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

Optical fiber sensor based on Bloch surface wave in photonic crystals

Open Access Open Access

Abstract

A new optical fiber sensor based on Bloch surface wave was theoretically proposed. An omnidirectional one-dimensional photonic crystal was designed as the multilayer coated on the outer surface of the optical fiber. Taking advantages of the omnidirectional reflection band, there is only surface mode resonance in the transmission spectrum, while guided mode resonance is avoided. The performance of the designed fiber sensor was analyzed theoretically with a ray transmission model. The presented sensor has comparable sensitivity but much higher figure of merit than other fiber sensors. The resolution can reach about 10−6 RIU or even higher.

© 2016 Optical Society of America

1. Introduction

Surface electromagnetic waves have been widely applied during the past few decades, such as surface plasmon polariton (SPP) [1–3]. As the fields of these surface modes are highly localized at the interface, they are very sensitive to the properties of the surface, which allows us to develop efficient optical sensors with high performance [4–6]. Recently, another surface-like wave, the Bloch surface wave (BSW) in photonic crystals (PC) [7–9], has attracted much attention, with many investigations having been conducted, both theoretically [10–14] and experimentally [15–21]. Subsequently, new sensors based on BSW were proposed [22–29]. A direct comparison of BSW and SPP sensors is given in [30].

Traditional evanescent field sensors were generally built on prism coupling configurations, such as the Otto or Kretschmann configurations. These structures are very complicated with large volumes, causing experimental inconvenience in practical applications. Naturally, fiber sensors (FS) based on evanescent fields came into being, which overcame the discrepancies of prism-coupling configurations [31–33].

The BSW in a fiber-like cylindrical geometry has also been experimentally observed [34], however, the potential applications of BSW in FS have not attracted much attention. In this work we present a new FS based on periodic multi-layer structure, i.e. one-dimensional (1D) PC, which works by the excitation of BSW in 1DPC. As the input lights transmitted in the optical fiber sensor have different incident angles, an omnidirectional 1DPC is designed to realize omnidirectional reflection. A cap layer is adopted to modify the surface of 1DPC and to control the excitation of BSW. The influence of the thickness of the cap layer is also investigated. The resonant wavelength (RW) of BSW in FS is designed to be around 1550 nm by appropriate structure parameter settings of the 1DPC. By using a ray transmission model, we evaluate the performance of the designed FS, and find that its sensitivity is comparable to other fiber sensors, but its figure of merit (FOM) is much higher. The resolution of the sensor can reach 10−6 RI Unit (RIU) or even higher with the intensity interrogation method.

2. Omnidirectional 1DPC design

Omnidirectional PC refers to PC that have an overlapping band gap for different incident angles, i.e. omnidirectional reflection band (ORB), which can be applied as perfect mirrors that totally reflect light with any polarization and orientation, within a specific frequency range. Compared to other omnidirectional reflectors such as metallic mirrors, the omnidirectional PC has a wider frequency range of total reflection and lower loss. However, it used to be believed that PC with ORB could only be realized in two- or three-dimensional photonic structures, while the fabrication of 2D or 3DPC is technically inconvenient. Fortunately, more detailed analysis shows that ORB certainly can be achieved in 1DPC [35], which is similar to the periodical optical film system. It is therefore possible for us to design an omnidirectional 1DPC by periodically layered structures.

As shown in Fig. 1, the 1DPC consists of an array of two dielectric layers, with RI n1 andn2, respectively. The thickness of the two layers is denoted byd1 andd2, whilen0 represents the RI of the coupling prism. Let β=n0k0sinθ0 represent the parallel wave vector, i.e. the propagation constant, where θ0 and k0 refer to the incident angle in the prism and the wave vector in vacuum, respectively. The propagation equation of 1DPC is given as

cospd=cosκ1d1cosκ2d212(η1η2+η2η1)sinκ1d1sinκ2d2.
Here κi=ni2k02β2refers to the perpendicular wave vector in the related medium. ηi represents corresponding effective optical admittance, which is different for s-polarized light (SPL) and p-polarized light (PPL). The subscript i=0,1,2,m denotes prism, layer 1, layer 2 and sensed medium, respectively. The symbol p stands for the Bloch wave vector, which satisfies |cospd|<1 in the allowed energy band. The band gap is determined by

 figure: Fig. 1

Fig. 1 Schematic of 1DPC structure for exciting BSW.

