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

Numerical investigations of a near-infrared plasmonic refractive index sensor with extremely high figure of merit and low loss based on the hybrid plasmonic waveguide-nanocavity system

Open Access Open Access

Abstract

With vertically slotted hybrid plasmonic waveguides (VSHPWs)-nanocavity system fabricated on the silicon-on-insulator platform, a near-infrared surface plasmon resonances (SPRs)-based refractive index (RI) sensor with extremely high figure of merit FOM = 224.3 and transmission efficiency T = 97.6% is proposed and investigated. Based on the finite element method, effective mode index behaviors together with spectral properties are calculated to analyze and optimize the sensing performance. Within near-infrared region, the wavelength sensitivity (S) and optical resolution (FWHM) can be achieved as S = 1817.5nm/RIU and FWHM = 7.4nm. A mechanism of synergy between propagating SPRs and localized SPRs is also presented for further improving the sensitivity (as high as 2647.5nm/RIU). In addition, the VSHPWs-based RI sensor can be fully realized by CMOS-compatible fabrication technology. In general, the high FOM, S and T achieved by our designed structure may have extensive applications in nanophotonic circuits, environmental monitoring and even pharmaceutical research.

© 2016 Optical Society of America

1. Introduction

Surface plasmon polaritons (SPPs) are surface waves that are bound to and propagate along metal-insulator interfaces, when incident electromagnetic waves are coupled into the collective oscillations of free electrons in a metal [1]. A variety of metal-insulator-metal (MIM)-based plasmonic nanostructures have been designed for guiding SPP modes so far, such as metallic gap (hetero)waveguides, slot waveguides, trench waveguides and V-shaped grooves [2]. Since surface evanescent waves have electric field components polarized perpendicular to the propagation direction, SPP-based plasmonic waveguides can show ultra-compact mode area, strongly enhanced mode field and high sensitivity to the refractive index (RI) variation of surrounding dielectrics [3]. However, the only disadvantage of these traditional plasmonic waveguides (TPWs) is their intrinsic absorption loss caused by metals [4]. Conversely, photonic modes guided in dielectric waveguides present extremely low propagation loss and relatively large mode area. Therefore, incorporating the advantages of both SPP modes and photonic modes, the hybrid plasmonic waveguides (HPWs) based on metal-insulator-semiconductor structures becoming an excellent alternative are extensively investigated [5–9]. Typically, the HPWs are horizontally or vertically slotted slab structures including a metal cladding, an in-between low-index insulator slot and a high-index semiconductor substrate, presenting mode properties dependent on the slot size and substrate dimensions. Metal cladding and high-index semiconductor substrate can squeeze light power into the very thin low-index slot, in which an extremely low-loss and sensitive hybrid mode is established [10]. Physically, this feature can be explained from two aspects. On the one hand, owing to the insulator-semiconductor interface is instead of an insulator-metal one in the TPWs, the hybrid mode has low losses. Thus, the electric field components can spread over the insulator region rather than tightly bounding to the metal surface. On the other hand, the discontinuity of electric field components at high-index-contrast interfaces can generate polarization charges, which will effectively interact with plasmon oscillations excited at the metal-insulator interface [11]. In other words, the hybrid mode is a result of coupling between two evanescent waves with discontinuous transverse electric field components.

In recent years, a great many HPWs-based functional components have been investigated, for example, directional couplers [12], polarization beam splitters [13], nanolasers [14], power splitters [15] and plasmonic sensors [16–24]. Regarding plasmonic RI sensors, by modulating surface plasmon resonances (SPRs) including propagating surface plasmon resonances (PSPRs) and localized surface plasmon resonances (LSPRs), a linear relationship between resonant wavelength of spectral shift and RI of material under detection can be established for realizing RI sensing, which is the most commonly used modulation method [11]. As [19–24] illustrated, many RI sensors are designed by exploiting SPRs-based gratings and periodic nanostructures, like metallic slit (nanoring) arrays acting as perfect absorbers to confine light. However, compared with those complex structures, a combination of HPWs and resonant nanocavities (denoted as waveguide-cavity system for short) is of absolute superiority because of its excellent spectral properties and thus sensing performance. Firstly, featuring best trade-off between propagation loss and mode confinement, HPWs can give rise to extremely confined and enhanced hybrid mode that are very sensitive to the RI change of surrounding medium. Secondly, the high-quality cavity with metallic walls can pick out the required wavelength accurately, showing favorable spectral properties like narrow line-width. Thirdly, this type of RI sensors is more ease of integration. For example, in the case that the incoming light signal is TE-polarized, a polarization converter [25] is only required to integrate ahead with the input waveguide of RI sensor, enabling the TE-polarized signal to convert into the desired TM-polarized one for RI sensing. Therefore, plasmonic RI sensors based on the waveguide-cavity system are essential components with high-performance and easy-to-integrate features, which are deserving of particularly investigated.

