Home Strongly coupled Raman scattering enhancement revealed by scattering-type scanning near-field optical microscopy
Article Open Access

Strongly coupled Raman scattering enhancement revealed by scattering-type scanning near-field optical microscopy

  • Kang Qin ORCID logo , Kai Liu , Sheng Peng , Zongyan Zuo , Xiao He , Jianping Ding , Yanqing Lu , Yongyuan Zhu and Xuejin Zhang ORCID logo EMAIL logo
Published/Copyright: March 30, 2023
Become an author with De Gruyter Brill

Abstract

Recent advances in near-field technology with an ultrahigh spatial resolution breaking optical diffraction limit, make it possible to further identify surface-enhanced Raman scattering (SERS) enhancement theories, and to monitor the SERS substrates. Here we verify the electromagnetic enhancement mechanism for SERS with a close-up view, using scattering-type scanning near-field optical microscopy. The array of metal-insulator-metal (MIM) subwavelength structures is studied, in which the field enhancement comes from the strong coupling between gap plasmon polariton and surface plasmon polariton modes. The near-field optical measurements reveal that SERS enhancement factor (EF) varies from one MIM subwavelength unit to another in a finite array. Besides the enhancement of isolated unit, the loss exchange phenomenon in strong coupling with a large Rabi splitting can give rise to an additional enhancement of more than 2 orders of magnitude in periodic arrays and close to 3 orders of magnitude in finite arrays. The SERS EF of the array composed of only 5 units is demonstrated to yield the best SERS performance. Our near-field optical measurements show evidence that finite-size structures embodied with strong coupling effect are a key way to develop practical high-performance SERS substrates.

1 Introduction

Surface-enhanced Raman scattering (SERS) spectroscopy is a surface-sensitive technique to track the spectroscopic fingerprint information of analytes [1]. Traditionally, the physical mechanism of the enhancement effect of SERS is based on the amplification of local electromagnetic (EM) field with the localized surface plasmon resonances [2]. SERS systems have been predominantly based on dimer structures, i.e. nanoparticle pairs of noble metals or nanoparticle-on-mirror configurations [35]. The amplified signals which are usually called ‘hot spots’ occur on the surface of the nanoparticles or in the gaps of the dimer structures [68]. Single-molecule SERS has been detected within extremely narrow gaps [9, 10].

The SERS enhancement factor (EF) is very sensitive to the gap configurations and geometric parameters while regular arrangement of nanogaps will promote uniformity of SERS signals. At the same time, such SERS substrates support some kinds of EM modes, such as the band-edge of surface plasmon polariton (SPP) mode [11], gap plasmon polariton (GPP) mode [12], and their coupling modes. Within the coupling regime, the linewidth of resonances can be narrower, indicating lower dissipation. The coupling between dissipative-dissipative modes takes place in most practical optical systems. The Hamiltonian of the two dissipative levels (damping rates are γ1 and γ2, respectively) in a reciprocal system can be expressed as [13]

(1) H = H L + H exchange = i γ + / 2 0 0 i γ + / 2 + β 1 + i γ / 2 g * g β 2 i γ / 2 ,

where β1,2 are the energies of two levels, g and g* are the reciprocal coupling coefficients, γ+ = γ1 + γ2, γ = γ1γ2, H L  is the background damping, Hexchange is the loss exchange term. When close to the coupling region (β1β2 = β0), the eigenvalues can then be written as

(2) β ± = β 0 + i γ + 2 ± g 2 + i γ 2 2 .

In addition to the inevitable background damping (γ+), coupling strength also determines the system’s loss. Reducing the resonant loss can improve the EM enhancement effect of the structures [14], so the loss changing caused by mode coupling is of great significance for modulating the field enhancement effect. The higher the quality factor (Q-factor) and the smaller the mode volume (V), then the stronger the coupling. The strong coupling takes place when Q/√V overcomes a threshold [15, 16]. In the strong coupling regime, energy exchanges between the original modes, accompanying with the generation of hybrid polariton bands and Rabi splitting. In general, V will govern the coupling strength for plasmonic structures with obvious dissipation [17]. Compared to the SPP mode, the GPP mode has a higher spatial confinement to the EM field and thus a smaller V, but its Q-factor is rather lower [18].

