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

Strong and weak couplings in molecular vibration–plasmon hybrid structures

Open Access Open Access

Abstract

Molecular vibration–plasmon couplings in a hybrid structure, which are composed of a silver grating filled with polymethyl methacrylate (PMMA) molecules (SG–PMMA), have been investigated theoretically. It is found that the interaction between the vibrational transitions and plasmons can transform from weak coupling into strong coupling by reducing the distance between the elements. When the space between grating elements is large, the localized surface plasmon resonance (LSP) of the silver elements greatly enhances the absorption of the PMMA molecules. As the gap between elements becomes small, the plasmonic nanocavity (NC) mode emerges and couples strongly with the molecular vibrational mode of PMMA. The strong coupling results in two new hybridized modes and the Rabi splitting energy is about 15 meV. Our work provides an effective way to alter the coupling strength of the molecular vibration-plasmon hybrid system and may be beneficial to the further biochemical and biophysical applications.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In the past decade, couplings between excitonic materials and metal nanostructures have attracted increasing interest because of their potential applications in the biological and chemical sensing [1–4], surface enhanced spectroscopy [5,6], quantum information processing [7,8], and solar cells [9,10]. In general, optical responses in metal nanostructures include localized surface plasmon (LSP), surface plasmon polariton (SPP), and plasmonic nanocavity (NC) [11], whereas excitons mainly exist in semiconductor materials [12], quantum wells [13], quantum dots [14], biomolecules [15], etc. There are two types of coupling: weak coupling and strong coupling [16]. The weak coupling often enhances absorption or light-emitting of the excitonic materials [17–21]. In the strong coupling regime, new hybrid modes are formed and separated by the Rabi splitting [22,23]. Such new modes have different optical properties and have been widely used in quantum cryptography [24], nanoscale lasers [25], plasmon-enhanced light harvesting [26], and chemical reaction promotion [27].

Similar to the exciton–plasmon interactions in visible, couplings between molecular vibrational transitions and plasmons have received great attention recently due to fundamental and practical interests [22,23,28]. Cetin et al. [22] reported that two vibrational modes of PMMA molecules can couple with two plasmonic modes, providing more reliable sensing information. Wan et al. [23] not only achieved the strong couplings between the LSP mode and the PMMA vibrational mode, but also demonstrated that the coupling strength can be modulated by changing the thickness of the PMMA film and hence the forbidden energy gap due to the mode splitting could be adjusted accordingly. Memmi et al. [28] demonstrated that the strong coupling between SPPs and the vibration of the ketone-based polymer arises new eigenstates, which hold strong potential for applications in chemistry and optoelectronics. However, the weak couplings in the molecular vibration–plasmon structure were seldom studied, especially the mutual transformation between the weak coupling and the strong coupling was not reported. As we know, there are various plasmonic modes which are closely related with metal structure parameters. There is great opportunity that coupling regimes can be changed by utilizing different plasmonic modes in the same system. It was found in some exciton–plasmon couplings in visible region that the coupling strength could be controlled by the geometry [18], exciton concentrations [29,30] and metal oxidation [31], etc.

In this paper, we investigate the mid-infrared molecular vibration–plasmon coupling in a hybrid structure, consisting of a silver grating filled with polymethyl methacrylate (PMMA) molecules (SG–PMMA). The localized plasmon resonance could be modulated to match with the vibrational band (C = O) of PMMA molecules by tuning the height of the grating elements. We further study the influences of the grating structure parameters on the molecular vibration–plasmon couplings in the SG–PMMA. It is surprising that the molecular vibration–plasmon coupling transfers from the weak coupling into the strong coupling by reducing the distance between adjacent elements. These couplings are different from those in the traditional cavity quantum electrodynamics (CQED) systems, which is essential for interesting physics in quantum information science [32,33]. Our system is semi-classical and can be studied by the coupled oscillator model [34,35].

