Home Selection rule engineering of forbidden transitions of a hydrogen atom near a nanogap
Article Open Access

Selection rule engineering of forbidden transitions of a hydrogen atom near a nanogap

  • Hyunyoung Y. Kim and Daisik S. Kim EMAIL logo
Published/Copyright: July 22, 2017
Become an author with De Gruyter Brill

Abstract

We perform an analytical study on the allowance of forbidden transitions for a hydrogen atom placed near line dipole sources, mimicking light emanating from a one-dimensional metallic nanogap. It is shown that the rapid variation of the electric field vector, inevitable in the near zone, completely breaks the selection rule of Δl=±1. While the forbidden transitions between spherically symmetric S states, such as 2S to 1S or 3S to 1S (Δl=0), are rather robust against selection rule breakage, Δl=±2 transitions such as between 3D and 1S or 3D and 2S states are very vulnerable to the spatial variation of the perturbing electric field. Transitions between 2S and 3D states are enhanced by many orders of magnitude, aided by the quadratic nature of both the perturbing Hamiltonian and D wavefunctions. The forbidden dipole moment, which approaches one Bohr radius times the electric charge in the vicinity of the gap, can be written in a simple closed form owing to the one-dimensional nature of our gap. With large enough effective volume together with the symmetric nature of the excited state wavefunctions, our work paves way towards atomic physics application of infinitely long nanogaps.

1 Introduction

Long wavelength approximation is at the heart of well-known selection rules in atomic spectroscopy. The wavelength of light is much larger than the atom size, so that the light-atom interaction Hamiltonian can safely ignore the spatial variation in the scale of the wavelength, resulting in an effective Hamiltonian in the form of pE=exεE0eiωt where p is the dipole moment operator, E the electric field of light, ε the polarization direction unit vector, E0 the amplitude of the electric field, and ω the angular frequency of light. Thereby, spontaneous emission, stimulated emission, and absorption are all proportional to the matrix element ψf|x|ψi (ψf , ψi =final and initial state wavefunctions, respectively), from which all selection rules follow. The most important selection rule Δl=±1; Δm=±1, 0 originates from the position operator x being represented by the spherical harmonics of order 1. While this selection rule can be broken by magnetic dipole transition, electric quadrupole transition, or by two photon transitions, these forbidden transitions are typically several orders of magnitudes weaker. For instance, the 2P-1S allowed transition lifetime of 2 ns for a hydrogen atom compares favorably with 4.6 days for the forbidden magnetic dipole transition lifetime of 2S to 1S. On the other hand, the electric quadrupole transition between 3D and 2S is somewhat less forbidden, taking 20 ms [1], [2]. Of some practical importance, this one-photon lifetime of the forbidden magnetic dipole transition between 2S and 1S is so impractically long that it is easily superseded by the two-photon lifetime of 0.15 s [3], which played an important role in the measurements of the Lamb shift [4].

While the spatial variation of electromagnetic waves in free space occurs within the wavelength scale, close to the induced sources such as surface current and surface charges which naturally occur in metallic nano objects, electric field vectors can vary in length scale much smaller than their vacuum wavelength, in the length scale of the nano objects themselves or the gap size between the metallic objects [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18], [19]. Of particular interest in the present paper is the one-dimensional metallic nano- and sub-nanogaps whose widths can be in the 1–0.1 nm regime [17], [18], [19], comparable to the spatial extents of hydrogen atom wavefunctions while maintaining a macroscopic length of 1 mm to 1 cm. Electric fields emanating from these gaps possess rapidly varying electric fields, both in magnitude and in direction, in the length scale of the gap itself, creating a potentially very useful field configuration for the purpose of breaking down well-known selection rules, thereby facilitating forbidden transitions in large enough volumes to be experimentally detectable.