In this letter, we investigate strong coupling effect between SPP and GPP modes, which combines the characteristics of high Q-factor and small V, rendering a larger SERS EF than that of single mode resonance. The comprehensive information about highly confined EM field and SERS performance are explored by adjusting geometric parameters and directly visualizing ‘hot spots’ with high spatial resolution technique [19]. The localized EM field and the SERS signal are simultaneously mapped with sub-10 nm resolution using scattering-type scanning near-field optical microscopy (s-SNOM) [20], in which a sharp Si tip is used as both the near-field probe and the Raman analyte. Our results demonstrate that the structures with strong coupling effect manifest superior SERS performance. Furthermore, the favorable structures for SERS applications are finite arrays with several units, rather than periodic ones. The former could provide extremely small V, although the latter has largest Q-factor.

2 Methods

2.1 Near-field optical setup

The s-SNOM system with capability of collecting both near-field electric field and near-field Raman scattering signals is shown in Figure 1(a). Standard Si tip (Nanoworld, Arrow NCR) with apex diameter of less than 20 nm was used to scatter the near-field signals. Dielectric tip usually scatters purer structure signals than metal tip [21, 22]. The detected light can describe the electric field distributions, which are associated with SERS ‘hot spots’ [23]. We used a green laser (532 nm wavelength) as the source. The parabolic mirror focuses the incident laser onto the scanning Si-tip and collects the backscattered light. The tip oscillates vertically at the frequency Ω of cantilever’s mechanical resonance which is 285 kHz typically at our experiments. The background noises can be suppressed by demodulating the detected signals at a higher harmonic order of the tapping frequency, nΩ [24]. We collected the magnitude s4 and phase φ4 of fourth-order signals through a pseudoheterodyne Michelson interferometric detection scheme. For the inelastic Raman scattering detection, we implemented an electron-multiplying CCD (EMCCD, Princeton Instruments) and a PI (Princeton Instruments) spectrometer. A flip mirror reflects the signals into the spectrometer. And a long-pass filter blocks the elastic tip-scattered light. In order to collect different components of the electric field, a polarizer and a half-wave plate were introduced to the light path [2527]. We study one-dimensional arrayed metal-insulator-metal (MIM) subwavelength structures, composed of grooves engraved on Ag surface, which support GPP, SPP, and their coupling modes. The narrow grooves, i.e. MIM subwavelength structures, were fabricated by focused ion beam (FIB) etching on the solution-synthesized single crystalline Ag plates [28]. The arrayed MIM subwavelength structures are illustrated in Figure 1(b). In general, the narrower the groove, the stronger the EM field for the GPP mode. Here we consider grooves with the width of 20 nm, which is accessible by common FIB etching technology, and other geometric parameters can be optimized both theoretically and experimentally (Figure S1).

Figure 1: 
Schematics of near-field optical setup and MIM subwavelength array. (a) Scheme of the s-SNOM setup. The laser firstly goes through an isolator and a beam expander (BE). Then, a half-wave plate (HWP) is used to choose s- or p-polarized incident light. Parabolic mirror focuses the light onto the Si tip and collects elastic and inelastic scattered signals from the tip. Near-field signals are transmitted into the detector after a polarizer (P) or into the spectrometer and the EMCCD after a filter (F) through a flip mirror (FM). Pseudoheterodyne interferometric detection unit composed of a beam-splitter (BS) and a piezo actuator is used to record phase signals. (b) Geometry of the MIM subwavelength structure. A unit cell and geometric parameters are shown in the inset with period (P), width (W), and depth (D). GPP modes can be supported with appropriate W and D. The backscattered light is actually detected in the experiment.
Figure 1:

Schematics of near-field optical setup and MIM subwavelength array. (a) Scheme of the s-SNOM setup. The laser firstly goes through an isolator and a beam expander (BE). Then, a half-wave plate (HWP) is used to choose s- or p-polarized incident light. Parabolic mirror focuses the light onto the Si tip and collects elastic and inelastic scattered signals from the tip. Near-field signals are transmitted into the detector after a polarizer (P) or into the spectrometer and the EMCCD after a filter (F) through a flip mirror (FM). Pseudoheterodyne interferometric detection unit composed of a beam-splitter (BS) and a piezo actuator is used to record phase signals. (b) Geometry of the MIM subwavelength structure. A unit cell and geometric parameters are shown in the inset with period (P), width (W), and depth (D). GPP modes can be supported with appropriate W and D. The backscattered light is actually detected in the experiment.