2. Molecular vibration–plasmon hybrid model

Figure 1(a) shows the cross-section schematic of the SG–PMMA hybrid structure. The period and element width of the 1–D silver grating are defined as P and w, respectively. The height of the grating elements is h and PMMA molecules are filled in grooves. The dielectric function of silver could be described by the Drude model εAg=εωp2/(ω2+iγ1ω), where ω is the angular frequency, the plasma frequency ωp = 9 eV, and the damping constant γ1 = 0.1 eV. The high frequency constant term ε is set as 4.6 [36]. The permittivity of the PMMA medium could be modeled by the Lorentz oscillator model εPMMA=ε2Aω02/(ω2ω02+iγ2ω) [22], where background permittivity ε2 = 2.36, oscillator strength A = 0.0115, Lorentz resonance frequency ω0 = 0.215 eV (5.77 μm), and damping frequency γ2 = 0.00245 eV. These parameters make the model fit the reported experimental data well [22]. The Lorentz model means there is a narrow absorption band around ω0, but the absorption is weak.

 figure: Fig. 1

Fig. 1 (a) Schematic of the cross section of SG-PMMA hybrid structure. (b) Absorption spectra of the SG–PMMA, the silver grating, and the PMMA molecules in the grooves, respectively. Here, the period P = 3000 nm, the element width w = 2000 nm, and the height of the elements h = 775 nm.

Download Full Size | PDF

Finite element method (FEM) has been performed to calculate the optical responses and electric field distributions of the SG–PMMA hybrid structure based on the software COMSOL Multiphysics. A TM plane wave with electric field E0 normally impinges onto the SG–PMMA hybrid structure. The periodic boundary conditions are employed on both left and right to account for the periodic arrangement. Perfectly matched layers are set at the top and bottom to avoid multiple reflections.

3. Results and discussion

In Fig. 1(b), the solid, dashed, and dotted lines represent the absorption spectra of the SG–PMMA, the silver grating filled with dielectric material of ℇ2 = 2.36, and the PMMA molecules, respectively. Here, the period P = 3000 nm, the element width w = 2000 nm, and the element height h = 775 nm. The absorption is defined as A = 1–RT, where R is the reflection and T is the transmission. Two virtual boundaries Btop and Bbottom are drawn in the upstate and substrate to calculate the reflection R and transmission T, respectively. Each of them is 50 nm away from the PMLs to avoid numerical errors [37]. The thickness of the silver substrate is much larger than the skin depth of the silver, which makes the transmission T = 0. The pure PMMA molecules are kept as many as those in the hybrid structure. There is a weak absorption peak at about 5.77 μm for PMMA molecules due to vibrational transitions. We observe two plasmon modes in the absorption spectrum of the silver grating. One peak at around 3 μm is the SPP mode, which is directly determined by the grating period under normal incidence. The reciprocal grating vector provides the wavenumber for SPP kSPP = 2mπ/P [6]. Here m = 1. The broad peak at around 5.77 μm arises from the LSPs of the grating elements [38]. This mode originates from the collective oscillations of free electrons in metallic elements and is determined by the element width w and the height h [39]. Here, the values of w and h are well designed so that the LSP matches with the molecular vibrational mode of PMMA. As the dielectric material in the silver grating is replaced by PMMA molecules, the absorption around 5.77 μm is increased to nearly 1 and the other peak at about 3 μm is almost unchanged. Even if the metal absorption from the silver grating is subtracted, the remaining absorption at 5.77 μm is 7 times larger than that of the pure PMMA molecules. Usually, such molecule absorption enhancement results from the weak couplings between the molecular vibrational transitions and the plasmon resonances [40].

Then we systematically investigate the absorption enhancement in the SG–PMMA hybrid structure to clarify the weak coupling. Figure 2(a) shows the contour plot of the absorption spectra of the silver grating as a function of grating element height h. Here, the period P and width w are fixed at 3000 and 2000 nm, respectively. As the h-value increases from 600 to 900 nm, the LSP redshifts almost lineally from about 4.7 to 6.5 μm, moving across the vibrational mode of PMMA (~5.77μm). As the h-value increases, the relaxation effect in metal elements grows and hence the LSP shows a redshift accordingly [39]. Figure 2(b) represents the contour plot of the absorption spectra of the SG–PMMA as a function of h-value. The dashed and dotted lines denote the uncoupled LSP mode and the molecular vibrational mode, respectively. It is observed that the spectrum position of the absorption peak at around 5.77 μm isn’t affected by the approaching LSP mode. This means the molecular vibrational transition and the plasmon do not exchange energies and their interaction is definitely in the weak coupling regime. The absorption of the SG–PMMA at around 5.77 μm increases first and then decreases with the increase of the h-value, as shown in Fig. 2(c). At h = 775 nm, the LSP appears at about 5.77 μm while the local electric fields near the grating elements is the largest, resulting in the greatest absorption.