Download Full Size | PDF

|cosκ1d1cosκ2d212(η1η2+η2η1)sinκ1d1sinκ2d2|>1.

As is believed in optical film design, the quarter-wave film system produces the largest reflection band [36]. In order to obtain a wide band gap, we adopted this configuration as our template, i.e.n1d1=n2d2=L0/4, where L0 represents the center wavelength.

To achieve ORB, there are generally two key conditions that must be satisfied. The first is a sufficiently large RI contrast of the two basical layers. The second is a suitable coupling material. In Fig. 1, this refers to the prism.

We finally setn1=4.300, n2=2.460, such as Ge and ZnSe at near-infrared wavelength. The center wavelength is determined as L0=1600nm. Fused silica is used as our prism, n0=1.445. The corresponding band structure is presented in Fig. 2.

 figure: Fig. 2

Fig. 2 Band structure of 1DPC. The left side represents PPL, while the right side represents SPL. (a) Band diagram presented in ω-β relationship. The slanted black line stands for the light line in the prism with refractive index n0 = 1.445. The red line stands for the critical condition for producing ORB. The blue line stands for the Brewster line, which crosses the intersection point of the boundaries of band gap of PPL. The gray areas refer to the energy band. (b) While n0 = 1.445, the band gap varies with the angle of incidence. The gray areas refer to band gaps, and the black area denotes ORB. (c) Same as above, except for n0 = 1.837, which is the critical condition. The range of ORB reduces to zero in this circumstance. (d) Same as above, except for n0 = 2.135, which is the case of the Brewster line. Band gap of PPL vanishes as the incident angle increases to 90°.

Download Full Size | PDF

As can be seen in Fig. 2(a), there is an intersection point at the band boundary of PPL, which stands for the case of Brewster angle. The existence of the intersection implies that the band gap will close at this point, as the blue line indicates. The critical position of the light line to achieve ORB is determined by the intersection point of the boundary of the band gap of PPL and the tangent line of the bottom of higher order energy band, as the red line indicates. This critical line is generally above the Brewster line. To realize omnidirectional 1DPC, the RI of the prism must be smaller than this critical value.

Our calculations give the RI of prism for critical line asn0=1.837, while that of Brewster line isn0=2.135. When the RI of the prism is smaller than the critical value, i.e. n0<1.837, ORB occurs. As shown in Fig. 2(b), n0=1.445<1.837, the light line is above the critical line. The ORB is marked in the black rectangular area, ranging from 1361.1 to 1606.4 nm. Generally, the lower the RI of the prism, the larger the ORB will be. In the critical case, as Fig. 2(c) shows, as the incidence angle increases to 90°, though band gap of PPL still exists, the ORB reduces to zero. The upper boundary of the band gap of PPL in glancing incidence is exactly 1361.1 nm, which is the same as the lower boundary of the band gap at normal incidence, as the black tangent line indicates. As the RI of the prism exceeds the critical line and reaches the Brewster line, the band gap of PPL vanishes at glancing incidence, which is the case of Fig. 2(d). It is not difficult to infer that, once the RI continuously grows, the band gap of PPL will close at a specific angle.

3. BSW excitation

The BSW of 1DPC is very sensitive to the environment of the interface. Therefore, a cap layer is generally adopted to modify the surface of the PC, as shown in Fig. 1. And the resonant condition of BSW can be modulated by the RI and thickness of the cap layer [15]. Assume that the sensed medium is water, with RI nm=1.330. The cap layer is adopted as ZnS, with RInc=2.280+0.0002i and thickness dc=615nm. The period of the bilayer is initially set as six. As presented above, when n0=1.445, the range of ORB is 1361.1–1606.4 nm, indicating that we can use this wavelength range to investigate BSW at any incident angle. The critical angle in this case is θc=66.987.