2.2 Numerical simulations and calculations

In order to explain and verify our experimental results, we obtain the reflection spectrum from the far-field and the near-field electric field distribution results by numerical calculations. The numerical simulation software packages used were Lumerical FDTD Solutions (Lumerical, Inc.). The material properties of Ag used therein are experimental data from Johnson and Christy [29]. The p-polarized light impinges on the sample surface. For finite-size structures, the boundary conditions used are perfectly matched layers (PMLs), and for periodic structures, the boundary conditions are periodic boundary conditions.

3 Results and discussion

3.1 Elastic near-field optical measurements

The acquisitions of near-field optical measurements include both intensity and phase distributions of different electric field components on the surface of structures [2527]. The electric field within the MIM subwavelength structure is dominated by the component pointing from one metal to another metal. During near-field optical measurements, s- and p-polarized components were detected with p-polarized 60° incidence through adding a polarizer in front of the detector, which can be referred to Figure 1(b), but collecting the backscattered field. Under this circumstance, s-polarized s4 signal reflects well the in-plane component E x 2 while p-polarized s4 signal the out-of-plane component E z 2. Figure 2(a) is the surface morphology measured by AFM. Figure 2(b) shows s4 signal as a function of the tip-Ag surface distance, in which the 1/e decay behavior conforms the pure near-field feature of the acquired s4 signal. When the number of units N is large enough, the property of finite MIM subwavelength array is close to that of infinite periodic one. To investigate the periodic array, we kept N > 100 in the experiments, took the isolated MIM subwavelength structure with the width of 20 nm and depth of 35 nm, and changed the array period. For the MIM subwavelength structure with the period of 200 nm, the measured near-field distributions of s-polarized and p-polarized s4 components are shown in Figure 2(c) and (d) respectively. It can be seen that the strongest s-polarized s4 component is concentrated on the inner dielectric domain of the MIM subwavelength structure, while the strongest p-polarized s4 component is distributed on both sides of the MIM subwavelength structure, which correspond to the behavior of E x 2 and E z 2 components. To match those in experiments, the calculated results shown in Figure 2(c) and (d), as well as in the rest of the text, are represented as the near-field signals extracted 20 nm above the sample surface unless other-wise specialized. The asymmetry in measured p-polarized s4 distribution comes from the oblique incidence of laser (Figure S2). Above results show that the MIM subwavelength structure can generate remarkable localized field enhancement.

Figure 2: 
Elastic near-field optical measurements with polarization resolution. (a) Morphology of MIM subwavelength structures by AFM. (b) s4-approach curve measured on the Ag surface. (c) s-polarized near-field intensity mapping and line profile of E
x
2 component. (d) p-polarized near-field intensity mapping and line profile of E
z
2 component. The upper gives the near-field intensity distribution. The lower gives experimental and theoretical line profiles along the white line. The scale bar is 500 nm.
Figure 2:

Elastic near-field optical measurements with polarization resolution. (a) Morphology of MIM subwavelength structures by AFM. (b) s4-approach curve measured on the Ag surface. (c) s-polarized near-field intensity mapping and line profile of E x 2 component. (d) p-polarized near-field intensity mapping and line profile of E z 2 component. The upper gives the near-field intensity distribution. The lower gives experimental and theoretical line profiles along the white line. The scale bar is 500 nm.

3.2 Near-field Raman scattering measurements

The SERS EF can be defined as EF = | E e|4/| E 0|4, where E 0 is the incident electric field and E e is the enhanced electric field. The total EF can be scaled as F2 = F1 × F2, where F1 is the EF of isolated MIM subwavelength structure and F2 is the additional EF that related to coupling interaction. The Stokes shifted Raman scattering spectra can be detected when the scattered near-field signals are transmitted from the tip to the EMCCD via a long pass filter. Figure 3(a) shows a typical near-field Raman scattering spectrum measured from Si tip. We scanned the Si tip along the direction perpendicular to the grooves (x-axis) and collected the spectra at each position with a 10 nm interval. By this way, we obtained the near-field intensity distribution of Raman scattering signals of the samples. Figure 3(b) shows the near-field intensity distribution of Si Raman scattering signal within one period, in which the curve of measured SERS EF is closer to the curve of s42 rather than s4.