 figure: Fig. 2

Fig. 2 (a) Contour plot of the absorption spectra of the silver grating as a function of element height h. (b) Contour plot of the absorption spectra of the SG–PMMA as a function of h. The dashed and dotted lines represent variations of the LSP mode and molecular vibrational mode, respectively. (c) Maximal absorption of the SG–PMMA at 5.77 μm as a function of h. Here, the period P and width w are fixed at 3000 nm and 2000 nm, respectively.

Download Full Size | PDF

Next, we focus on the influence of the grating period P on the plasmon modes of the silver grating. Figure 3(a) shows the absorption spectra of the silver gratings with period P of 3000 nm (solid line), 2700 nm (short dashed line), 2400 nm (short dotted line). Here the element width w and height h are fixed at 2000 and 700 nm, respectively. It is observed that by decreasing the P-value from 3000 to 2400 nm the SPP of the silver grating shows a distinct blue-shift from around 3 to 2.4 μm, which follows the resonance condition of the SPP [6]. Meanwhile, the LSP shows a negligible blue–shift. The LSP is mainly determined by the element width and height and the interaction between neighboring elements causes this weak shift [41]. Figure 3(b) shows the near-field distribution of the silver grating with P = 3000 nm at the LSP resonance (5.30 μm). The electric fields are strongly localized at the top corners and extend outward, which is one feature of the LSP reported in [29]. With the decreasing P-value, the distance between grating elements grows narrow as the element width w is fixed at 2000 nm. When the groove width becomes very small, plasmonic nanocavity (NC) mode of the silver grating is achieved and generates stronger localized field enhancements [11,42]. Figure 3(c) shows the absorption spectra of the silver gratings with P-values of 2100, 2090, and 2080 nm. The SPP at around 2.1 μm is almost unchanged because of the very small variation in the P-value. On the other hand, another peak red shifts with the decrease of P-value, which means this mode is different from the LSP mode in Fig. 3(a). Strong interactions between neighboring grating elements form the NC mode, which is related with the shape of the cavity. As the groove height h is fixed here, narrower groove leads the resonance to move towards longer wavelength [43,44]. Figure 3(d) shows the electric field distribution of the silver grating with P = 2100 nm at the NC resonance of 5.73 μm. The cavity between grating elements is filled with higher electric fields compared with Fig. 3(b). Also, it has strongest electric field at the mouth of the groove and the weakest field at the bottom [45]. Therefore, by decreasing the grating period, the LSP mode of grating elements could be changed to the NC mode between grating elements, accompanied with higher electric fields in grooves. By this way, the light-matter interactions should be strengthened.

 figure: Fig. 3

Fig. 3 (a) Absorption spectra of the silver gratings with grating period P of 3000 nm (solid line), 2700 nm (short dashed line), 2400 nm (short dotted line). (b) Electric field distribution in the silver gratings with P = 3000 nm at resonance wavelength of 5.30 μm. (c) Absorption spectra of the silver gratings with grating period P of 2100 nm (solid line), 2090 nm (short dashed line), 2080 nm (short dotted line). (d) Electric field distribution in the silver grating with P = 2100 nm at resonance wavelength of 5.73 μm. Here, the element width w and element height h are fixed at 2000 nm and 700 nm, respectively.