2 Materials and methods

To model spatial variation of the electric field emanating from nano- and sub-nanometer gaps, we first consider a line dipole with a line charge density λ and a gap width of w, fed by an alternating current source of angular frequency ω and surface current density Keiωt with the charge conservation relationship K=iωλ. In the extreme subwavelength regime of our interest, we can ignore the retarded time, so that the electric field is approximated by the near-field term only [20], [21], [22], [23], [24], [25], [26]:

(1)E(x,z,t)=λeiωt2πε0(r^1r1r^2r2)=λeiωt2πε0[(xw2,z)(xw2)2+z2(x+w2,z)(x+w2)2+z2],

as plotted in Figure 1A for w=1 nm. By replacing λ with σdz and integrating over a film thickness of h=100 nm, we obtain a realistic field profile of a capacitative nanogap of surface charge density σ, as shown in Figure 1B. For large enough h<<w, this field is well approximated by a simple form:

Figure 1: (A) Electric field lines above a line dipole of gap width 1 nm. Line charges are located at x=±0.5 nm, z=0. (B) Electric field profile obtained by an integration of the line dipoles from z=−100 to 0 nm, well fitted with an analytical form E→(x,z)=σw2πε0(z,−x)x2+z2$\vec E(x,z) = {{\sigma w} \over {2\pi {\varepsilon _0}}}{{(z, - x)} \over {{x^2} + {z^2}}}$ when far away from the gap. (C) Electric field lines E→(x,z)−E→(x0,z0)$\vec E(x,z) - \vec E({x_0},{z_0})$ around a center position of (x0, z0)=(3 nm, 3 nm) together with |ψ320(x, 0, z)|2. The color scale is in arbitrary units and we use the integrated field profile of Eq. (2). (D) Contour plot of an xz cross section of the overlap integrand ψ320Hint(2)ψ200${\psi _{320}}H_{{\mathop{\rm int}} }^{(2)}{\psi _{200}}$ around the center position of (x0, z0) (3 nm, 3 nm). The color scale is in arbitrary units.
Figure 1:

(A) Electric field lines above a line dipole of gap width 1 nm. Line charges are located at x=±0.5 nm, z=0. (B) Electric field profile obtained by an integration of the line dipoles from z=−100 to 0 nm, well fitted with an analytical form E(x,z)=σw2πε0(z,x)x2+z2 when far away from the gap. (C) Electric field lines E(x,z)E(x0,z0) around a center position of (x0, z0)=(3 nm, 3 nm) together with |ψ320(x, 0, z)|2. The color scale is in arbitrary units and we use the integrated field profile of Eq. (2). (D) Contour plot of an xz cross section of the overlap integrand ψ320Hint(2)ψ200 around the center position of (x0, z0) (3 nm, 3 nm). The color scale is in arbitrary units.

(2)E(x,z,t)=σw2πε0(z,x)x2+z2eiωt

for distances larger than w but smaller than h. Unless otherwise indicated, calculations are performed using the integrated field profile of Eq. (2) using Mathematica and Matlab.

3 Results

We now place a hydrogen atom at a position (x0, z0) in a field profile given as Eq. (3), and ask how the transition rates between different states will change relative to the case of a plane wave excitation. Near the source, say, at a position of (x0, z0)=(3 nm, 3 nm), the field lines are curved in the scale of the 3D wavefunctions of a hydrogen atom. It behooves us to examine the behavior of the electric field and the interaction Hamiltonian at the near field. Plotted in Figure 1C at (x0, z0)=(3 nm, 3 nm) are the electric field lines minus their central value, E(x,z)E(x0,z0), with an xz cross section of the 3D wavefunction |ψ320|2 (n=3; l=2; m=0) in an area of 2 nm by 2 nm squared. The field lines approximate those of a vector field (x′, −z′), with (x, z)≡(x0+x′, z0+z′), producing an interaction Hamiltonian with a symmetry of x′2z′2, as can be seen clearly when we expand around (x0, z0):

(3)Hint=eϕ(x,z)eϕ(x0,z0)=en=11n!(xx+zz)nϕ(x+x0,z+z0)|x=z=0Hint(1)+Hint(2)+,

where e is the electron charge.