Figure 3: 
Near-field Raman scattering measurements and strong coupling mechanism. (a) Detected near-field Raman scattering signal of Si tip. (b) Normalized line profile of near-field Raman scattering intensity, elastically scattered s4, and s42 along x-axis. (c) Period-dependent near-field Raman scattering intensity and elastically scattered s42 in the center of the MIM subwavelength structure. (d) The strong coupling between GPP and SPP modes. The color represents the maximum SERS EF of MIM subwavelength unit.
Figure 3:

Near-field Raman scattering measurements and strong coupling mechanism. (a) Detected near-field Raman scattering signal of Si tip. (b) Normalized line profile of near-field Raman scattering intensity, elastically scattered s4, and s42 along x-axis. (c) Period-dependent near-field Raman scattering intensity and elastically scattered s42 in the center of the MIM subwavelength structure. (d) The strong coupling between GPP and SPP modes. The color represents the maximum SERS EF of MIM subwavelength unit.

3.3 Strong coupling phenomenon

The maximum SERS EF takes place around the period of 250 nm, and the SERS EF curve matches well with the curve of s42, shown in Figure 3(c). As the period becomes smaller, the coupling between adjacent GPP modes will be stronger, which results in the blueshift of the resonant wavelength of GPP modes [30]. When the period becomes so small as to be close to the mean free path of electron gas, the probability of collisions will increase and spatial non-locality emerges [31, 32], causing additional losses and leading the SERS EF to drop. The case of isolated GPP mode is close to that of very large period. Within above two limits, SPP modes are involved, altering original resonant wavelength of GPP modes. The mode coupling process can be described by coupled mode theory. With the increase of the ratio of coupling strength to the difference of damping rates, the coupling type turns from weak coupling to strong coupling. From Equation (2), reversible exchange of energy, i.e. Rabi oscillation arises between the modes, and the system’s damping rates will degenerate into the average damping rate in the strong coupling regime (Section 3 in Supporting Information). Figure 3(d) shows a significant anticrossing behavior between GPP and SPP modes, forming two new hybrid polariton bands with a Rabi splitting energy of 408.5 meV, which indicates a strong coupling phenomenon (Figure S3) [33, 34]. From Figure 3(d), the loss-exchange provides an additional enhancement of F2 ∼ 380 at the period of 250 nm by comparing with the EF at a very large period. Such enhancement benefits from the relative low damping rate of SPP mode. In addition, the SERS EF is also very sensitive to the width of MIM subwavelength structure (Figure S4).

3.4 Finite-size effects

For finite-size array structures, additional enhancement F2 changes with N and varies from one MIM subwavelength unit to another. Figure 4(a) shows the calculated maximum SERS EF on surface with N for both the central unit and whole array, which reveals an oscillation feature. Here the geometric parameters of the MIM subwavelength unit are fixed as 20 nm in width, 35 nm in depth and 250 nm in spacing. Most notably, the apparent oscillations for small N produce large additional enhancement F2. The oscillation of both tends to be stable when N become large, so that a finite array of MIM subwavelength units can be practically regarded as a periodic one as long as N > 20. When N > 10, oscillation period of N is typically around 4. The additional enhancement F2 of maximum SERS EF are around 380 for central unit and 645 for whole array respectively. The difference between above two values can be ascribed to edge effect, in which the edge provides a channel for energy leakage, making stronger EM field [35]. It is noteworthy that the effective polarizability and resonant wavelength of the edge units are also different from those in the middle [36].

Figure 4: 
Finite-size effects. (a) Calculated SERS EF with N. (b) Measured and calculated maximum SERS EF of central unit with N. (c) Measured and calculated near-field distribution along the x-axis of an array with 5 MIM subwavelength units. (d) Measured and calculated near-field distribution along the x-axis of an array with 8 MIM subwavelength units.
Figure 4:

Finite-size effects. (a) Calculated SERS EF with N. (b) Measured and calculated maximum SERS EF of central unit with N. (c) Measured and calculated near-field distribution along the x-axis of an array with 5 MIM subwavelength units. (d) Measured and calculated near-field distribution along the x-axis of an array with 8 MIM subwavelength units.