Download Full Size | PDF

Finally, we study the couplings between the molecular vibrational mode and the plasmonic NC mode in the SG–PMMA. Figure 4(a) shows the contour plot of the absorption spectra of the uncoupled silver grating as a function of the element height h. Here, the element width w and period P is fixed at 2000 and 2100 nm, respectively. When the h-value increases from 610 to 810 nm, the NC mode of the silver grating redshifts from 4.9 μm to 6.8 μm almost linearly. The NC mode is mainly determined by the ratio of the cavity [43,44]. Since the width of the gap is fixed, it red shifts with the increase of the gap depth h. Figure 4(b) shows the contour plot of the absorption spectra of the SG–PMMA as a function of h. The uncoupled NC modes and molecular vibrational modes are represented by the dash and dotted lines, respectively. When the h-value is small, the broad NC mode is far away from the molecular vibrational mode and hence the interaction between these two modes results in a Fano-like resonance, i.e., the Fano lineshape occurs. As the h-value increases to around 710 nm, the strong molecular vibration–plasmon coupling results in two new hybrid modes [46] and the Rabi splitting happens, as shown in Fig. 4(b). Such transition from Fano resonance to Rabi splitting has been carefully investigated in previous papers [47,48]. The solid and dash-dot lines show the hybridized high-energy mode (HM) and low-energy mode (LM), respectively. There is an obvious anti-crossing behavior and the obtained Rabi splitting energy ΩRabi is about 15 meV. The full-width at half-maximum (FWHM) of the uncoupled plasmon resonance and molecular vibrational transition are Γplasmon ≈22.90 meV and ΓPMMA ≈2.45 meV, respectively. It is known that the Rabi splitting occurs when the Rabi splitting energy ΩRabi is larger than (Γplasmon + ΓPMMA)/2 [49]. Based on the above calculations, we can confirm that the coupling in Fig. 4 can satisfy this criterion, as (Γplasmon + ΓPMMA)/2 ≈12.68 meV. Furthermore, we investigate the strong coupling strength in the SG–PMMA structure, which can be studied by the classic coupled oscillator model [34]:

(E1+iΓplasmon/2ggE2+iΓPMMA/2)(αβ)=E(αβ),
where E1 and E2 are the energies of the NC and molecular vibrational modes, respectively. E is the eigenvalues corresponding to the energies of the hybrid modes. α and β construct the eigenvectors, which represent the weighting coefficients of the original modes (|α|2 + |β|2 = 1). g represents the strong coupling strength. From Eq. (1), we can obtain two eigenvalues and the difference between them at E1 = E2 is the Rabi splitting ΩRabi=2g2(ΓplasmonΓPMMA)2/16. Then, the obtained coupling strength is g = 9.1 meV, which satisfies the strong coupling condition g2>(Γplasmon2+ΓPMMA2)/16 [35]. As reviewed in previous papers, the strong molecular vibration–plasmon interactions hold a strong potential for applications in chemistry and optoelectronics.

 figure: Fig. 4

Fig. 4 (a) Contour plot of the absorption spectra of the silver grating as a function of h. (b) Contour plot of the absorption spectra of the SG–PMMA hybrid structure as a function of h. Here, the period P and width w are fixed at 2100 nm and 2000 nm, respectively. The solid and dash-dot lines show the hybridized HM and LM, respectively. The dashed and dotted lines represent the uncoupled nanocavity mode (NC) of the silver grating and molecular vibrational mode of PMMA, respectively.

Download Full Size | PDF

4. Conclusion

We have investigated the molecular vibration–plasmon couplings in the SG–PMMA hybrid structure in mid-infrared region. When the period of the silver grating is large, the weak coupling between the LSP of grating elements and PMMA molecular vibration could greatly enhance the absorption of the PMMA molecules. The maximal absorption enhancement is more than 7 times when the LSP matches with the molecular vibrational mode. As the grating period decreases and hence the distance between adjacent elements becomes small, the LSP mode turns into the plasmonic nanocavity mode. This plasmonic mode introduces higher local electric fields and promotes light-matter interactions greatly in the SG-PMMA hybrid structure. Strong couplings between molecular vibrations and plasmon NC modes are achieved and the obtained Rabi splitting energy is about 15 meV. The tunable couplings in this molecular vibration–plasmon hybrid structure may be beneficial to their further biochemical and biophysical applications.