The first term is the dipole approximation Hamiltonian that gives rise to the usual selection rules, whereas the second term contains all the salient features:

(4)Hint(2)=e12(xx+zz)2ϕ(x+x0,z+z0)|(x=0,z=0)=e(x2z2)2Ezz|(x=x0,z=z0)exzEzx|(x=x0,z=z0)

having taken advantage of the divergence relation Exx+Ezz=0. To see how this Hamiltonian consisting of two-dimensional quadrutic polynomials can be taken advantage of by the D waves, we resume our interest in the 2S to 3D transition. With the photon energy of 13.6 eV(1419)=1.89 eV (656 nm) well within the visible range, we can produce the essentially cylindrical field profile near the nanogap using common transition metals. In Figure 1D, the y cross section of the overlap integrand ψ320Hint(2)ψ200 is shown in a 2 nm by 2 nm area with the hydrogen nucleus at (x0, z0)=(3 nm, 3 nm). The integrand ψ320Hint(2)ψ200 stays mostly positive, because the symmetries of z′2x′2 from the Hamiltonian and 2z′2x′2 from ψ320 are quite similar. This result suggests that there will be a significant transition matrix element between 2S and 3D states, especially at the vicinity of the gap.

Note that ψ322ψ32+2+ψ3222 also couples to ψ200 through z′2x′2, with a matrix element smaller by a factor 3, whereas the exzEzx|(x=x0,z=z0) part of the Hamiltonian does not participate significantly along the x=z line since |Ezx|0 along this line. Clearly, for general directions we also need to consider excitations into ψ32+1 and ψ32−1 through xz′.

To quantify how strong the forbidden transition matrix elements are between 2S and 3D states, we recall transition dipole moments of allowed excitations. Choosing a local electric field orientation as the z direction, a relevant dipole moment is defined as

dallowed=|ψf|Hint|ψi|E0=|ψf|eE0z|ψi|E0=|ψf|ez|ψi|~eaB,

where E0 is the electric field strength at the hydrogen nucleus and aB is the Bohr radius. Analogously, we define the transition dipole moment of a forbidden 2S to 3D excitation such that

d200320=|ψ320|Hint|ψ200|E0(x0,z0).

In Figure 2, we quantify forbidden dipole moments of the 2S to 3D transitions. We calculate the total forbidden dipole/transition moment d2S3D=md20032m2 along the x=z line, as shown in Figure 2A. Calculations using the full Hint are represented by blue squares, while those using only Hint(2) are represented by a blue line, displaying near perfect agreement. Finally, taking advantage of |Ezz||E(x,z)|1|x| and |Ezx|0 along this line, we reach the simple closed form approximation for the total forbidden dipole moment:

Figure 2: (A) Total forbidden dipole moment along the x=z line, calculated from the second-order Hamiltonian (blue line), the full interaction Hamiltonian (blue squares), and a fit to the analytical expression 0.719/2z$0.719/\sqrt 2 z$ (red line). (B) Contour plot of the forbidden transition moment d200→320=|〈ψ320|Hint|ψ200〉|E0(x0,z0)${d_{200 \to 320}} = {{\left| {\left\langle {{\psi _{320}}} \right|{H_{{\mathop{\rm int}} }}\left| {{\psi _{200}}} \right\rangle } \right|} \over {{E_0}({x_0},{z_0})}}$ in units of eaB where aB is the Bohr radius, in a 20 nm by 20 nm area starting from z=1 nm. (C) Contour plot of the forbidden transition moment d200→321=|〈ψ321|Hint|ψ200〉|E0(x0, z0).${d_{200 \to 321}} = {{\left| {\left\langle {{\psi _{321}}} \right|{H_{{\mathop{\rm int}} }}\left| {{\psi _{200}}} \right\rangle } \right|} \over {{E_0}({x_0},{\rm{ }}{z_0})}}.$ (D) Contour plot of the total forbidden transition moment d2S→3D=d200→3202+d200→3222+d200→3212.${d_{{\rm{2S}} \to {\rm{3D}}}} = \sqrt {d_{200 \to 320}^2 + d_{200 \to 322}^2 + d_{200 \to 321}^2} .$ The cylindrical symmetry is largely restored.
Figure 2:

(A) Total forbidden dipole moment along the x=z line, calculated from the second-order Hamiltonian (blue line), the full interaction Hamiltonian (blue squares), and a fit to the analytical expression 0.719/2z (red line). (B) Contour plot of the forbidden transition moment d200320=|ψ320|Hint|ψ200|E0(x0,z0) in units of eaB where aB is the Bohr radius, in a 20 nm by 20 nm area starting from z=1 nm. (C) Contour plot of the forbidden transition moment d200321=|ψ321|Hint|ψ200|E0(x0,z0). (D) Contour plot of the total forbidden transition moment d2S3D=d2003202+d2003222+d2003212. The cylindrical symmetry is largely restored.

(5)d2S3D|ψ320|x2z22|ψ200|2+|ψ322|x2z22|ψ200|2eE0x0E0=2163458eaB22x013.6eaBaBρ0,

represented by a red line. In unit of eaB the red line corresponds to 0.7192x0=0.719ρ0, where x0 and the distance from the origin ρ0 are in nanometers. All three results agree rather well.

To see the angular dependences of various forbidden excitations, we plot d2003202+d2003222 in Figure 2B, demonstrating that indeed excitations into ψ322 and ψ320 are maximum along the x=z line. Transition dipole moments into ψ321=ψ32+1+ψ3212,d200→321, are plotted in Figure 2C, showing an almost orthogonal angle dependence from that of d2003202+d2003222. Adding all the forbidden dipoles, an almost isotropic d2S3D=d2003202+d2003212+d2003222 is obtained (Figure 2D), fitted rather well with the simple analytical expression of 0.719ρ0 now applicable to all directions within an error of 0.2–2%, as we move away from the z-axis towards the x-axis.

We now study the width dependence of the forbidden transitions. Since the cylindrical symmetry of the total forbidden dipole moment is only approximate in our geometry, we expect to find deviations as we increase the gap width. Figure 3A depicts the case of d2S→3D for w=3 nm. Away from the gap, the cylindrical symmetry is recovered, whereas for distances less than 5 nm, angular deviations and weaker moments are evident. For w=10 nm in Figure 3B, the deviations are more pronounced, but again, at distances larger than 10 nm, the forbidden dipole moment converges to those of narrower gaps. Figure 3C plots the forbidden dipole moment along the z-axis for several gap widths. For gap widths of 3, 5, and 10 nm, the forbidden dipole moments eventually converge to the 1/z line at z~w. Scanning along the x-direction for a fixed z=1 nm, the behavior is very different. At x=0, forbidden dipole moments are smaller mainly because along the middle of the gap, field curvatures are less. For w=3, 5, and 10 nm, d2S→3D peaks at xw2 because sharp edges of the charge distributions are located at (x=±w2,z=0). Again, at all instances d2S→3D recovers the 0.719ρ0 dependence for x>10 nm.

Figure 3: Total forbidden dipole moment d2S→3D plotted (A) when w=3 nm and (B) when w=10 nm. (C) Total forbidden dipole moment along the z-axis when w=1, 3, 5, and 10 nm. (Inset: a log-log plot). (D) Total forbidden dipole moment when the observation is along the x-axis keeping z=1 nm for w=1, 3, 5, and 10 nm.
Figure 3:

Total forbidden dipole moment d2S→3D plotted (A) when w=3 nm and (B) when w=10 nm. (C) Total forbidden dipole moment along the z-axis when w=1, 3, 5, and 10 nm. (Inset: a log-log plot). (D) Total forbidden dipole moment when the observation is along the x-axis keeping z=1 nm for w=1, 3, 5, and 10 nm.