In this calculation, the transfer matrix method (TMM) [37] is applied, which gives the transfer matrix of a single layer as

Mi=[cosκidiisinκidiηiiηisinκidicosκidi].

The total transfer matrix of the system is a product of the transfer matrices of all layers, i.e.

M=Mi=(M1M2)NMC.

Then, the reflection coefficient can be obtained by

r=(M11+M21ηm)η0(M21+M22ηm)(M11+M21ηm)η0+(M21+M22ηm).

Figure 3 gives the spectrum of the designed 1DPC when the incident angle is 85°, which is larger than the critical angle. As we can see, the resonant wavelengths are 1379.0 nm for SPL and 1593.1 nm for PPL, which are both located in the band gap and represent the excitation of BSW of SPL and PPL, respectively.

 figure: Fig. 3

Fig. 3 Reflection spectrum of designed 1DPC. Blue and red lines represent the reflectance of SPL and PPL, respectively.

Download Full Size | PDF

As mentioned above, the resonant condition relies heavily on the cap layer. Therefore, in Fig. 4, we present the dependence of RW on the thickness of the cap layer. As can be seen in the figures, as the thickness of the cap layer changes from 605 nm to 625 nm, RW of SPL increases from 1361.2 nm to 1396.4 nm, while RW of PPL increases from 1581.2 nm to 1604.3 nm, nearly linearly. This indicates that we can adjust RW by controlling the thickness of the cap layer.

 figure: Fig. 4

Fig. 4 Thickness of cap layer affecting the resonant condition of BSW. (a) Reflectance of SPL varying with thickness of cap layer. (b) Reflectance of PPL varying with thickness of cap layer.

Download Full Size | PDF

Many studies have successfully excited BSW in various ways. However, many of these spectra contain not only the surface mode, but also some guided modes located outside the band gap. These guided modes also manifest themselves as dips in the reflection spectrum, resulting in multi-dip curves. In order to show the characteristics of BSW clearly, it is better to avoid guided mode resonance. In our results, there is only one dip, which exactly represents BSW. The difference comes from the band structure of 1DPC. In our calculations, the 1DPC is designed with an ORB, and the wavelength we used is always within ORB for any incidence angle. As guided modes cannot occur in the band gap, only surface modes can be excited in this circumstance.

4. FS design and analysis

In the designed omnidirectional 1DPC, BSW can be excited for both SPL and PPL with high detectability. Now we can apply it in FS. Figure 5 shows the structure of the designed BSW FS. The multilayer is coated on the outer surface of the core of a multi-mode optical fiber, which could be experimentally fabricated via the atomic layer deposition technique [34, 38]. As the diameter of the fiber that we designed is much larger than the wavelength of input light, we adopted a ray transmission model in the following calculations [39, 40].

 figure: Fig. 5

Fig. 5 Schematic and ray transmission model of designed fiber sensor.

Download Full Size | PDF

Assume that the input light has a standard Gaussian profile, with a power distribution expressed as

PIn(φ)=P0exp(φ2φ02),
where φ is the incident angle of a specific plane wave component, and φ0 is the divergence angle. The power distribution of output light is given as
POut=PIn(φ)RN(φ),
where R(φ) stands for the single reflectance of the light with incident angle φ. Here we only take the meridian rays into consideration. N refers to the number of reflections, which is determined by the incident angle and the size of the fiber. It can be calculated by N=Lcotθ/D, where L and D denote the length and diameter of the fiber, respectively. represents the practical incidence angle on the outer surface of the fiber core, with the relationship sinφ=n0cosθ. Note that, in order to show simultaneous excitation of BSW of SPL and PPL, this calculation is dependent and equal for SPL and PPL. Then, the total transmittance of the FS can be expressed as
T=0π2POut(φ)dφ0π2PIn(φ)dφ.
The divergence angle is assumed as φ0=3.