Funding

National Natural Science Foundation of China (NSFC) (11874222, 11674175, 11834008); Major Project of Nature Science Research for Colleges and Universities in Jiangsu Province (15KJA140002); “333” Project of Jiangsu Province (BRA2017451); Program of Natural Science Research for Colleges and Universities in Jiangsu Province (17KJB510032)

References

1. K. A. Willets and R. P. Van Duyne, “Localized surface plasmon resonance spectroscopy and sensing,” Annu. Rev. Phys. Chem. 58(1), 267–297 (2007). [CrossRef]   [PubMed]  

2. A. Hatef, S. M. Sadeghi, É. Boulais, and M. Meunier, “Quantum dot-metallic nanorod sensors via exciton-plasmon interaction,” Nanotechnology 24(1), 015502 (2013). [CrossRef]   [PubMed]  

3. S. Zeng, D. Baillargeat, H. P. Ho, and K. T. Yong, “Nanomaterials enhanced surface plasmon resonance for biological and chemical sensing applications,” Chem. Soc. Rev. 43(10), 3426–3452 (2014). [CrossRef]   [PubMed]  

4. J. Martínez, A. Ródenas, M. Aguiló, T. Fernandez, J. Solis, and F. Díaz, “Mid-infrared surface plasmon polariton chemical sensing on fiber-coupled ITO coated glass,” Opt. Lett. 41(11), 2493–2496 (2016). [CrossRef]   [PubMed]  

5. G. Dayal, X. Y. Chin, C. Soci, and R. Singh, “High-Q plasmonic Fano resonance for multiband surface-enhanced infrared absorption of molecular vibrational sensing,” Adv. Opt. Mater. 5(2), 1600559 (2017). [CrossRef]  

6. T. Wang, V. H. Nguyen, A. Buchenauer, U. Schnakenberg, and T. Taubner, “Surface enhanced infrared spectroscopy with gold strip gratings,” Opt. Express 21(7), 9005–9010 (2013). [CrossRef]   [PubMed]  

7. N. P. de Leon, B. J. Shields, C. L. Yu, D. E. Englund, A. V. Akimov, M. D. Lukin, and H. Park, “Tailoring light-matter interaction with a nanoscale plasmon resonator,” Phys. Rev. Lett. 108(22), 226803 (2012). [CrossRef]   [PubMed]  

8. M. L. Andersen, S. Stobbe, A. S. Sørensen, and P. Lodahl, “Strongly modified plasmon–matter interaction with mesoscopic quantum emitters,” Nat. Phys. 7(3), 215–218 (2011). [CrossRef]  

9. A. Bharti, R. Bhardwaj, A. K. Agrawal, N. Goyal, and S. Gautam, “Monochromatic X-ray induced novel synthesis of plasmonic nanostructure for photovoltaic application,” Sci. Rep. 6(1), 22394 (2016). [CrossRef]   [PubMed]  

10. C. Hägglund, S. P. Apell, and B. Kasemo, “Maximized optical absorption in ultrathin films and its application to plasmon-based two-dimensional photovoltaics,” Nano Lett. 10(8), 3135–3141 (2010). [CrossRef]   [PubMed]  

11. A. Polyakov, K. F. Thompson, S. D. Dhuey, D. L. Olynick, S. Cabrini, P. J. Schuck, and H. A. Padmore, “Plasmon resonance tuning in metallic nanocavities,” Sci. Rep. 2(1), 933 (2012). [CrossRef]   [PubMed]  

12. P. Bhattacharya, T. Frost, S. Deshpande, M. Z. Baten, A. Hazari, and A. Das, “Room temperature electrically injected polariton laser,” Phys. Rev. Lett. 112(23), 236802 (2014). [CrossRef]   [PubMed]  

13. J. M. Ménard, C. Poellmann, M. Porer, U. Leierseder, E. Galopin, A. Lemaître, A. Amo, J. Bloch, and R. Huber, “Revealing the dark side of a bright exciton-polariton condensate,” Nat. Commun. 5(1), 4648 (2014). [CrossRef]   [PubMed]  

14. J. P. Reithmaier, G. Sek, A. Löffler, C. Hofmann, S. Kuhn, S. Reitzenstein, L. V. Keldysh, V. D. Kulakovskii, T. L. Reinecke, and A. Forchel, “Strong coupling in a single quantum dot-semiconductor microcavity system,” Nature 432(7014), 197–200 (2004). [CrossRef]   [PubMed]  