We now consider the excited state wavefunction

(6)|ψex|ψ321ψ321|Hint|ψ200+|ψ320ψ320|Hint|ψ200+|ψ322ψ322|Hint|ψ200

at various locations and conditions. The near-perfect cylindrical symmetry of d2S→3D for the 1 nm gap case suggests a strong symmetry for |ψex as well. Figure 4A displays the excited state wavefunction squared using the full interaction Hamiltonian at three different locations. The excited state wavefunctions faithfully reproduce a pure |ψ321 state within a coordinate system defined by the local field orientation. While physically intuitive, mathematically it is because (x2z2)2Ezz=xzEzx when using the field profile of Eq. (2). We found that even with larger gap widths, excited wavefunctions’ orientations following the local field orientation remain largely unaffected except for right above the sharp edge (Figure 4B). In stark contrast, as shown in Figure 4C, the single wire case described by Eq. (1) displays wavefunctions not rotating with the local field orientation in a simplistic way. In spite of this complication, the excited wavefunctions remain a pure |ψ321 state at a properly rotated coordinate system, which directly follows from the Hamiltonian of Eq. (4) containing only two quadratic terms: (x2z2); 2xz. Replacing our source by a simple point dipole and using Eq. (6) with three-dimensional Hamiltonian produce a whole combination of D wavefunctions evident in Figure 4D. On the equator relative to the dipole orientation, pure |ψ321 still get excited, whereas at most of other directions, all three states contribute.

Figure 4: Plot of the xz cross section of the excited state wavefunction as defined in Eq. (6), at strategic locations under various conditions.(A) w=1 nm with a cylindrical field profile described in Figure 1B, generated by integrating line dipoles from z=−100 to 0 nm, approximated by Eq. (2). (B) Same as (A) except w=4 nm. (C) w=1 nm using the line dipole field of Eq. (1), depicted in Figure 1A. (D) Excited state wavefunctions with a point dipole source. Unlike those excited by the infinite line sources, the wavefunctions are not of a uniform shape.
Figure 4:

Plot of the xz cross section of the excited state wavefunction as defined in Eq. (6), at strategic locations under various conditions.

(A) w=1 nm with a cylindrical field profile described in Figure 1B, generated by integrating line dipoles from z=−100 to 0 nm, approximated by Eq. (2). (B) Same as (A) except w=4 nm. (C) w=1 nm using the line dipole field of Eq. (1), depicted in Figure 1A. (D) Excited state wavefunctions with a point dipole source. Unlike those excited by the infinite line sources, the wavefunctions are not of a uniform shape.

4 Discussion and conclusion

With the 2S to 3D transition being essentially allowed near the gap, we estimate the spontaneous decay lifetime from 3D to 2S states. An effective dipole moment of one Bohr radius gives rise to a lifetime of 44 ns, six orders of magnitudes faster than the quadrupole transition in the vacuum. The physics of this spontaneous emission modification by nanostructures [27], [28] is clear in our case: in vacuum, the quadrupole transition is weaker than the dipole transition by (2πaBλ0)2, where λ0 is the wavelength of light; on the other hand, near the nanogap with a distance of ρ, we replace this factor by (2πaBρ)2, resulting in fast lifetimes comparable to those of the allowed transitions. Our infinite nanogaps have the advantage over point gaps in that it stretches to millimeter to centimeter length scale in the y-axis [11], [12], offering more robustness and million times larger effective areas than point gaps [29]. In attempts to break selection rules by going to shorter wavelength light, for example, X-rays [30], the transitions necessarily involve core orbitals of comparable short length scale, so that the effect is less dramatic than that presented here. Finally, while our technique applies to any S to D transitions, it may not apply to 1S to 3S transitions such as described in [31]. This is because the quadratic potential still gives rise to zero matrix element between two S states because of symmetry.

Our two-dimensional quadratic potentials have, in addition to the obviously larger volume, another advantage over point source dipoles that also give rise to forbidden transitions in surface-enhanced Raman scattering and infrared absorption [23], [24], [25], [26] in molecules. The excited wavefunctions are all of one nature, as shown in Figure 4A–C, which can give rise to constructive interference of quadrupole radiations. Finally, while an analytical field profile has been used throughout our paper, a COMSOL calculation assuming a 1 nm gap sandwiched by aluminum layers of 100 nm thickness at 656 nm produces a field profile of a cylindrical symmetry well described by Eq. (2). Finite-difference-time-domain calculations as well as vector field mapping experiments also support this picture [5], [32], [33], [34]. We therefore expect similar quantum mechanical results under finite elements electromagnetic simulations.