Figure 6 gives the transmission spectrum of the fiber sensor. As SPL and PPL are equally mixed in the fiber, the two BSWs can be both excited while incident with broadband light. Obviously, RW is located at about 1375.7 nm and 1587.4 nm, indicating the resonance of BSW of SPL and PPL, respectively. As can be seen, the two dips are easily detectable for full width at half maximum (FWHM) less than 3 nm.

 figure: Fig. 6

Fig. 6 Spectrum of designed fiber sensor.

Download Full Size | PDF

In general, sensing with only one resonant dip, either SPL or PPL, is adequate. In order to make use of fiber communication wavelengths near 1550 nm where there are many high-performance detectors and laser sources, the RWs can be moved to 1550 nm by appropriately setting the thickness of the cap layer. According to the calculation, the thicknesses of exciting BSW of SPL and PPL at 1550 nm are about 717 nm and 585 nm, respectively. For the sake of convenience, in the following calculation, we assume that the input light is pure SPL or PPL when referring to the resonance of SPL or PPL, respectively. The experimental realization of such SPL or PPL in a fiber-like cylindrical geometry can be achieved via the radial polarization convertor [41], or some optical fiber based techniques [42, 43].

It is necessary to analyze the detectability of these two dips at 1550 nm. As the divergence angle can be experimentally tunable via collimation, expansion or focusing, we demonstrated the influence of divergence angle of incident light in Fig. 7(a). From the circles we can see that, as divergence angle increases from 1° to 5°, FWHM generally increases, from 0.5 nm to 2.9 nm for SPL, and 0.6 nm to 3.4 nm for PPL, which is the result of energy diverging. Meanwhile, the detectability is also related to the depth of the dip, i.e. the minimum transmittance (MT). The triangles indicate that the range of MT is 73.18–87.87% for SPL, and 20.43–34.55% for PPL. Obviously, though FWHM of PPL is a little larger than SPL, it has a much smaller MT when RW is shifted to 1550 nm.

 figure: Fig. 7

Fig. 7 Detectibility of resonant dips for SPL and PPL. (a) Influence of divergence angle on FWHM and minimunm transmittance, with period number set as six. (b) Influence of number of periods on FWHM and minimum transmittance, with divergance angle equalling 3°. Blue and red lines stand for SPL and PPL, while circles and triangles denote FWHM and minimum transmittance, respectively.

Download Full Size | PDF

Furthermore, we investigated the effect of the period of the bilayer. As Fig. 7(b) shows, FWHM decreases as the number of bilayers increases. This is because FWHM represents the loss characteristic of resonance. More layers indicate less energy loss, leading to a smaller width of resonant dip. However, MT increases if more layers are added. Considering the fabrication difficulty, six periods will be a good tradeoff. Therefore, in the following discussion, we set the divergence angle as 3°, and period number as six.

Another important performance of sensor is sensitivity. In order to study the sensitivity of the fiber sensor, we calculated the RW under different RIs of the sensed medium. The results are presented in Fig. 8. As we can see, while the RI of the sensed medium varies from 1.320 to 1.340, RW changes from 1547.3 nm to 1553.0 nm for SPL, and from 1544.8 nm to 1555.5 nm for PPL, roughly linearly. Defining the sensitivity as S=δλ/δn, we obtained the sensitivity as about 285 nm/RIU for SPL, and 535 nm/RIU for PPL, which is comparable to other fiber sensors. However, FWHM of BSW fiber sensors is significantly smaller. In our calculation, it is 1.5 nm for SPL and 1.7 nm for PPL. Therefore, the figure of merit, defined as FOM = S/FWHM, is extremely large for our sensor, reaching about 190.0 RIU−1 for SPL and 314.7 RIU−1 for PPL, which is much better than for other fiber sensors [44, 45]. One thing to note is the sensitivity we discussed above is the bulk sensitivity instead of the surface sensitivity which is also important for some biosensing applications [28].

 figure: Fig. 8

Fig. 8 Dependence of RW on RI of sensed medium for both SPL and PPL. The two inset graphs indicate the transmission spectrum under the corresponding sensed medium.

