 1673–1687 (2010). [CrossRef]  

28. I. V. Bondarev, M. F. Gelin, and A. V. Meliksetyan, “Tunable plasmon nanooptics with carbon nanotubes,” in Dekker Encyclopedia of Nanoscience and Nanotechnology, S. E. Lyshevski, ed. (3, CRC, 2014), pp. 4989–5001.

29. T. Pichler, M. Knupfer, M. S. Golden, J. Fink, A. Rinzler, and R. E. Smalley, “Localized and delocalized electronic states in single-wall carbon nanotubes,” Phys. Rev. Lett. 80, 4729–4732 (1998). [CrossRef]  

30. S. Tasaki, K. Maekawa, and T. Yamabe, “π-band contribution to the optical properties of carbon nanotubes: Effects of chirality,” Phys. Rev. B 57, 9301–9318 (1998). [CrossRef]  

31. Z. M. Li, Z. K. Tang, H. J. Liu, N. Wang, C. T. Chan, R. Saito, S. Okada, G. D. Li, J. S. Chen, N. Nagasawa, and S. Tsuda, “Polarized absorption spectra of single-walled 4 Å carbon nanotubes aligned in channels of an AlPO4–5 single crystal,” Phys. Rev. Lett. 87, 127401 (2001). [CrossRef]  

32. I. V. Bondarev, G.Ya. Slepyan, and S. A. Maksimenko, “Spontaneous decay of excited atomic states near a carbon nanotube,” Phys. Rev. Lett. 89, 115504 (2002). [CrossRef]   [PubMed]  

33. A. G. Marinopoulos, L. Reining, A. Rubio, and N. Vast, “Optical and loss spectra of carbon nanotubes: Depolarization effects and intertube interactions,” Phys. Rev. Lett. 91, 046402 (2003). [CrossRef]   [PubMed]  

34. I. V. Bondarev and Ph. Lambin, “Spontaneous-decay dynamics in atomically doped carbon nanotubes,” Phys. Rev. B 70, 035407 (2004). [CrossRef]  

35. Actual DOS resonance frequencies are slightly red shifted relative to their respective plasmon resonance frequencies [cf. Figs. 1(a) and 1(b)]. The shifts are within plasmon resonance widths though, and so are neglected, thereby reducing the number of relevant theory parameters here.

36. V. B. Berestetskii, E. M. Lifshitz, and L. P. Pitaevskii, Quantum Electrodynamics (Pergamon, 1982).

37. P. Y. Yu and M. Cardona, Fundamentals of Semiconductors, 4. (Springer-Verlag, 2010). [CrossRef]  

38. S. V. Gaponenko, V. N. Bogomolov, E. P. Petrov, A. M. Kapitonov, D. A. Yarotsky, I. I. Kalosha, A. A. Eychmueller, A. L. Rogach, J. McGilp, U. Woggon, and F. Gindele, “Spontaneous emission of dye molecules, semiconductor nanocrystals, and rare-earth ions in opal-based photonic crystals,” J. Lightwave Technol. 17, 2128–2138 (1999). [CrossRef]  

39. H. Schniepp and V. Sandoghdar, “Spontaneous emission of europium ions embedded in dielectric nanospheres,” Phys. Rev. Lett. 89, 257403 (2002). [CrossRef]   [PubMed]  

40. N. Noginova, G. Zhu, M. Mavy, and M. A. Noginov, “Magnetic dipole based systems for probing optical magnetism,” J. Appl. Phys. 103, 07E901 (2008). [CrossRef]  

41. V. Delgado and J. M. Gomez Llorente, “Weak-coupling-like time evolution of driven four-level systems in the strong-coupling regime,” Phys. Rev. A 68, 022503 (2003). [CrossRef]  

42. J. McKeever, A. Boca, A. D. Boozer, J. R. Buck, and H. J. Kimble, “Experimental realization of a one-atom laser in the regime of strong coupling,” Nature 425, 268–271 (2003). [CrossRef]   [PubMed]  