In this paper, by sandwiching a rectangular cavity between two vertically slotted hybrid plasmonic waveguides (VSHPWs), a novel near-infrared plasmonic RI sensor with excellent performance (high sensitivity, narrow line-width and good degree of linearity) is proposed and investigated. Compared with other types of waveguide-cavity system-based RI sensors, we introduce a mechanism of synergy between LSPRs and PSPRs that can further improve the sensitivity to as high as 2647.5nm/RIU from 1817.5nm/RIU. With CMOS-compatible technology, the proposed RI sensor can be designed and fully realized on the silicon-on-insulator (SOI) platform.

2. Horizontally slotted hybrid plasmonic waveguide (HSHPW)

2.1 Hybrid structure

Figure 1 shows the cross-sectional view of a HSHPW with fabrication procedure on the right. For such a commonly used multilayer slab structure, as described in [26–28], the fabrication technology is CMOS-compatible and fully realizable. With a high-index Si substrate, different combinations of metals (such as gold, silver, copper and aluminum) and insulators (such as SiO2, SiC and MgF2) can present diverse mode and transmission characteristics. More detailed descriptions about these combinations can be found in [29]. The heights of Si rib, SiO2 slot and metal cladding are hrib = 300nm, hslot = 50nm and hm = 100nm, respectively. The widths of Si rib, SiO2 slot and metal cladding are wrib = 250nm. The dielectric constants of air, Si and SiO2 are εair = 1, εsi = 12.11, and εSiO2 = 2.25. Metal of silver (Ag) is selected due to its smallest imaginary part in the near-infrared region (NIR), which leads to relatively low absorption loss. Here, Ag is characterized by the Drude model ε(ω)=εωp2/[ω(ω+iγ)], where ε = 3.7, ωp = 9.1[eV], and γ = 0.018[eV] [30].

 figure: Fig. 1

Fig. 1 (Left) Cross-sectional view of the HSHPW structure. (Right) CMOS-compatible fabrication technology a-d in terms of the practical procedure.

Download Full Size | PDF

2.2 Mode characteristic

By varying slot size, the HSHPW can support three modes: Si-based photonic mode (mode S), SPP-based plasmonic mode (mode P) and HSHPW-based hybrid mode (mode H). As eigenmodes existing in the Si rib and at the Ag-SiO2 interface, mode S and mode P are featured by nearly lossless propagation and nanosize mode confinement, respectively. As an efficient hybridization of mode P and leaky mode S, mode H exhibits extremely low loss, ultra-compact mode area and high sensitivity to RI change. In addition, field evolution from mode H to mode S is shown in Fig. 2. For hslot < 50nm, mode H presenting more plasmonic-like is mainly confined in the slot. However, for 50nm < hslot <300nm, mode H starts to leak into the Si rib, thus presenting more photonic-like. Finally, for hslot = 500nm, a pure mode S is formed in the Si rib. To show mode S clearly, the field evolution distributions are all solved in a symmetric HSHPW structure being outlined in Fig. 2(a), where we use a SiO2 substrate with width wrib instead of the previous SOI platform. Here, the incident wavelength is 1550nm.

 figure: Fig. 2

Fig. 2 (a)-(f) Field evolution of the SPP modes from hybrid mode H to pure photonic mode S. Values of hslot are marked in the yellow blocks.