In our experiments, the maximum SERS EF of the central unit was collected for arrays with different number of units N, as shown in Figure 4(b). Figure 4(b) shows the agreement between experiment and calculation, which can be further confirmed (Figure S5). It can be seen that the maximum SERS EF of both reaches its maximum when N = 5, whose near-field spatial distribution is shown in Figure 4(c). When N ≤ 5, the largest SERS EF appears in the central unit, which is more than 800 times than that of an isolated one. From Figure 4(b), the EF of middle unit reaches valley point when N = 8, whose near-field spatial distribution is shown in Figure 4(d). In this case, the largest EF does not exist in the middle unit, but move towards the boundary units.

3.5 EM confinement in time and in space

The strong coupling between GPP and SPP modes accounts for localized near-field enhancement by confining EM waves in space and in time. The confinement characteristics of cavity quantum electrodynamics provide valuable insights into the light–matter interaction process of the SERS. The Q-factor governs the temporal confinement while V the spatial confinement. Q-factor represents the damping rate of the resonance mode, and is an important indicator in non-Hermitian physical processes [13, 37, 38]. For MIM subwavelength structures, we calculated scattering cross section of the central unit for different numbers of units N [37, 38], as shown in Figure 5(a). It shows that prominent resonance peaks near 532 nm and the resonance blueshifts with N, which is similar to the case of reducing the period shown in Figure 3(d). Figure 5(b) show the Q-factor extracted from the scattering cross section varies with N. For N ≤ 5, the Q-factor increases abruptly with N. When N = 5, the Q-factor is already close to that of infinite structures. Very small V gives birth to enormous EM field intensity and SERS EF. Figure 5(c) shows the calculated V of the central unit with N [39]. All the minimum versus situate at around 532 nm, which means that the finite MIM subwavelength arrays resonate at around 532 nm, with strong EM field confinement in space. Figure 5(d) shows the calculated V at 532 nm for different N, in which V oscillates with N. For small N, V decreases significantly with N, and reaches the minimum when N = 5. Therefore, compared to isolated MIM subwavelength structure, an array with a few MIM subwavelength units will increase Q-factor, reduce V, and thus boost the EM field confinement of the central unit, being in line with the SERS EF shown in Figure 4(b).

Figure 5: 
EM confinement in time and in space. (a) Calculated scattering cross section with N. (b) Calculated Q-factor with N at 532 nm. (c) Calculated V with wavelength and N. (d) Calculated V with N at 532 nm. The geometric parameters of MIM subwavelength unit are 20 nm in width, 35 nm in depth, and 250 nm in spacing respectively.
Figure 5:

EM confinement in time and in space. (a) Calculated scattering cross section with N. (b) Calculated Q-factor with N at 532 nm. (c) Calculated V with wavelength and N. (d) Calculated V with N at 532 nm. The geometric parameters of MIM subwavelength unit are 20 nm in width, 35 nm in depth, and 250 nm in spacing respectively.

4 Conclusions

In conclusion, we have taken the GPP mode as the isolated resonance, and specially investigated the contribution of strong coupling between GPP and SPP modes in finite and infinite arrays to the SERS EF using the s-SNOM. The spatial distribution of SERS EF transforms with the spacing and number of MIM subwavelength units. By virtue of strong coupling, periodically arranged MIM subwavelength structures provide more than 645 times additional SERS EF compared to an isolated one. Nevertheless, it is not necessary to utilize the array with a large number of MIM subwavelength units. The additional SERS EF of the array composed of only 5 units can be more than 800. Our results demonstrate that finite arrays with strong coupling effect will facilitate the development of high-performance devices, such as SERS substrates for single-molecule detection.


Corresponding author: Xuejin Zhang, National Laboratory of Solid State Microstructures, Collaborative Innovation Center of Advanced Microstructures, Key Laboratory of Intelligent Optical Sensing and Manipulation, Jiangsu Key Laboratory of Artificial Functional Materials, College of Engineering and Applied Sciences, Nanjing University, No. 22, Hankou Road, 210093 Nanjing, P.R. China, E-mail:

Funding source: National Natural Science Foundation of China

Funding source: National Key R&D Program of China

Award Identifier / Grant number: 2017YFA0303700

Award Identifier / Grant number: 2018YFA0306200

Award Identifier / Grant number: 2020YFA0211404

Award Identifier / Grant number: 2022YFA1405004

Funding source: Natural Science Foundation of Jiangsu Province, Major Project

Award Identifier / Grant number: BK20212004

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This work was supported by the National Key R&D Program of China (2022YFA1405004, 2020YFA0211404, 2018YFA0306200, and 2017YFA0303700), the National Natural Science Foundation of China (11274159 and 11374150), and the Natural Science Foundation of Jiangsu Province, Major Project (BK20212004).