Download Full Size | PDF

Besides wavelength interrogation, the sensor can also work via intensity interrogation under a proper laser source for its high FOM. As mentioned above, PPL in this case has a smaller MT than SPL, which is more suitable for intensity interrogation. Therefore, we set the thickness of the cap layer as 585 nm in order to make use of BSW of PPL. The predicted detection spectrum is presented in Fig. 9, which is set up under a wavelength of 1550 nm. As can be seen, the transmittance rapidly varies as the RI of sensed medium changes, indicating that the transmittance is very sensitive to the sensed medium. For example, while RI increases by 0.001, from 1.330 to 1.331, the transmittance significantly increases from 21.51% to 64.72%, i.e. about 40%. Assuming that the detection limit of transmittance at 1550 nm is 0.1%, the average resolution of our sensor is approximately 2.5 × 10−6 RIU. In fact, in this wavelength range, many detectors have much lower detection limits, so the practical resolution could be even higher.

 figure: Fig. 9

Fig. 9 Predicted detection spectrum under transmittance-RI relationship. The inset graph shows the wavelength spectrum at corresponding points.

Download Full Size | PDF

5. Conclusions

In conclusion, an optical fiber sensor based on BSW has been presented. The sensor is coated with an omnidirectional 1DPC composed of Ge and ZnSe periodical layers and a ZnS cap layer for the excitation of BSW. The important role that the cap layer plays in controlling the RW is analyzed. Performances including sensitivity, FOM and resolution of the designed sensor are evaluated theoretically with a ray transmission model. Its resolution is estimated to be 10−6 RIU. The material selection of the 1DPC and cap layer could vary as needed, but the theoretical analysis method still applies. The results of this work could contribute to the application of BSW in fiber sensors.

Funding

Shanghai Natural Science Foundation (Grant No. 15ZR1404100).

References and links

1. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424(6950), 824–830 (2003). [CrossRef]   [PubMed]  

2. A. V. Zayats, I. I. Smolyaninov, and A. A. Maradudin, “Nano-optics of surface plasmon polaritons,” Phys. Rep. 408(3-4), 131–314 (2005). [CrossRef]  

3. P. Berini, “Long-range surface plasmon polaritons,” Adv. Opt. Photonics 1(3), 484 (2009). [CrossRef]  

4. J. Homola, S. S. Yee, and G. Gauglitz, “Surface plasmon resonance sensors: review,” Sens. Actuators B Chem. 54(1-2), 3–15 (1999). [CrossRef]  

5. G. Nenninger, P. Tobiška, J. Homola, and S. Yee, “Long-range surface plasmons for high-resolution surface plasmon resonance sensors,” Sens. Actuators B Chem. 74(1-3), 145–151 (2001). [CrossRef]  

6. R. Slavík and J. Homola, “Ultrahigh resolution long range surface plasmon-based sensor,” Sens. Actuators B Chem. 123(1), 10–12 (2007). [CrossRef]  

7. P. Yeh, A. Yariv, and A. Cho, “Optical surface waves in periodic layered media,” Appl. Phys. Lett. 32(2), 104–105 (1978). [CrossRef]  

8. R. D. Meade, K. D. Brommer, A. M. Rappe, and J. D. Joannopoulos, “Electromagnetic Bloch waves at the surface of a photonic crystal,” Phys. Rev. B Condens. Matter 44(19), 10961–10964 (1991). [CrossRef]   [PubMed]  

9. W. M. Robertson, G. Arjavalingam, R. D. Meade, K. D. Brommer, A. M. Rappe, and J. D. Joannopoulos, “Observation of surface photons on periodic dielectric arrays,” Opt. Lett. 18(7), 528–530 (1993). [CrossRef]   [PubMed]  

10. A. P. Vinogradov, A. V. Dorofeenko, S. G. Erokhin, M. Inoue, A. A. Lisyansky, A. M. Merzlikin, and A. B. Granovsky, “Surface state peculiarities in one-dimensional photonic crystal interfaces,” Phys. Rev. B 74(4), 045128 (2006). [CrossRef]  

11. M. Bergmair and K. Hingerl, “Band structure and coupled surface states in one-dimensional photonic crystals,” J. Opt. A 9(9), 339–344 (2007). [CrossRef]  