Download Full Size | PDF

Next, based on the finite element method, effective mode index (denoted as neff) of the HSHPW is calculated by mode solver. With simulation domain enclosed by perfectly matched layers, free triangular meshing with maximum (minimum) element size of 10nm (1nm) is used for calculation convergence. Since the SiO2 slot plays an important role in controlling the coupling effect, neff-wrib relationships for different hslot are plotted in Fig. 3(a). Thinner slot has larger neff because more power penetrates into metal. Figure 3(b) shows that thicker metal cladding has no impact on neff as long as hm is larger than silver’s skin depth (typically about 20nm in the NIR). In Fig. 3(c), for constant hslot, slot with larger wHPW tends to concentrate more power, resulting in larger neff. In Fig. 3(d), Pnorm-wrib relationships are plotted for indicating power confinement capability of the slot. As defined in [31], surface integrals of the merely slot domain and whole simulation domain are introduced to calculate the power values of Pslot and Ptotal, respectively. Then the power density in slot is defined as Pdensity = Pslot / Aslot, where Aslot is the area of slot domain. Therefore, the normalized power density is calculated by Pnorm = Pdensity / Ptotal. For hslot = 10nm, Pnorm = 156[1/μm2] is larger than Pnorm values obtained in [27,28].

 figure: Fig. 3

Fig. 3 Real part of neff varies as a function of (a) wrib and (b)-(c) hslot, for varying (a) hslot (b) hm and (c) wrib. (d) Pnorm varies as a function of wrib for different hslot.

Download Full Size | PDF

3. Near-infrared plasmonic RI sensor based on VSHPWs-nanocavity system

3.1 Structure design

The cross-sectional and three dimensional views of the proposed RI sensor based on VSHPWs-nanocavity system are shown in Fig. 4(a) and 4(b), respectively. The geometric parameters are set as l1 = l2 = 500nm, wslot = 25nm, wrib = 50nm, wHPW = 100nm, wm = 100nm and hrib = 250nm. In Fig. 4(b), the sensing cavity is sandwiched by input VSHPW1 and output VSHPW2. With the metal gaps, an incident power with Pin and y-polarized electric field can be efficiently coupled into and out of the cavity. Note that, the cavity length lc is calculated by the expression λm=neffs/m derived in [30], together with neff-n0 relationships shown in Fig. 4(d). The hollow sensing cavity has neff = 1.2 and n0 = 1.0, where n0 is the RI of material under detection. Thus, the cavity length (lc = s/2) is calculated as 646nm using a near-infrared resonant wavelength λm = 1550nm. Here, lc is set as 650nm. In Fig. 4(c), for wslot = 25nm, the hrib = 240nm is corresponds to the minimum of imag(neff) and thus a maximum of transmission efficiency.

 figure: Fig. 4

Fig. 4 (a) Cross-sectional and (b) three dimensional views of the proposed RI sensor. (c) Effective mode index behavior of the VSHPW with different wslot for varying hrib. (d) Linear relationship between neff of the cavity and n0 of the material under detection, for several random incident wavelength λ0 (For clarity, small vertical displacements are made and labeled on the right side of each curve).

Download Full Size | PDF

3.2 Spectral property

In our simulation, on the basis of boundary mode analysis, Pout is calculated as an integral of Poynting vector’s x-component along the output boundary of VSHPW2. Pin is set as the default value of 1[W] [30]. Experimentally, the incident power can be injected into VSHPW1 by a nanofiber, while the output power can be detected from VSHPW2 by a spectrometer [32]. As shown in Fig. 5(a), adjusting width d of two metal gaps for desirable coupling effect and resonant peak profile, transmission efficiency T (defined as T = Pout / Pin) of the whole system reduces to 65% from 97.6% as d increases from 5nm to 30nm [33]. For d = 30nm, line-widths of the first-order (mode 1 at λm = 1667nm) and second-order (mode 2 at λm = 841nm) resonant peaks are 7.4nm and 10.3nm, respectively. This can be physically explained that narrower metal gap leads to more power penetration into dielectrics, resulting in lower absorption loss, larger T and wider line-width. However, slightly wider metal gap can support efficiently excited and coupled SPP modes, showing decreased T and narrower line-width. Thus, moderate d contributes to acceptable T and small FWHM (defined in section 3.4). In subsequent investigations, as sensing indicator, mode 1 peak for d = 30nm is selected for its good optical resolution. Next, for several wslot, spectral shifts of mode 1 caused by the RI variation of different material is demonstrated in Fig. 5(b). By following the direction of black solid arrows, there is a one-to-one correspondence between peaks under n0 = 1.0 (n0 = 1.4) tag and wslot values listed above the top arrow. Notice that for smaller wslot the λ0 is getting larger, because constant lc and smaller wslot can result in larger neff and thus λ0.