In conclusion, we have shown that the 2S-3D forbidden transition is allowed for all practical purposes, near the vicinity of a metallic nanogap. The relevant scale of this quadrupole transition becomes not the wavelength of light but the gap width and the distance of the atom from the gap. With million times larger effective volume than point gaps, together with the highly symmetric excited state wavefunctions, we foresee an intimate interaction between atomic spectroscopy and now mature nanogap technology in the near future, especially with free standing gaps. With the advantage of metallic nanogaps of infinite length with an ultimate field enhancement [32] whereby electromagnetic waves from microwaves to ultraviolet have all the same near-field profile [33], [34], up to the plasma frequency of metal, selection rule-free spectroscopy of atoms, molecules, and quantum dots will become of wide use.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP: NRF-2015R1A3A2031768) (MOE: BK21 Plus Program-21A20131111123). DSK acknowledges COMSOL simulations of Taehee Kang and helpful discussions with professors Do-Young Noh, Yongil Shin, Changyoung Kim, and Jae-Hoon Park.

References

[1] Jitrik O, Bunge CF, Carlos F. Transition probabilities for hydrogen-like atoms. J Phys Chem Ref Data 2004;33:1059–70.10.1063/1.1796671Search in Google Scholar

[2] Wiese WL, Fuhr JR. Accurate atomic transition probabilities for hydrogen, helium, and lithium. J Phys Chem Ref Data 2009;38:565–719.10.1063/1.3077727Search in Google Scholar

[3] Hansch TW, Lee SA, Wallenstein R, Wieman C. Doppler-free two-photon spectroscopy of hydrogen 1S-2S*. Phys Rev Lett 1974;34:307–9.10.1142/9789812813787_0002Search in Google Scholar

[4] Lamb WE Jr, Retherford RC. Fine structure of the hydrogen atom by a microwave method. Phys Rev 1947;72:241–3.10.1103/PhysRev.72.241Search in Google Scholar

[5] Lee KG, Kihm HW, Kihm JE, et al. Vector field microscopic imaging of light. Nat Photon 2007;1:53–6.10.1038/nphoton.2006.37Search in Google Scholar

[6] Choi SB, Park DJ, Jeong YK, et al. Directional control of surface plasmon polariton waves propagating through an asymmetric Bragg resonator. Appl Phys Lett 2009;94: 063115–7.10.1063/1.3068502Search in Google Scholar

[7] Park HR, Park YM, Kim HS, et al. Terahertz nano-resonators: giant field enhancement and ultrabroadband performance. Appl Phys Lett 2010;96:121106–8.10.1063/1.3368690Search in Google Scholar

[8] Seo MA, Park HR, Koo SM, et al. Terahertz field enhancement by a metallic nano slit operating beyond the skin-depth limit. Nat Photon 2009;3:152–6.10.1038/nphoton.2009.22Search in Google Scholar

[9] Koo SM, Kumar MS, Shin JH, Kim DS, Park NK. Extraordinary magnetic field enhancement with metallic nanowire: role of surface impedance in Babinet’s principle for ub-skin-depth regime. Phys Rev Lett 2009;103:263901–4.10.1103/PhysRevLett.103.263901Search in Google Scholar PubMed

[10] Jeong JY, Rhie JY, Jeon WJ, Hwang CS, Kim DS. High-throughput fabrication of infinitely long 10 nm slit arrays for terahertz applications. J Infrared Millim Te 2015;36:262–8.10.1007/s10762-014-0135-3Search in Google Scholar

[11] Lee KH, Jeong JY, Bahk YM, et al. Microwave funneling through sub-10 nm nanogaps. ACS Photon 2016;3:537–42.10.1021/acsphotonics.6b00047Search in Google Scholar