12. Y. C. Hsu and L. W. Chen, “Bloch surface wave excitation based on coupling from photonic crystal waveguide,” J. Opt. 12(9), 095709 (2010). [CrossRef]  

13. J. Barvestani, M. Kalafi, A. Soltani-Vala, and A. Namdar, “Backward surface electromagnetic waves in semi-infinite one-dimensional photonic crystals containing left-handed materials,” Phys. Rev. A 77(1), 013805 (2008). [CrossRef]  

14. L. L. Doskolovich, E. A. Bezus, and D. A. Bykov, “Phase-shifted Bragg gratings for Bloch surface waves,” Opt. Express 23(21), 27034–27045 (2015). [CrossRef]   [PubMed]  

15. W. M. Robertson, “Experimental measurement of the effect of termination on surface electromagnetic waves in one-dimensional photonic bandgap arrays,” J. Lightwave Technol. 17(11), 2013–2017 (1999). [CrossRef]  

16. Y. A. Vlasov, N. Moll, and S. J. McNab, “Observation of surface states in a truncated photonic crystal slab,” Opt. Lett. 29(18), 2175–2177 (2004). [CrossRef]   [PubMed]  

17. E. Descrovi, F. Giorgis, L. Dominici, and F. Michelotti, “Experimental observation of optical bandgaps for surface electromagnetic waves in a periodically corrugated one-dimensional silicon nitride photonic crystal,” Opt. Lett. 33(3), 243–245 (2008). [CrossRef]   [PubMed]  

18. M. Ballarini, F. Frascella, F. Michelotti, G. Digregorio, P. Rivolo, V. Paeder, V. Musi, F. Giorgis, and E. Descrovi, “Bloch surface waves-controlled emission of organic dyes grafted on a one-dimensional photonic crystal,” Appl. Phys. Lett. 99(4), 043302 (2011). [CrossRef]  

19. Y. Wan, Z. Zheng, W. Kong, X. Zhao, Y. Liu, Y. Bian, and J. Liu, “Nearly three orders of magnitude enhancement of Goos-Hanchen shift by exciting Bloch surface wave,” Opt. Express 20(8), 8998–9003 (2012). [CrossRef]   [PubMed]  

20. F. Michelotti, A. Sinibaldi, P. Munzert, N. Danz, and E. Descrovi, “Probing losses of dielectric multilayers by means of Bloch surface waves,” Opt. Lett. 38(5), 616–618 (2013). [CrossRef]   [PubMed]  

21. D. A. Shilkin, E. V. Lyubin, I. V. Soboleva, and A. A. Fedyanin, “Direct measurements of forces induced by Bloch surface waves in a one-dimensional photonic crystal,” Opt. Lett. 40(21), 4883–4886 (2015). [CrossRef]   [PubMed]  

22. F. Villa, L. E. Regalado, F. Ramos-Mendieta, J. Gaspar-Armenta, and T. Lopez-Ríos, “Photonic crystal sensor based on surface waves for thin-film characterization,” Opt. Lett. 27(8), 646–648 (2002). [CrossRef]   [PubMed]  

23. M. Shinn and W. M. Robertson, “Surface plasmon-like sensor based on surface electromagnetic waves in a photonic band-gap material,” Sens. Actuators B Chem. 105(2), 360–364 (2005). [CrossRef]  

24. E. Descrovi, F. Frascella, B. Sciacca, F. Geobaldo, L. Dominici, and F. Michelotti, “Coupling of surface waves in highly defined one-dimensional porous silicon photonic crystals for gas sensing applications,” Appl. Phys. Lett. 91(24), 241109 (2007). [CrossRef]  

25. V. N. Konopsky and E. V. Alieva, “Photonic crystal surface waves for optical biosensors,” Anal. Chem. 79(12), 4729–4735 (2007). [CrossRef]   [PubMed]  

26. F. Giorgis, E. Descrovi, C. Summonte, L. Dominici, and F. Michelotti, “Experimental determination of the sensitivity of Bloch surface waves based sensors,” Opt. Express 18(8), 8087–8093 (2010). [CrossRef]   [PubMed]  