 figure: Fig. 5

Fig. 5 (a) The dependence of transmission efficiency T on λ0 for different metal gap width d. (b) Spectral shifts caused by RI change as n0 increases from 1.0 to 1.4, for varying wslot (d = 30nm).

Download Full Size | PDF

In addition, the normalized electric field Enorm (defined as Enorm = |E| / |E|max) and power flux density (x-component) distributions of mode1, mode2, mode3 and a non-resonant mode are presented in Fig. 6(a)-6(d), respectively. We can see clearly the PSPRs distributions in cavity and transmission behaviors of the SPP modes from VSHPW1 to VSHPW2.

 figure: Fig. 6

Fig. 6 Normalized Enorm and power flux density (x-component) distributions of the (a) first-order (b) second-order (c) third-order resonant SPP modes and (d) non-resonant SPP modes.

Download Full Size | PDF

3.3 Sensing performance

The sensing performance can be characterized by three performance indices. First, the wavelength sensitivity S of a RI sensor is defined as Δλm/Δn0 (nm/RIU), representing the amount of resonant spectral shifts per refractive index unit. Second, the line-width of a resonant peak is indicated by FWHM (full width at half maximum), showing how accurate is the λm can be identified. Third, however, figure of merit FOM that defined as the ratio of S to FWHM is the most comprehensive one to evaluate the sensing performance. The larger FOM is, the better trade-off between sensitivity and resolution is achieved. Concluded from the spectral shifts shown in Fig. 5(b), the detailed sensing behaviors are listed in Table1. With wslot increasing from 5nm to 45nm, S (FWHM) decreases from 1817.5nm/RIU (11.8nm) to 1587.5nm/RIU (7.4nm). As a result, the proposed RI sensor with FOM as high as 224.3 can be achieved for wslot = 25nm. The last but equally important is, except for S, FWHM and FOM, the good degree of linearity of λm-n0 curves shown in Fig. 7(a). Beyond the NIR, a simple prediction of λm for different lc can be realized by the linear relationship of λm-lc curves presented in Fig. 7(b).

Tables Icon

Table 1. Sensing Performance of the Proposed Plasmonic RI Sensor

 figure: Fig. 7

Fig. 7 (a) Linear relationship between λm and n0 for different wslot. (b) Linear relationship between λm of mode 1 and cavity length lc (n0 = 1.0).

Download Full Size | PDF

Considering practical application, the perfect symmetry of the structure may get broken for the precision limitation during fabrication process. Thus the fabrication tolerance and alignment tolerance are calculated. First, concluded from Table 1, ± 5nm variations in slot width (wslot = 25nm) can lead to average fluctuations of ± 1.45% for S, ± 8.05% for FWHM and ± 8.0% for FOM. It is acceptable that, the relatively large fabrication tolerance still maintains a FOM larger than 200. Second, numerical simulation results show that a y-direction deviation of ± 20nm between VSHPWs and cavity is acceptable, which leads to merely <1nm variation in S. In general, both fabrication tolerance and alignment tolerance are relaxed for the structure.