  3. Conflict of interest statement: The authors declare no competing financial interest.

References

[1] Y. C. Cao, R. C. Jin, and C. A. Mirkin, “Nanoparticles with Raman spectroscopic fingerprints for DNA and RNA detection,” Science, vol. 297, no. 5586, pp. 1536–1540, 2002. https://doi.org/10.1126/science.297.5586.1536.Search in Google Scholar PubMed

[2] K. A. Willets and R. P. V. Duyne, “Localized surface plasmon resonance spectroscopy and sensing,” Annu. Rev. Phys. Chem., vol. 58, pp. 267–297, 2007. https://doi.org/10.1146/annurev.physchem.58.032806.104607.Search in Google Scholar PubMed

[3] J. S. Huang, V. Callegari, P. Geisler, et al.., “Atomically flat single-crystalline gold nanostructures for plasmonic nanocircuitry,” Nat. Commun., vol. 1, p. 150, 2010. https://doi.org/10.1038/ncomms1143.Search in Google Scholar PubMed

[4] F. Benz, C. Tserkezis, L. O. Herrmann, et al.., “Nanooptics of molecular-shunted plasmonic nanojunctions,” Nano Lett., vol. 15, no. 1, pp. 669–674, 2015. https://doi.org/10.1021/nl5041786.Search in Google Scholar PubMed PubMed Central

[5] R. W. Taylor, F. Benz, D. O. Sigle, et al.., “Watching individual molecules flex within lipid membranes using SERS,” Sci. Rep., vol. 4, p. 5940, 2014. https://doi.org/10.1038/srep05940.Search in Google Scholar PubMed PubMed Central

[6] S. L. Kleinman, E. Ringe, N. Valley, et al.., “Single-molecule surface-enhanced Raman spectroscopy of crystal violet isotopologues: theory and experiment,” J. Am. Chem. Soc., vol. 133, no. 11, pp. 4115–4122, 2011. https://doi.org/10.1021/ja110964d.Search in Google Scholar PubMed

[7] J. Kneipp, H. Kneipp, and K. Kneipp, “SERS—a single-molecule and nanoscale tool for bioanalytics,” Chem. Soc. Rev., vol. 37, no. 5, pp. 1052–1060, 2008. https://doi.org/10.1039/b708459p.Search in Google Scholar PubMed

[8] X. M. Qian and S. M. Nie, “Single-molecule and single-nanoparticle SERS: from fundamental mechanisms to biomedical applications,” Chem. Soc. Rev., vol. 37, no. 5, pp. 912–920, 2008. https://doi.org/10.1039/b708839f.Search in Google Scholar PubMed

[9] S. M. Nie and S. R. Emory, “Probing single molecules and single nanoparticles by surface-enhanced Raman scattering,” Science, vol. 275, no. 5303, pp. 1102–1106, 1997. https://doi.org/10.1126/science.275.5303.1102.Search in Google Scholar PubMed

[10] K. Kneipp, Y. Wang, H. Kneipp, et al.., “Single molecule detection using surface-enhanced Raman scattering (SERS),” Phys. Rev. Lett., vol. 78, pp. 1667–1670, 1997. https://doi.org/10.1103/PhysRevLett.78.1667.Search in Google Scholar

[11] L. R. Yuan, K. Qin, J. Tan, et al.., “Plasmonic band-edge modulated surface enhanced Raman scattering,” Appl. Phys. Lett., vol. 111, p. 051601, 2017. https://doi.org/10.1063/1.4997303.Search in Google Scholar

[12] J. L. Perchec, P. Quémerais, A. Barbara, and T. López-Ríos, “Why metallic surfaces with grooves a few nanometers deep and wide may strongly absorb visible light,” Phys. Rev. Lett., vol. 100, p. 066408, 2008. https://doi.org/10.1103/physrevlett.100.066408.Search in Google Scholar