27. A. Sinibaldi, E. Descrovi, F. Giorgis, L. Dominici, M. Ballarini, P. Mandracci, N. Danz, and F. Michelotti, “Hydrogenated amorphous silicon nitride photonic crystals for improved-performance surface electromagnetic wave biosensors,” Biomed. Opt. Express 3(10), 2405–2410 (2012). [CrossRef]   [PubMed]  

28. R. Rizzo, N. Danz, F. Michelotti, E. Maillart, A. Anopchenko, and C. Wächter, “Optimization of angularly resolved Bloch surface wave biosensors,” Opt. Express 22(19), 23202–23214 (2014). [CrossRef]   [PubMed]  

29. W. Kong, Z. Zheng, Y. Wan, S. Li, and J. Liu, “High-sensitivity sensing based on intensity-interrogated Bloch surface wave sensors,” Sens. Actuators B Chem. 193, 467–471 (2014). [CrossRef]  

30. A. Sinibaldi, N. Danz, E. Descrovi, P. Munzert, U. Schulz, F. Sonntag, L. Dominici, and F. Michelotti, “Direct comparison of the performance of Bloch surface wave and surface plasmon polariton sensors,” Sens. Actuators B Chem. 174, 292–298 (2012). [CrossRef]  

31. J. Homola, “Optical fiber sensor based on surface plasmon excitation,” Sens. Actuators B Chem. 29(1-3), 401–405 (1995). [CrossRef]  

32. D. Monzón-Hernández and J. Villatoro, “High-resolution refractive index sensing by means of a multiple-peak surface plasmon resonance optical fiber sensor,” Sens. Actuators B Chem. 115(1), 227–231 (2006). [CrossRef]  

33. A. K. Sharma, R. Jha, and B. D. Gupta, “Fiber-optic sensors based on surface plasmon resonance: a comprehensive review,” IEEE Sens. J. 7(8), 1118–1129 (2007). [CrossRef]  

34. M. Roussey, E. Descrovi, M. Häyrinen, A. Angelini, M. Kuittinen, and S. Honkanen, “One-dimensional photonic crystals with cylindrical geometry,” Opt. Express 22(22), 27236–27241 (2014). [CrossRef]   [PubMed]  

35. Y. Fink, J. N. Winn, S. Fan, C. Chen, J. Michel, J. D. Joannopoulos, and E. L. Thomas, “A dielectric omnidirectional reflector,” Science 282(5394), 1679–1682 (1998). [CrossRef]   [PubMed]  

36. H. A. Macleod, Thin-Film Optical Filters (CRC Press, 2010).

37. Z. Y. Li and L. L. Lin, “Photonic band structures solved by a plane-wave-based transfer-matrix method,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 67(4), 046607 (2003). [CrossRef]   [PubMed]  

38. Y. Zhao, F. Pang, Y. Dong, J. Wen, Z. Chen, and T. Wang, “Refractive index sensitivity enhancement of optical fiber cladding mode by depositing nanofilm via ALD technology,” Opt. Express 21(22), 26136–26143 (2013). [CrossRef]   [PubMed]  

39. Y. Matsuura, A. Hongo, M. Saito, and M. Miyagi, “Loss characteristics of circular hollow waveguides for incoherent infrared light,” J. Opt. Soc. Am. A 6(3), 423–427 (1989). [CrossRef]  

40. A. K. Sharma and B. D. Gupta, “Theoretical model of a fiber optic remote sensor based on surface plasmon resonance for temperature detection,” Opt. Fiber Technol. 12(1), 87–100 (2006). [CrossRef]  

41. Z. Bomzon, G. Biener, V. Kleiner, and E. Hasman, “Radially and azimuthally polarized beams generated by space-variant dielectric subwavelength gratings,” Opt. Lett. 27(5), 285–287 (2002). [CrossRef]   [PubMed]  

42. W. Chen, W. Han, D. Abeysinghe, R. Nelson, and Q. Zhan, “Generating cylindrical vector beams with subwavelength concentric metallic gratings fabricated on optical fibers,” J. Opt. 13(1), 015003 (2010). [CrossRef]  