3.4 Further improvement

For better sensing performance, a cavity embedded with metal nanoparticles (NPs) is proposed. As shown in Fig. 8(a), within the cavity with lc = 450nm, three Ag NPs with radius r = 45nm are placed in the cavity with equal spacing, keeping other geometric parameters unchanged. Comparing Enorm distributions shown in Fig. 6(a)-6(c) and Fig. 8(b)-8(d), as can be expected for a strongly enhanced coupling effect in the dielectric cavity, we find that most power of resonant SPP modes is concentrated in the gap region between Ag slab and NPs. When LSPRs are excited by PSPRs in the cavity, standing wave distributions are coupled with individual dipole-like LSPRs. The LSPRs-PSPRs synergy can greatly enhances coupling efficiency of SPP modes and thus the sensitivity. As the transmission spectrum presented in Fig. 8(e), mode 1 exhibits a sensitivity of 2230nm/RIU and a FWHM of 17nm. For lc = 650nm, the sensitivity is increased from 1665nm/RIU to 2647.5nm/RIU for r increasing from 5nm to 45nm. Note that, lc = 650nm and r = 45nm are maximal structure size due to λm = 2240nm of mode 1. Furthermore, a good degree of linearity of λm-n0 curve is also presented in Fig. 8(f). Within the NIR, plasmonic RI sensors based on the TPWs-cavity system, can realize a S of 1496nm/RIU and a FOM of 124.6 by a fillet cavity in [30], a S of 1476.3nm/RIU by a ring cavity in [33], and a S of 1563nm/RIU and a FOM of 38.6 by a hexagonal cavity in [34]. In addition, values of S and FOM achieved by other types of nano-structured plasmonic RI sensors are concluded in [35].

 figure: Fig. 8

Fig. 8 (a) Perspective view of the proposed RI sensor with a modified cavity. Top view of the Enorm distributions of the (b) first-order (c) second-order (d) third-order resonant SPP modes in cavity. (e) Spectral shifts in the transmission spectrum for lc = 450nm and r = 45nm. (f) Linear relationship between λm of mode 1 and n0 of material under detection.

Download Full Size | PDF

According to fabrication technologies and manipulation approaches presented in [36], we present a basically feasible manufacturing process as shown in Fig. 9. Firstly, two Si ribs with 50nm in width and 500nm in length are etched on the top of SOI platform by electron beam lithography (EBL). Secondly, SiO2 ribs adjacent to Si ribs are fabricated by sputter coating or plasmon-enhanced chemical vapor deposition. Then EBL can be used to tailor the width and length of SiO2 ribs. Thirdly, after placing a protection cover on the position of sensing cavity, Ag cladding can be formed by magnetron sputtering method. Finally, three prepared Ag NPs are placed one by one using contact nanomanipulation techniques, such as multi-probe scanning electron microscopy (SEM) and scanning atomic force microscopy. For example, with ultra-high resolution and relatively low energy, SEM manipulation enables good accuracy in controlling NPs within the proposed sensing cavity with moderate size. It is worth noting that repulsive interactions resulted from interparticle and metallic wall-particle gaps can provide a good force balance, which keeps the Ag NPs in a relatively steady state. As described in [37], recently the most popular way of preparing noble metal NPs is the chemical colloidal synthesis. In addition to the large fabrication tolerance analyzed above, the Ag NPs with diameters of 90nm exhibit desirable preparation feasibility and operating controllability. In general, lithography technique with ultra-high resolution and excellent chemical colloidal synthesis method contribute greatly to the manufacture and preparation of plasmonic nanostructures [37].

 figure: Fig. 9

Fig. 9 (Left) Schematic views for the fabrication process. (Right) From a to e, the detailed manufacture and preparation approaches are listed according to actual operation.

Download Full Size | PDF

By employing PSPRs or LSPRs resulted from metallic nanostructures, the proposed NPs-embedded plasmonic sensor allows tremendous applications in environmental monitoring [38] and disease detection [39]. For example, when monitoring a liquid sample containing heavy metal ions with a certain concentration, the RI sensing can be fulfilled on a micro/nanofluidic platform with spectrum response detected by SPRs spectrometer [38].

4. Conclusion

In this paper, on the premise of CMOS-compatible fabrication feasibility, we propose a near-infrared plasmonic RI sensor based on waveguide-cavity system, which consists of two Ag-SiO2-Si VSHPWs coupled with a rectangular cavity. By adjusting the metal gap width between VSHPWs and cavity, an extremely confined and enhanced hybrid mode can be efficiently coupled in and out of the cavity. As a result, the sensing function is realized with high-performance including high sensitivity of 1660nm/RIU, narrow line-width of 7.4nm, large FOM of 224.3 and good degree of linearity. To further improve sensitivity, a mechanism of PSPRs-LSPRs synergy is presented, which can give rise to a sensitivity as high as 2647.5nm/RIU. However, a moderate FOM of 156 is obtained because of the relatively wide line-width of 17nm. Notice that, an increasing sensitivity is always accompanying with a decreasing resolution and vice versa. Therefore, achieving good balance between sensitivity and resolution is still a significant work. In conclusion, the high sensitivity, good resolution and efficient transmission achieved by our design may pave the way for further investigations in nanophotonic circuits, environmental monitoring and even pharmaceutical research.