[12] Olislager L, Kubo W, Tanaka T, et al. Propagation and survival of frequency-binentangled photons in metallic nanostructures. Nanophotonics 2015;4:324–31.10.1515/nanoph-2015-0011Search in Google Scholar

[13] Rui G, Zhan, Q. Tailoring optical complex fields with nano-metallic structures. Nanophotonics 2015;4:2–25.10.1515/nanoph-2014-0018Search in Google Scholar

[14] Khurgin JB, Sun G. How small can “nano” be in a “nanolaser”? Nanophotonics 2012;1:3–8.10.1515/nanoph-2012-0017Search in Google Scholar

[15] Shalaby M, Merbold H, Peccianti M, et al. Concurrent field enhancement and high transmission of THz radiation in nanoslit arrays. Appl Phys Lett 2011;99:041110.10.1063/1.3617476Search in Google Scholar

[16] Novitsky A, Zalkovskij M, Malureanu R, Lavrinenko A. Microscopic model of the THz field enhancement in a metal nanoslit. Opt Commun 2011;23:5495–500.10.1016/j.optcom.2011.08.019Search in Google Scholar

[17] Chen X, Park HR, Pelton M, et al. Atomic layer lithography of wafer-scale nanogap arrays for extreme confinement of electromagnetic waves. Nat Commun 2013;4:2361–6.10.1038/ncomms3361Search in Google Scholar PubMed

[18] Kim JY, Kang BJ, Park JH, et al. Terahertz quantum plasmonics of nanoslot antennas in nonlinear regime. Nano Lett 2015;15:6683–8.10.1021/acs.nanolett.5b02505Search in Google Scholar PubMed

[19] Bahk YM, Kang BJ, Kim YS, et al. Electromagnetic saturation of angstrom-sized quantum barriers at terahertz frequencies. Phys Rev Lett 2015;115:125501.10.1103/PhysRevLett.115.125501Search in Google Scholar PubMed

[20] Jackson JD. Classical electrodynamics. New York, Wiley, 1999.Search in Google Scholar

[21] Bao W, Staffaroni M, Boker J, et al. Plasmonic near-field probes: a comparison of the campanile geometry with other sharp tips. Opt Exp 2013;21:8166–76.10.1364/OE.21.008166Search in Google Scholar PubMed

[22] Cuche A, Drezet A, Sonnefraud Y, et al. Near-field optical microscopy with a nanodiamond-based single-photon tip. Opt Exp 2009;17:19969–80.10.1364/OE.17.019969Search in Google Scholar PubMed

[23] Takase M, Ajiki H, Mizumoti Y, et al. Selection-fuel breakdown in plasmon-induced electronic excitation of an isolated single-walled carbon nanotube. Nat Photon 2013;7:550–4.10.1038/nphoton.2013.129Search in Google Scholar

[24] Iida T, Aiba Y, Ishihara H. Anomalous optical selection rule of an organic molecule controlled by extremely localized light field. Appl Phys Lett 2011;98:053108.10.1063/1.3551710Search in Google Scholar

[25] Moskovits M, DiLella, DP. Surface-enhanced Raman spectroscopy of benzene and benzene-d6 adsorbed on silver. J Chem Phys 1980;73:6068–75.10.1063/1.440142Search in Google Scholar

[26] Isawa T, Takenaka M, Taketsugu T. Generalized theoretical method for the interaction between arbitrary nonuniform electric field and molecular vibrations: toward near-field infrared spectroscopy and microscopy. J Chem Phys 2016;144:124116.10.1063/1.4944937Search in Google Scholar PubMed

[27] Pelton M. Modified spontaneous emission in nanophotonic structures. Nat Photon 2015;9:427–35.10.1038/nphoton.2015.103Search in Google Scholar

[28] Hakhumyan G, Leroy C, Mirzoyan R, Pashayan-Leroy Y, Sarkisyan D. Study of “forbidden” atomic transitions on D2 line using Rb nano-cell placed in external magnetic field. Eur Phys J D 2012;66:119.10.1140/epjd/e2012-20687-2Search in Google Scholar