15. R. Chikkaraddy, B. de Nijs, F. Benz, S. J. Barrow, O. A. Scherman, E. Rosta, A. Demetriadou, P. Fox, O. Hess, and J. J. Baumberg, “Single-molecule strong coupling at room temperature in plasmonic nanocavities,” Nature 535(7610), 127–130 (2016). [CrossRef]   [PubMed]  

16. E. Cao, W. Lin, M. Sun, W. Liang, and Y. Z. Song, “Exciton-plasmon coupling interactions: from principle to applications,” Nanophotonics 7(1), 145–167 (2018). [CrossRef]  

17. B. Sharma, R. R. Frontiera, A. I. Henry, E. Ringe, and R. P. Van Duyne, “SERS: Materials, applications, and the future,” Mater. Today 15(1–2), 16–25 (2012). [CrossRef]  

18. S. Balci, E. Karademir, C. Kocabas, and A. Aydinli, “Absorption enhancement of molecules in the weak plasmon-exciton coupling regime,” Opt. Lett. 39(17), 4994–4997 (2014). [CrossRef]   [PubMed]  

19. J. Lee, A. O. Govorov, J. Dulka, and N. A. Kotov, “Bioconjugates of CdTe nanowires and Au nanoparticles: plasmon−exciton interactions, luminescence enhancement, and collective effects,” Nano Lett. 4(12), 2323–2330 (2004). [CrossRef]  

20. X. Yang, H. Yu, X. Guo, Q. Ding, T. Pullerits, R. Wang, G. Zhang, W. Liang, and M. Sun, “Plasmon-exciton coupling of monolayer MoS2-Ag nanoparticles hybrids for surface catalytic reaction,” Mater. Today Energy 5, 72–78 (2017). [CrossRef]  

21. A. Kumar, R. Srivastava, P. Tyagi, D. S. Mehta, and M. N. Kamalasanan, “Efficiency enhancement of organic light emitting diode via surface energy transfer between exciton and surface plasmon,” Org. Electron. 13(1), 159–165 (2012). [CrossRef]  

22. A. E. Cetin, S. Korkmaz, H. Durmaz, E. Aslan, S. Kaya, R. Paiella, and M. Turkmen, “Quantification of multiple molecular fingerprints by dual-resonant perfect absorber,” Adv. Opt. Mater. 4(8), 1274–1280 (2016). [CrossRef]  

23. W. Wan, X. Yang, and J. Gao, “Strong coupling between mid-infrared localized plasmons and phonons,” Opt. Express 24(11), 12367–12374 (2016). [CrossRef]   [PubMed]  

24. H. K. Lo and H. F. Chau, “Unconditional security of quantum key distribution over arbitrarily long distances,” Science 283(5410), 2050–2056 (1999). [CrossRef]   [PubMed]  

25. M. A. Noginov, G. Zhu, A. M. Belgrave, R. Bakker, V. M. Shalaev, E. E. Narimanov, S. Stout, E. Herz, T. Suteewong, and U. Wiesner, “Demonstration of a spaser-based nanolaser,” Nature 460(7259), 1110–1112 (2009). [CrossRef]   [PubMed]  

26. S. D. Standridge, G. C. Schatz, and J. T. Hupp, “Distance dependence of plasmon-enhanced photocurrent in dye-sensitized solar cells,” J. Am. Chem. Soc. 131(24), 8407–8409 (2009). [CrossRef]   [PubMed]  

27. Q. Ding, Y. Shi, M. Chen, H. Li, X. Yang, Y. Qu, W. Liang, and M. Sun, “Ultrafast dynamics of plasmon−exciton interaction of Ag nanowire−graphene hybrids for surface catalytic reactions,” Sci. Rep. 6(1), 32724 (2016). [CrossRef]   [PubMed]  

28. H. Memmi, O. Benson, S. Sadofev, and S. Kalusniak, “Strong coupling between surface plasmon polaritons and molecular vibrations,” Phys. Rev. Lett. 118(12), 126802 (2017). [CrossRef]   [PubMed]  

29. S. Balci, B. Kucukoz, O. Balci, A. Karatay, C. Kocabas, and G. Yaglioglu, “Tunable plexcitonic nanoparticles: a model system for studying plasmon−exciton interaction from the weak to the ultrastrong coupling regime,” ACS Photonics 3(11), 2010–2016 (2016). [CrossRef]  