Funding

Ministry of Science and technology of China (2016YFA0301300); National Natural Science Foundation of China (NSFC) (61275201, 61372037); Fund of State Key Laboratory of Information Photonics and Optical Communications (Beijing University of Posts and Telecommunications), P. R. China (IPOC2016ZZ03); Fundamental Research Funds for the Central Universities (2016RC24); Beijing Excellent Ph.D. Thesis Guidance Foundation (20131001301); the BUPT Excellent PhD. Students Foundation (CX2016204).

References and links

1. L. Novotny and B. Hecht, Principles of Nano-Optics (Cambridge University, 2012).

2. S. I. Bozhevolnyi, Plasmonic Nanoguides and Circuits (CRC, 2008).

3. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007).

4. J. A. Schuller, E. S. Barnard, W. Cai, Y. C. Jun, J. S. White, and M. L. Brongersma, “Plasmonics for extreme light concentration and manipulation,” Nat. Mater. 9(3), 193–204 (2010). [CrossRef]   [PubMed]  

5. M. Wu, Z. Han, and V. Van, “Conductor-gap-silicon plasmonic waveguides and passive components at subwavelength scale,” Opt. Express 18(11), 11728–11736 (2010). [CrossRef]   [PubMed]  

6. J. Zhang, L. Cai, W. Bai, Y. Xu, and G. Song, “Hybrid plasmonic waveguide with gain medium for lossless propagation with nanoscale confinement,” Opt. Lett. 36(12), 2312–2314 (2011). [CrossRef]   [PubMed]  

7. L. Chen, T. Zhang, X. Li, and W. Huang, “Novel hybrid plasmonic waveguide consisting of two identical dielectric nanowires symmetrically placed on each side of a thin metal film,” Opt. Express 20(18), 20535–20544 (2012). [CrossRef]   [PubMed]  

8. Z. Zhang and J. Wang, “Long-range hybrid wedge plasmonic waveguide,” Sci. Rep. 4, 6870 (2014). [CrossRef]   [PubMed]  

9. L. Ding, J. Qin, K. Xu, and L. Wang, “Long range hybrid tube-wedge plasmonic waveguide with extreme light confinement and good fabrication error tolerance,” Opt. Express 24(4), 3432–3440 (2016). [CrossRef]   [PubMed]  

10. D. Dai and S. He, “A silicon-based hybrid plasmonic waveguide with a metal cap for a nano-scale light confinement,” Opt. Express 17(19), 16646–16653 (2009). [CrossRef]   [PubMed]  

11. E. P. Li and H. S. Chu, Plasmonic Nanoeletronics and Sensing (Cambridge University, 2014).

12. C. H. Du and Y. P. Chiou, “Vertical directional couplers with ultra-short coupling length based on hybrid plasmonic waveguides,” J. Lightwave Technol. 32(11), 2065–2071 (2014). [CrossRef]  

13. B. Sun, Y. Wang, Y. Liu, S. Liu, C. Liao, and M. Y. Chen, “Compact device employed a hybrid plasmonic waveguide for polarization-selective splitting,” Opt. Commun. 334, 240–246 (2015). [CrossRef]  

14. H. Lu, Y. Liu, Z. Yu, C. Ye, and J. Wang, “Hybrid plasmonic waveguides for low-threshold nanolaser applications,” Chin. Opt. Lett. 12(11), 103–106 (2014).