43. R. Zheng, C. Gu, A. Wang, L. Xu, and H. Ming, “An all-fiber laser generating cylindrical vector beam,” Opt. Express 18(10), 10834–10838 (2011). [CrossRef]   [PubMed]  

44. B. H. Liu, Y. X. Jiang, X. S. Zhu, X. L. Tang, and Y. W. Shi, “Hollow fiber surface plasmon resonance sensor for the detection of liquid with high refractive index,” Opt. Express 21(26), 32349–32357 (2013). [CrossRef]   [PubMed]  

45. Y. X. Jiang, B. H. Liu, X. S. Zhu, X. L. Tang, and Y. W. Shi, “Long-range surface plasmon resonance sensor based on dielectric/silver coated hollow fiber with enhanced figure of merit,” Opt. Lett. 40(5), 744–747 (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 (9)

Fig. 1
Fig. 1 Schematic of 1DPC structure for exciting BSW.
Fig. 2
Fig. 2 Band structure of 1DPC. The left side represents PPL, while the right side represents SPL. (a) Band diagram presented in ω-β relationship. The slanted black line stands for the light line in the prism with refractive index n0 = 1.445. The red line stands for the critical condition for producing ORB. The blue line stands for the Brewster line, which crosses the intersection point of the boundaries of band gap of PPL. The gray areas refer to the energy band. (b) While n0 = 1.445, the band gap varies with the angle of incidence. The gray areas refer to band gaps, and the black area denotes ORB. (c) Same as above, except for n0 = 1.837, which is the critical condition. The range of ORB reduces to zero in this circumstance. (d) Same as above, except for n0 = 2.135, which is the case of the Brewster line. Band gap of PPL vanishes as the incident angle increases to 90°.
Fig. 3
Fig. 3 Reflection spectrum of designed 1DPC. Blue and red lines represent the reflectance of SPL and PPL, respectively.
Fig. 4
Fig. 4 Thickness of cap layer affecting the resonant condition of BSW. (a) Reflectance of SPL varying with thickness of cap layer. (b) Reflectance of PPL varying with thickness of cap layer.
Fig. 5
Fig. 5 Schematic and ray transmission model of designed fiber sensor.
Fig. 6
Fig. 6 Spectrum of designed fiber sensor.
Fig. 7
Fig. 7 Detectibility of resonant dips for SPL and PPL. (a) Influence of divergence angle on FWHM and minimunm transmittance, with period number set as six. (b) Influence of number of periods on FWHM and minimum transmittance, with divergance angle equalling 3°. Blue and red lines stand for SPL and PPL, while circles and triangles denote FWHM and minimum transmittance, respectively.
Fig. 8
Fig. 8 Dependence of RW on RI of sensed medium for both SPL and PPL. The two inset graphs indicate the transmission spectrum under the corresponding sensed medium.
Fig. 9
Fig. 9 Predicted detection spectrum under transmittance-RI relationship. The inset graph shows the wavelength spectrum at corresponding points.

Equations (8)

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

cospd=cos κ 1 d 1 cos κ 2 d 2 1 2 ( η 1 η 2 + η 2 η 1 )sin κ 1 d 1 sin κ 2 d 2 .
| cos κ 1 d 1 cos κ 2 d 2 1 2 ( η 1 η 2 + η 2 η 1 )sin κ 1 d 1 sin κ 2 d 2 |>1.
M i =[ cos κ i d i isin κ i d i η i i η i sin κ i d i cos κ i d i ].
M= M i = ( M 1 M 2 ) N M C .
r= ( M 11 + M 21 η m ) η 0 ( M 21 + M 22 η m ) ( M 11 + M 21 η m ) η 0 +( M 21 + M 22 η m ) .
P In ( φ )= P 0 exp( φ 2 φ 0 2 ),
P Out = P In ( φ ) R N ( φ ),
T= 0 π 2 P Out ( φ ) dφ 0 π 2 P In ( φ )dφ .
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.