30. T. J. Antosiewicz, S. P. Apell, and T. Shegai, “Plasmon−exciton interactions in a core−shell geometry: from enhanced absorption to strong coupling,” ACS Photonics 1(5), 454–463 (2014). [CrossRef]  

31. F. Todisco, S. D’Agostino, M. Esposito, A. I. Fernández-Domínguez, M. De Giorgi, D. Ballarini, L. Dominici, I. Tarantini, M. Cuscuná, F. Della Sala, G. Gigli, and D. Sanvitto, “Exciton−plasmon coupling enhancement via metal oxidation,” ACS Nano 9(10), 9691–9699 (2015). [CrossRef]   [PubMed]  

32. T. Yoshie, A. Scherer, J. Hendrickson, G. Khitrova, H. M. Gibbs, G. Rupper, C. Ell, O. B. Shchekin, and D. G. Deppe, “Vacuum Rabi splitting with a single quantum dot in a photonic crystal nanocavity,” Nature 432(7014), 200–203 (2004). [CrossRef]   [PubMed]  

33. K. Hennessy, A. Badolato, M. Winger, D. Gerace, M. Atatüre, S. Gulde, S. Fält, E. L. Hu, and A. Imamoğlu, “Quantum nature of a strongly coupled single quantum dot-cavity system,” Nature 445(7130), 896–899 (2007). [CrossRef]   [PubMed]  

34. X. Liu, T. Galfsky, Z. Sun, F. Xia, E. Lin, Y. Lee, S. Kéna-Cohen, and V. M. Menon, “Strong light–matter coupling in two-dimensional atomic crystals,” Nat. Photonics 9(1), 30–34 (2015). [CrossRef]  

35. E. M. Roller, C. Argyropoulos, A. Högele, T. Liedl, and M. Pilo-Pais, “Plasmon−exciton coupling using DNA templates,” Nano Lett. 16(9), 5962–5966 (2016). [CrossRef]   [PubMed]  

36. A. Delga, J. Feist, J. Bravo-Abad, and F. J. Garcia-Vidal, “Quantum emitters near a metal nanoparticle: strong coupling and quenching,” Phys. Rev. Lett. 112(25), 253601 (2014). [CrossRef]   [PubMed]  

37. H. Yang, J. Yao, X. Wu, D. Wu, and X. Liu, “Strong plasmon−exciton−plasmon multimode couplings in three-layered Ag−J-Aggregates−Ag nanostructures,” J. Phys. Chem. C 121(45), 25455–25462 (2017). [CrossRef]  

38. F. B. Barho, F. Gonzalez-Posada, M. J. Milla-Rodrigo, M. Bomers, L. Cerutti, and T. Taliercio, “All-semiconductor plasmonic gratings for biosensing applications in the mid-infrared spectral range,” Opt. Express 24(14), 16175–16190 (2016). [CrossRef]   [PubMed]  

39. W. A. Murray, S. Astilean, and W. L. Barnes, “Transition from localized surface plasmon resonance to extended surface plasmon-polariton as metallic nanoparticles merge to form a periodic hole array,” Phys. Rev. B Condens. Matter Mater. Phys. 69(16), 165407 (2004). [CrossRef]  

40. S. Kalusniak, S. Sadofev, and F. Henneberger, “Resonant interaction of molecular vibrations and surface plasmon polaritons: the weak coupling regime,” Phys. Rev. B Condens. Matter Mater. Phys. 90(12), 125423 (2014). [CrossRef]  

41. L. Du, X. Zhang, T. Mei, and X. Yuan, “Localized surface plasmons, surface plasmon polaritons, and their coupling in 2D metallic array for SERS,” Opt. Express 18(3), 1959–1965 (2010). [CrossRef]   [PubMed]  

42. R. Feng, W. Ding, L. Liu, L. Chen, J. Qiu, and G. Chen, “Dual-band infrared perfect absorber based on asymmetric T-shaped plasmonic array,” Opt. Express 22(S2), A335–A343 (2014). [CrossRef]   [PubMed]  