15. J. Wang, X. Guan, Y. He, Y. Shi, Z. Wang, S. He, P. Holmström, L. Wosinski, L. Thylen, and D. Dai, “Sub-μm2 power splitters by using silicon hybrid plasmonic waveguides,” Opt. Express 19(2), 838–847 (2011). [CrossRef]   [PubMed]  

16. B. You, J. Y. Lu, T. A. Liu, and J. L. Peng, “Hybrid terahertz plasmonic waveguide for sensing applications,” Opt. Express 21(18), 21087–21096 (2013). [CrossRef]   [PubMed]  

17. X. Sun, D. Dai, L. Thylén, and L. Wosinski, “High-sensitivity liquid refractive-index sensor based on a Mach-Zehnder interferometer with a double-slot hybrid plasmonic waveguide,” Opt. Express 23(20), 25688–25699 (2015). [CrossRef]   [PubMed]  

18. X. Sun, D. Dai, L. Thylén, and L. Wosinski, “Double-slot hybrid plasmonic ring resonator used for optical sensors and modulators,” Photonics 2(4), 1116–1130 (2015). [CrossRef]  

19. S. Luo, J. Zhao, D. Zuo, and X. Wang, “Perfect narrow band absorber for sensing applications,” Opt. Express 24(9), 9288–9294 (2016). [CrossRef]   [PubMed]  

20. X. Lu, L. Zhang, and T. Zhang, “Nanoslit-microcavity-based narrow band absorber for sensing applications,” Opt. Express 23(16), 20715–20720 (2015). [CrossRef]   [PubMed]  

21. X. Lu, R. Wan, and T. Zhang, “Metal-dielectric-metal based narrow band absorber for sensing applications,” Opt. Express 23(23), 29842–29847 (2015). [CrossRef]   [PubMed]  

22. Q. Liu, J. S. Kee, and M. K. Park, “A refractive index sensor design based on grating-assisted coupling between a strip waveguide and a slot waveguide,” Opt. Express 21(5), 5897–5909 (2013). [CrossRef]   [PubMed]  

23. Y. Binfeng, H. Guohua, Z. Ruohu, and C. Yiping, “Design of a compact and high sensitive refractive index sensor base on metal-insulator-metal plasmonic Bragg grating,” Opt. Express 22(23), 28662–28670 (2014). [CrossRef]   [PubMed]  

24. X. Wang and C. K. Madsen, “Highly sensitive compact refractive index sensor based on phase-shifted sidewall Bragg gratings in slot waveguide,” Appl. Opt. 53(1), 96–103 (2014). [CrossRef]   [PubMed]  

25. Y. Chang and T. Yu, “Photonic-quasi-TE-to-hybrid-plasmonic-TM polarization mode converter,” J. Lightwave Technol. 33(20), 4261–4267 (2015). [CrossRef]  

26. M. S. Kwon, “Metal-insulator-silicon-insulator-metal waveguides compatible with standard CMOS technology,” Opt. Express 19(9), 8379–8393 (2011). [CrossRef]   [PubMed]  

27. D. Dai, Y. Shi, S. He, L. Wosinski, and L. Thylen, “Gain enhancement in a hybrid plasmonic nano-waveguide with a low-index or high-index gain medium,” Opt. Express 19(14), 12925–12936 (2011). [CrossRef]   [PubMed]  

28. D. Dai and S. He, “Low-loss hybrid plasmonic waveguide with double low-index nano-slots,” Opt. Express 18(17), 17958–17966 (2010). [CrossRef]   [PubMed]  

29. M. T. Noghani and M. H. V. Samiei, “Analysis and optimum design of hybrid plasmonic slab waveguides,” Plasmonics 8(2), 1155–1168 (2013). [CrossRef]  

30. L. Chen, Y. Liu, Z. Yu, D. Wu, R. Ma, Y. Zhang, and H. Ye, “Numerical analysis of a near-infrared plasmonic refractive index sensor with high figure of merit based on a fillet cavity,” Opt. Express 24(9), 9975–9983 (2016). [CrossRef]   [PubMed]  

31. V. R. Almeida, Q. Xu, C. A. Barrios, and M. Lipson, “Guiding and confining light in void nanostructure,” Opt. Lett. 29(11), 1209–1211 (2004). [CrossRef]   [PubMed]  