43. F. Xia, M. Steiner, Y.-M. Lin, and Ph. Avouris, “A microcavity-controlled, current-driven, on-chip nanotube emitter at infrared wavelengths,” Nature Nanotechn. 3, 609–613 (2008). [CrossRef]  

44. E. Gaufrès, N. Izard, X. Le Roux, S. Kazaoui, D. Marris-Morini, E. Cassan, and L. Vivien, “Optical microcavity with semiconducting single-wall carbon nanotubes,” Opt. Express 18, 5740–5745 (2010). [CrossRef]   [PubMed]  

45. R. Miura, S. Imamura, R. Ohta, A. Ishii, X. Liu, T. Shimada, S. Iwamoto, Y. Arakawa, and Y. K. Kato, “Ultralow mode-volume photonic crystal nanobeam cavities for high-efficiency coupling to individual carbon nanotube emitters,” Nature Commun. 5, 5580 (2014). [CrossRef]  

46. S. Münch, S. Reitzenstein, P. Franeck, A. Löffler, T. Heindel, S. Höfling, L. Worschech, and A. Forchel, “The role of optical excitation power on the emission spectra of a strongly coupled quantum dot-micropillar system,” Opt. Express 17, 12821–12828 (2009). [CrossRef]   [PubMed]  

47. M. Nomura, N. Kumagai, S. Iwamoto, Y. Ota, and Y. Arakawa, “Laser oscillation in a strongly coupled single-quantum-dot-nanocavity system,” Nature Phys. 6, 279–283 (2010). [CrossRef]  

48. L. Jensen, L. L. Zhao, J. Autschbach, and G. C. Schatz, “Theory and method for calculating resonance Raman scattering from resonance polarizability derivatives,” J. Chem. Phys. 123, 174110 (2005). [CrossRef]   [PubMed]  

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 (5)

Fig. 1
Fig. 1 (a) Fragment of the energy dependence of the dimensionless (normalized by e 2 / 2 π h ¯ ) axial surface conductivities σzz for the semiconducting (6,4), (10,0) and (11,0) nanotubes of increasing diameter. Peaks of Reσzz represent excitons (E 11, E 22, …); peaks of Re(1zz ) are inter-band plasmons (P 11, …). (b) Photonic DOS functions for the CNs in (a) with the TLS placed at the distance rA =Rcn +2b (see inset). Dimensionless energy is [Energy]/2γ 0. Conductivities are obtained using the (k · p)-scheme developed by Ando [12]. DOS functions are calculated as described by Bondarev and Lambin in [14, 15]. See text for notations.
Fig. 2
Fig. 2 Schematic of the energy level structure as given by Eq. (8) for the coupled four-level CN–TLS system. Thick red lines show the eigen energy levels as functions of xp . Thin red dashed lines indicate their broadening due to finite Δxp . Horizontal dotted lines are to show Rabi-splitting and the in-resonance strong-coupling solutions given by Eq. (8) with δ = 0.
Fig. 3
Fig. 3 Schematic of the Raman scattering process (top) in terms of the inter-level transitions (levels sketched in Fig. 2) of the coupled CN–TLS system given by Eqs. (3), (8) and (9), and the Feynman diagrams (bottom) for the scattering cross-section calculations.
Fig. 4
Fig. 4 Enhancement factor A ( δ , X , Δ x p ) / Δ x p 4 as given by Eq. (17) for Δxp = 0.005 (left) and for X =0.05 (right) to show the influence of the detuning δ ( = x ˜ A x p ) on the maximum intensity of the plasmon enhanced Raman scattering effect.
Fig. 5
Fig. 5 Raman scattering probability as a function of the incident xi and scattered xs photon energies as given by Eq. (16) with δ = X / 3 (maximum of A / Δ x p 4 in Fig. 4, see text) for x p = 0.35 [ P 11 ( 6 , 4 ) plasmon in Fig. 1(a)], with X and Δxp being varied independently [rows (a) and (b)]. TLS–plasmon coupling strength is represented by the ratio X/Δxp being greater or less than unity for strong and weak coupling, respectively. Raman scattering is seen to be manifestly indicative of the strong TLS–plasmon coupling, dramatically increasing as X/Δxp goes much greater than unity and disappearing when it is comparable with unity.