43. P. Bouchon, F. Pardo, B. Portier, L. Ferlazzo, P. Ghenuche, G. Dagher, C. Dupuis, N. Bardou, R. Haïdar, and J. L. Pelouard, “Total funneling of light in high aspect ratio plasmonic nanoresonators,” Appl. Phys. Lett. 98(19), 191109 (2011). [CrossRef]  

44. C. W. Cheng, M. N. Abbas, Z. C. Chang, M. H. Shih, C. M. Wang, M. C. Wu, and Y. C. Chang, “Angle-independent plasmonic infrared band-stop reflective filter based on the Ag/SiO2/Ag T-shaped array,” Opt. Lett. 36(8), 1440–1442 (2011). [CrossRef]   [PubMed]  

45. F. Pardo, P. Bouchon, R. Haïdar, and J. L. Pelouard, “Light funneling mechanism explained by magnetoelectric interference,” Phys. Rev. Lett. 107(9), 093902 (2011). [CrossRef]   [PubMed]  

46. M. Muallem, A. Palatnik, G. D. Nessim, and Y. R. Tischler, “Strong light-matter coupling between a molecular vibrational mode in a PMMA film and a low-loss mid-IR microcavity,” Ann. Phys. 528(3–4), 313–320 (2016). [CrossRef]  

47. N. T. Fofang, T. H. Park, O. Neumann, N. A. Mirin, P. Nordlander, and N. J. Halas, “Plexcitonic nanoparticles: plasmon-exciton coupling in nanoshell-J-Aggregate complexes,” Nano Lett. 8(10), 3481–3487 (2008). [CrossRef]   [PubMed]  

48. S. Savasta, R. Saija, A. Ridolfo, O. Di Stefano, P. Denti, and F. Borghese, “Nanopolaritons: vacuum Rabi splitting with a single quantum dot in the center of a dimer nanoantenna,” ACS Nano 4(11), 6369–6376 (2010). [CrossRef]   [PubMed]  

49. P. Törmä and W. L. Barnes, “Strong coupling between surface plasmon polaritons and emitters: a review,” Rep. Prog. Phys. 78(1), 013901 (2015). [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 (4)

Fig. 1
Fig. 1 (a) Schematic of the cross section of SG-PMMA hybrid structure. (b) Absorption spectra of the SG–PMMA, the silver grating, and the PMMA molecules in the grooves, respectively. Here, the period P = 3000 nm, the element width w = 2000 nm, and the height of the elements h = 775 nm.
Fig. 2
Fig. 2 (a) Contour plot of the absorption spectra of the silver grating as a function of element height h. (b) Contour plot of the absorption spectra of the SG–PMMA as a function of h. The dashed and dotted lines represent variations of the LSP mode and molecular vibrational mode, respectively. (c) Maximal absorption of the SG–PMMA at 5.77 μm as a function of h. Here, the period P and width w are fixed at 3000 nm and 2000 nm, respectively.
Fig. 3
Fig. 3 (a) Absorption spectra of the silver gratings with grating period P of 3000 nm (solid line), 2700 nm (short dashed line), 2400 nm (short dotted line). (b) Electric field distribution in the silver gratings with P = 3000 nm at resonance wavelength of 5.30 μm. (c) Absorption spectra of the silver gratings with grating period P of 2100 nm (solid line), 2090 nm (short dashed line), 2080 nm (short dotted line). (d) Electric field distribution in the silver grating with P = 2100 nm at resonance wavelength of 5.73 μm. Here, the element width w and element height h are fixed at 2000 nm and 700 nm, respectively.
Fig. 4
Fig. 4 (a) Contour plot of the absorption spectra of the silver grating as a function of h. (b) Contour plot of the absorption spectra of the SG–PMMA hybrid structure as a function of h. Here, the period P and width w are fixed at 2100 nm and 2000 nm, respectively. The solid and dash-dot lines show the hybridized HM and LM, respectively. The dashed and dotted lines represent the uncoupled nanocavity mode (NC) of the silver grating and molecular vibrational mode of PMMA, respectively.

Equations (1)

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

( E 1 +i Γ plasmon /2 g g E 2 +i Γ PMMA /2 )( α β )=E( α β ),
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.