32. T. Wu, Y. Liu, Z. Yu, H. Ye, Y. Peng, C. Shu, C. Yang, W. Zhang, and H. He, “A nanometeric temperature sensor based on plasmonic waveguide with an ethanol-sealed rectangular cavity,” Opt. Commun. 339, 1–6 (2015). [CrossRef]  

33. T. Wu, Y. Liu, Z. Yu, Y. Peng, C. Shu, and H. Ye, “The sensing characteristics of plasmonic waveguide with a ring resonator,” Opt. Express 22(7), 7669–7677 (2014). [CrossRef]   [PubMed]  

34. Y. Xie, Y. Huang, W. Zhao, W. Xu, and C. He, “A novel plasmonic sensor based on metal-insulator-metal waveguide with side-coupled hexagonal cavity,” IEEE Photonics J. 7(2), 1–12 (2015). [CrossRef]  

35. S. Roh, T. Chung, and B. Lee, “Overview of the characteristics of micro- and nano-structured surface plasmon resonance sensors,” Sensors (Basel) 11(2), 1565–1588 (2011). [CrossRef]   [PubMed]  

36. S. Gwo, H. Y. Chen, M. H. Lin, L. Sun, and X. Li, “Nanomanipulation and controlled self-assembly of metal nanoparticles and nanocrystals for plasmonics,” Chem. Soc. Rev. 10, 450 (2016). [CrossRef]   [PubMed]  

37. P. Yang, J. Zheng, Y. Xu, Q. Zhang, and L. Jiang, “Colloidal synthesis and applications of plasmonic metal nanoparticles,” Adv. Mater. 201, 1739 (2016). [PubMed]  

38. I. Choi, “Recent advances in nanoplasmonic sensors for environmental detection and monitoring,” J. Nanosci. Nanotechnol. 16(5), 4274–4283 (2016). [CrossRef]   [PubMed]  

39. S. Barizuddin, S. Bok, and S. Gangopadhyay, “Plasmonic sensors for disease detection – a review,” J. Nanomed. Nanotechnol. 7(3), 1000373 (2016).

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 (9)

Fig. 1
Fig. 1 (Left) Cross-sectional view of the HSHPW structure. (Right) CMOS-compatible fabrication technology a-d in terms of the practical procedure.
Fig. 2
Fig. 2 (a)-(f) Field evolution of the SPP modes from hybrid mode H to pure photonic mode S. Values of hslot are marked in the yellow blocks.
Fig. 3
Fig. 3 Real part of neff varies as a function of (a) wrib and (b)-(c) hslot, for varying (a) hslot (b) hm and (c) wrib. (d) Pnorm varies as a function of wrib for different hslot.
Fig. 4
Fig. 4 (a) Cross-sectional and (b) three dimensional views of the proposed RI sensor. (c) Effective mode index behavior of the VSHPW with different wslot for varying hrib. (d) Linear relationship between neff of the cavity and n0 of the material under detection, for several random incident wavelength λ0 (For clarity, small vertical displacements are made and labeled on the right side of each curve).
Fig. 5
Fig. 5 (a) The dependence of transmission efficiency T on λ0 for different metal gap width d. (b) Spectral shifts caused by RI change as n0 increases from 1.0 to 1.4, for varying wslot (d = 30nm).
Fig. 6
Fig. 6 Normalized Enorm and power flux density (x-component) distributions of the (a) first-order (b) second-order (c) third-order resonant SPP modes and (d) non-resonant SPP modes.
Fig. 7
Fig. 7 (a) Linear relationship between λm and n0 for different wslot. (b) Linear relationship between λm of mode 1 and cavity length lc (n0 = 1.0).
Fig. 8
Fig. 8 (a) Perspective view of the proposed RI sensor with a modified cavity. Top view of the Enorm distributions of the (b) first-order (c) second-order (d) third-order resonant SPP modes in cavity. (e) Spectral shifts in the transmission spectrum for lc = 450nm and r = 45nm. (f) Linear relationship between λm of mode 1 and n0 of material under detection.
Fig. 9
Fig. 9 (Left) Schematic views for the fabrication process. (Right) From a to e, the detailed manufacture and preparation approaches are listed according to actual operation.

Tables (1)

Tables Icon

Table 1 Sensing Performance of the Proposed Plasmonic RI Sensor

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.