[13] H. Zhao and L. Feng, “Parity-time symmetry photonics,” Natl. Sci. Rev., vol. 5, no. 2, pp. 183–199, 2018. https://doi.org/10.1093/nsr/nwy011.Search in Google Scholar

[14] S. A. Maier, “Plasmonic field enhancement and SERS in the effective mode volume picture,” Opt. Express, vol. 14, pp. 1957–1964, 2006. https://doi.org/10.1364/oe.14.001957.Search in Google Scholar PubMed

[15] J. P. Reithmaier, G. Sek, A. Loffler, et al.., “Strong coupling in a single quantum dot-semiconductor microcavity system,” Nature, vol. 432, pp. 197–200, 2004. https://doi.org/10.1038/nature02969.Search in Google Scholar PubMed

[16] I. Yoshie, A. Scherer, J. Hendrichson, et al.., “Vacuum Rabi splitting with a single quantum dot in a photonic crystal nanocavity,” Nature, vol. 432, pp. 200–203, 2004. https://doi.org/10.1038/nature03119.Search in Google Scholar PubMed

[17] T. Wu, M. Gurioli, and P. Lalanne, “Nanoscale light confinement: the Q’s and V’s,” ACS Photonics, vol. 8, pp. 1522–1538, 2021. https://doi.org/10.1021/acsphotonics.1c00336.Search in Google Scholar

[18] Y. Zhang and Z. Zhang, “Ultra-subwavelength and low loss in v-shaped hybrid plasmonic waveguide,” Plasmonics, vol. 12, no. 1, pp. 59–63, 2017. https://doi.org/10.1007/s11468-016-0228-z.Search in Google Scholar

[19] D. K. Lim, K. S. Jeon, H. M. Kim, J. M. Nam, and Y. D. Suh, “Nanogap-engineerable Raman-active nanodumbbells for single-molecule detection,” Nat. Mater., vol. 9, pp. 60–67, 2010. https://doi.org/10.1038/nmat2596.Search in Google Scholar PubMed

[20] P. Kusch, S. Mastel, N. S. Mueller, et al.., “Dual-scattering near-field microscope for correlative nanoimaging of SERS and electromagnetic hotspots,” Nano Lett., vol. 17, pp. 2667–2673, 2017. https://doi.org/10.1021/acs.nanolett.7b00503.Search in Google Scholar PubMed

[21] A. Garcia-Etxarri, I. Romero, F. J. Garcia de Abajo, R. Hillenbrand, and J. Aizpurua, “Influence of the tip in near-field imaging of nanoparticle plasmonic modes: weak and strong coupling regimes,” J. Phys. Rev. B Condens., vol. 79, no. 12, p. 125439, 2009. https://doi.org/10.1103/physrevb.79.125439.Search in Google Scholar

[22] P. Alonso-González, P. Albella, M. Schnell, et al.., “Resolving the electromagnetic mechanism of surface-enhanced light scattering at single hot spots,” Nat. Commun., vol. 3, pp. 684, 2012. https://doi.org/10.1038/ncomms1674.Search in Google Scholar PubMed PubMed Central

[23] R. Hillenbrand and F. Keilmann, “Complex optical constants on a subwavelength scale,” Phys. Rev. Lett., vol. 85, no. 14, pp. 3029–3032, 2000. https://doi.org/10.1103/physrevlett.85.3029.Search in Google Scholar

[24] N. Ocelic, A. Huber, and R. Hillenbrand, “Pseudoheterodyne detection for background-free near-field spectroscopy,” Appl. Phys. Lett., vol. 89, p. 101124, 2006. https://doi.org/10.1063/1.2348781.Search in Google Scholar

[25] Z. H. Kim and S. R. Leone, “Polarization-selective mapping of near-field intensity and phase around gold nanoparticles using apertureless near-field microscopy,” Opt. Express, vol. 16, pp. 1733–1741, 2008. https://doi.org/10.1364/oe.16.001733.Search in Google Scholar PubMed

[26] M. Schnell, A. García-Etxarri, A. J. Huber, K. B. Crozier, J. Aizpurua, and R. Hillenbrand, “Controlling the near-field oscillations of loaded plasmonic nanoantennas,” Nat. Photonics, vol. 3, pp. 287–291, 2009. https://doi.org/10.1038/nphoton.2009.46.Search in Google Scholar