[29] Jain PK, Ghosh D, Baer R, Rabani E, Alivisatos AP. Near-field manipulation of spectroscopic selection rules on the nanoscale. Proc Natl Acad Sci USA 2012;109:8016–9.10.1073/pnas.1121319109Search in Google Scholar PubMed PubMed Central

[30] Demekhin Ph V. On the breakdown of the electric dipole approximation for hard x-ray photoionization cross sections. J Phys B At Mol Opt Phys 2014;47:025602.10.1088/0953-4075/47/2/025602Search in Google Scholar

[31] Yost DC, Matveev A, Grinin A, et al. Spectroscopy of the hydrogen 1S-3S transition with chirped laser pulses. Phys Rev A 2016;93:042509.10.1103/PhysRevA.93.042509Search in Google Scholar

[32] Bahk YM, Han S, Rhie J, et al. Ultimate terahertz field enhancement of single nano slits. Phys Rev B 2017;95: 075424.10.1103/PhysRevB.95.075424Search in Google Scholar

[33] Ahn JS, Kang T, Singh DK, et al. Optical field enhancement of nanometer-sized gaps at near-infrared frequencies. Opt Exp 2015;23:4897–907.10.1364/OE.23.004897Search in Google Scholar PubMed

[34] Kang T, Rhie J, Park J, et al. Resonance tuning of electric field enhancement of nanogaps. Appl Phys Exp 2015;8: 092003.10.7567/APEX.8.092003Search in Google Scholar

Received: 2017-3-29
Revised: 2017-4-25
Accepted: 2017-5-2
Published Online: 2017-7-22
Published in Print: 2018-1-1

©2017, Daisik S. Kim et al., published by De Gruyter.

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Articles in the same Issue

  1. Review articles
  2. Advances in optoplasmonic sensors – combining optical nano/microcavities and photonic crystals with plasmonic nanostructures and nanoparticles
  3. Periodic array-based substrates for surface-enhanced infrared spectroscopy
  4. From isolated light-harvesting complexes to the thylakoid membrane: a single-molecule perspective
  5. Theory and applications of toroidal moments in electrodynamics: their emergence, characteristics, and technological relevance
  6. Multiple exciton generation in quantum dot-based solar cells
  7. All-integrated terahertz modulators
  8. Exciton-plasmon coupling interactions: from principle to applications
  9. Emerging technologies for high performance infrared detectors
  10. Research articles
  11. Magneto-optical response in bimetallic metamaterials
  12. Saturated evanescent-wave absorption of few-layer graphene-covered side-polished single-mode fiber for all-optical switching
  13. Reversible thermochromic response based on photonic crystal structure in butterfly wing
  14. Selection rule engineering of forbidden transitions of a hydrogen atom near a nanogap
  15. Heterogeneous terahertz quantum cascade lasers exceeding 1.9 THz spectral bandwidth and featuring dual comb operation
  16. Effect of temperature on the structural, linear, and nonlinear optical properties of MgO-doped graphene oxide nanocomposites
  17. Engineering light emission of two-dimensional materials in both the weak and strong coupling regimes
  18. Anisotropic excitation of surface plasmon polaritons on a metal film by a scattering-type scanning near-field microscope with a non-rotationally-symmetric probe tip
  19. Achieving pattern uniformity in plasmonic lithography by spatial frequency selection
  20. Controllable all-fiber generation/conversion of circularly polarized orbital angular momentum beams using long period fiber gratings
  21. High-efficiency/CRI/color stability warm white organic light-emitting diodes by incorporating ultrathin phosphorescence layers in a blue fluorescence layer
  22. Relative merits of phononics vs. plasmonics: the energy balance approach
  23. Efficiency enhancement of InGaN amber MQWs using nanopillar structures
  24. Color display and encryption with a plasmonic polarizing metamirror
  25. Experimental demonstration of an optical Feynman gate for reversible logic operation using silicon micro-ring resonators
  26. Specialized directional beaming through a metalens and a typical application
  27. Anomalous extinction in index-matched terahertz nanogaps
Downloaded on 22.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2017-0037/html?lang=en
Scroll to top button