Equations (27)

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

H ^ = H ^ F + H ^ A + H ^ A F = 0 d ω h ¯ ω d R f ^ ( R , ω ) f ^ ( R , ω ) + h ¯ ω ˜ A 2 σ ^ Z + 0 d ω d R [ g ( + ) ( r A , R , ω ) σ ^ g ( ) ( r A , R , ω ) σ ^ ] f ^ ( R , ω ) + h . c . ,
ω ˜ A = ω A [ 1 2 ( h ¯ ω A ) 2 0 d ω d R | g ( r A , R , ω ) | 2 ] .
g ( ± ) ( r A , R , ω ) = g ( r A , R , ω ) ± ω ω A g ( r A , R , ω ) , g ( ) ( r A , R , ω ) = i 4 ω A c 2 d z π h ¯ ω Re σ z z ( ω ) ( ) G z z ( r A , R , ω ) ,
( ) G z z ( r A , R , ω ) = d r δ z z ( ) ( r A r ) G z z ( r , R , ω )
δ α β ( r ) = α β 1 4 π | r | , δ α β ( r ) = δ α β δ ( r ) δ α β ( r )
α = r , φ , z ( × × ω 2 c 2 ) z α G α z ( r , R , ω ) = δ ( r R ) ,
d R | g ( ± ) ( r A , R , ω ) | 2 = h ¯ 2 2 π Γ 0 ( ω ) [ ξ ( r A , ω ) + ω A 2 ω 2 ξ ( r A , ω ) ] ,
ξ ( ) ( r A , ω ) = Im ( ) G z z ( ) ( r A , r A , ω ) Im G z z 0 ( ω )
( ) G z z ( ) ( r A , r A , ω ) = d r d r δ z z ( ) ( r A r ) G z z ( r , r , ω ) δ z z ( ) ( r r A )
| 0 = | l | { 0 } , | 1 , 2 = C u ( 1 , 2 ) | u | { 0 } + 0 d ω d R C l ( 1 , 2 ) ( R , ω ) | l | { 1 ( R , ω ) } , | 3 = | u | { 1 ( R , ω ) } .
E 0 = h ¯ ω ˜ A 2 , E 3 = h ¯ ω ˜ A 2 + h ¯ ω
{ ( h ¯ ω ˜ A 2 E ) C u ( 1 , 2 ) + 0 d ω d R g ( + ) ( r A , R , ω ) C l ( 1 , 2 ) ( R , ω ) = 0 , [ g ( + ) ( r A , R , ω ) ] * C u ( 1 , 2 ) + ( h ¯ ω ˜ A 2 + h ¯ ω E ) C l ( 1 , 2 ) ( R , ω ) = 0.
C l ( 1 , 2 ) ( R , ω ) = [ g ( + ) ( r A , R , ω ) ] * h ¯ ω ˜ A / 2 h ¯ ω + E C u ( 1 , 2 )
E = h ¯ ω ˜ A 2 + 0 d ω d R | g ( + ) ( r A , R , ω ) | 2 h ¯ ω ˜ A / 2 h ¯ ω + E = h ¯ ω ˜ A 2 + h ¯ 2 2 π 0 d ω Γ 0 ( ω ) ( 1 + ω A 2 / ω 2 ) ξ ( r A , ω ) h ¯ ω ˜ A / 2 h ¯ ω + E
0 d ω Γ 0 ( ω ) ( 1 + ω A 2 / ω 2 ) ξ ( r A , ω ) h ¯ ω ˜ A / 2 h ¯ ω + E Γ 0 ( ω p ) ( 1 + ω A 2 / ω p 2 ) ξ ( r A , ω p ) Δ ω 0 2 h ¯ ω ˜ A / 2 h ¯ ω p + E 0 d ω ( ω ω p ) 2 + Δ ω 0 2 ,