[27] M. Schnell, A. García-Etxarri, A. J. Huber, et al.., “Amplitude- and phase-resolved near-field mapping of infrared antenna modes by transmission-mode scattering-type near-field microscopy,” J. Phys. Chem. C, vol. 114, pp. 7341–7345, 2010. https://doi.org/10.1021/jp909252z.Search in Google Scholar

[28] C. Y. Wang, H. Y. Chen, L. Y. Sun, et al.., “Giant colloidal silver crystals for low-loss linear and nonlinear plasmonics,” Nat. Commun., vol. 6, p. 7734, 2015. https://doi.org/10.1038/ncomms8734.Search in Google Scholar PubMed PubMed Central

[29] P. B. Johnson and R. W. Christy, “Optical constant of the noble metals,” Phys. Rev. B, vol. 6, no. 12, pp. 4370–4379, 1972. https://doi.org/10.1103/physrevb.6.4370.Search in Google Scholar

[30] W. C. Tan, T. W. Preist, J. R. Sambles, and N. P. Wanstall, “Flat surface-plasmon-polariton bands and resonant optical absorption on short-pitch metal gratings,” Phys. Rev. B, vol. 59, no. 19, pp. 12661–12666, 1999. https://doi.org/10.1103/physrevb.59.12661.Search in Google Scholar

[31] K. Zhao, H. X. Xu, B. H. Gu, and Z. Y. Zhang, “One-dimensional arrays of nanoshell dimers for single molecule spec-troscopy via surface enhanced Raman scattering,” J. Chem. Phys., vol. 125, p. 081102, 2006. https://doi.org/10.1063/1.2229204.Search in Google Scholar PubMed

[32] G. Toscano, S. Raza, S. Xiao, et al.., “Surface-enhanced Raman spectroscopy: nonlocal limitations,” Opt. Lett., vol. 37, no. 13, pp. 2538–2540, 2012. https://doi.org/10.1364/ol.37.002538.Search in Google Scholar

[33] L. Novotny, “Strong coupling, energy splitting, and level crossings: a classical perspective,” Am. J. Phys., vol. 78, pp. 1199–1202, 2010. https://doi.org/10.1119/1.3471177.Search in Google Scholar

[34] P. Törmä and W. L. Barnes, “Strong coupling between surface plasmon polaritons and emitters: a review,” Rep. Prog. Phys., vol. 78, p. 013901, 2015. https://doi.org/10.1088/0034-4885/78/1/013901.Search in Google Scholar PubMed

[35] R. Heilmann, G. Salerno, J. Cuerda, T. K. Hakala, and P. Törmä, “Quasi-BIC mode lasing in a quadrumer plasmonic lattice,” ACS Photonics, vol. 9, pp. 224–232, 2022. https://doi.org/10.1021/acsphotonics.1c01416.Search in Google Scholar PubMed PubMed Central

[36] L. Zundel and A. Manjavacas, “Finite-size effects on periodic arrays of nanostructures,” J. Phys. Photonics, vol. 1, p. 015004, 2019. https://doi.org/10.1088/2515-7647/aae8a2.Search in Google Scholar

[37] S. Derom, R. Vincent, A. Bouhelier, and C. des Francs, “Resonance quality, radiative/ohmic losses and modal volume of Mie plasmons,” G. EPL, vol. 98, p. 47008, 2012. https://doi.org/10.1209/0295-5075/98/47008.Search in Google Scholar

[38] T. Christopoulos, O. Tsilipakos, G. Sinatkas, and E. E. Kriezis, “On the calculation of the quality factor in contemporary photonic resonant structures,” Opt. Express, vol. 27, pp. 14505–14522, 2019. https://doi.org/10.1364/oe.27.014505.Search in Google Scholar

[39] K. G. Cognée, W. Yan, F. La China, et al.., “Mapping complex mode volumes with cavity perturbation theory,” Optica, vol. 6, pp. 269–273, 2019. https://doi.org/10.1364/optica.6.000269.Search in Google Scholar


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/nanoph-2023-0016).


Received: 2023-01-09
Accepted: 2023-03-19
Published Online: 2023-03-30

© 2023 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 3.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2023-0016/html
Scroll to top button