ε 0 = x ˜ A 2 , ε 1 , 2 = 1 2 ( x p δ 2 + X 2 i Δ x p ) , ε 3 = x ˜ A 2 + x p i Δ x p .
| C u ( 1 , 2 ) | 2 + 0 d ω d R | C l ( 1 , 2 ) ( R , ω ) | 2 = 1 ,
C u ( 1 , 2 ) = [ 1 2 ( 1 + 1 1 + X 2 / δ 2 1 + X 2 / δ 2 1 + X 2 / δ 2 ) ] 1 / 2 , 0 d ω d R | C l ( 1 , 2 ) ( R , ω ) | 2 = 1 | C u ( 1 , 2 ) | 2 = ( X 2 / 2 δ 2 ) | C u ( 1 , 2 ) | 2 1 + X 2 / 2 δ 2 1 + X 2 / δ 2 .
n | H ^ R ( ω i ) | 0 = i c 2 π h ¯ ω i d z cos ϑ i C u ( n ) * , cos ϑ i = e i · e z , n = 1 , 2
1 | H ^ A F ( e ) | 2 = 0 d ω d R [ C l ( 1 ) ( R , ω ) g ( + ) ( r A , R , ω ) ] * C u ( 2 )
( 2 π h ¯ ) | 0 | H ^ R ( ω s ) | 1 1 | H ^ A F ( e ) | 2 2 | H ^ R ( ω i ) | 0 [ h ¯ ω i h ¯ ω p ( E 1 E 0 ) ] [ h ¯ ω i ( E 2 E 0 ) ] + 0 | H ^ R ( ω i ) | 1 1 | H ^ A F ( e ) | 2 2 | H ^ R ( ω s ) | 0 [ h ¯ ω s h ¯ ω p ( E 1 E 0 ) ] [ h ¯ ω s ( E 2 E 0 ) ] | 2 δ ( h ¯ ω i h ¯ ω p h ¯ ω s ) + | 0 | H ^ R ( ω s ) | 2 2 | H ^ A F ( a ) | 1 1 | H ^ R ( ω i ) | 0 [ h ¯ ω i + h ¯ ω p ( E 2 E 0 ) ] [ h ¯ ω i ( E 1 E 0 ) ] + 0 | H ^ R ( ω i ) | 2 2 | H ^ A F ( a ) | 1 1 | H ^ R ( ω s ) | 0 [ h ¯ ω s + h ¯ ω p ( E 2 E 0 ) ] [ h ¯ ω s ( E 1 E 0 ) ] | 2 δ ( h ¯ ω i + h ¯ ω p h ¯ ω s ) .
1 | H ^ A F ( e ) | 2 = 2 γ 0 C u ( 1 ) * C u ( 2 ) ( X 2 / 4 ) x ˜ A / 2 x p + ε 1 = 2 | H ^ A F ( a ) | 1 ,
0 | H ^ R | 1 1 | H ^ A F ( e ) | 2 2 | H ^ R | 0 = ( 0 | H ^ R | 2 2 | H ^ A F ( a ) | 1 1 | H ^ R | 0 ) = 2 π h ¯ ω i ω s c 2 d z 2 cos ϑ i cos ϑ s 2 γ 0 | C u ( 1 ) C u ( 2 ) | 2 ( X 2 / 4 ) x ˜ A / 2 x p + ε 1
| C u ( 1 ) C u ( 2 ) | 2 = X 2 / 4 δ 2 + X 2 ,
d σ d Ω s = ( 2 γ 0 ) 2 | d z | 4 h ¯ 4 c 4 cos 2 ϑ i cos 2 ϑ s P ( x i , x s ) ,
P ( x i , x s ) = x i x s 3 A ( δ , X , Δ x p ) { 1 [ ( x i x p δ + / 2 ) 2 + Δ x p 2 ] [ ( x s x p δ / 2 ) 2 + Δ x p 2 ] + 1 [ ( x i x p δ / 2 ) 2 + Δ x p 2 ] [ ( x s x p δ + / 2 ) 2 + Δ x p 2 ] } ,
A ( δ , X , Δ x p ) = X 8 2 6 ( δ 2 + X 2 ) 2 ( δ 2 + Δ x p 2 ) ,
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.