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

Hybrid layered polymer slot waveguide Young interferometer

Open Access Open Access

Abstract

We demonstrate a polymer slot waveguide Young interferometer coated with a bilayer of Al2O3/TiO2. The approach enables relaxed dimensions of the polymer waveguide which simplifies the fabrication of the structure with a resolution of 50 nm. The layers were coated by an atomic layer deposition technique. The feasibility of the device was investigated by exploiting the interferometric structure as a bulk refractive index sensor operating at 975 nm wavelength for detection of an ethanol-water solution. A refractive index change of 1 × 10−6 RIU with a sensing length of only 800 µm was detected. The approach confirms the possibility of realizing a low cost device with a small footprint and enhanced sensitivity by employing the TiO2 rails in the sides of the slot waveguide.

© 2016 Optical Society of America

1. Introduction

Polymers have been proved to be promising materials for single and multimode optical waveguide based devices [1]. Their interesting properties including low optical loss, transparency in a wide wavelength range, cost effectiveness, processing possibility and compatibility with various substrates and suitability for integration with other devices, have enabled these materials to be used in a variety of applications, i.e., electro-optical devices, sensors, couplers, modulators, and photonic crystals [2–6].

Slot waveguides are nanostructures that confine light at a subwavelength scale in a low refractive index material sandwiched between two narrow high refractive index rails. The field enhancement in these structures is based on the discontinuity of the electric field at the interface between high and low refractive index materials [7]. For that reason, most of the slot waveguides are fabricated in high refractive index materials such as silicon [8–10], in which the slot width should be around 100 nm in the 1.3 and 1.55 µm telecommunication windows. Unfortunately, the sub-micrometer dimensions may limit the application of the slot waveguides. Indeed, some infiltration difficulties may rise when the slot region has to be filled by non-linear materials or fluidic analytes [11]. Lower refractive index based slot waveguides allow a wider slot region and thus better infiltration possibilities. However, the trade-off is a possible leak of light in the rails and a lower slot mode confinement. Nevertheless, theoretical and experimental research results confirm the possibility of realizing these structures in low refractive index materials, such as polymers [12,13]. To achieve a good confinement in the slot region while maintaining a single mode operation, the slot walls have to be as high as possible [13,14]. A high aspect ratio (rail height versus width) leads to demanding fabrication methods that are not always compatible with low cost mass production processes [15,16].

Previously, we have shown theoretically that employing thin layers of high refractive index materials results in an enhanced light confinement in the slot region and relaxed dimensions by increasing the index contrast in the slot waveguide [17]. The thin high refractive index layer on polymer slot pushes the field in the slot area while reducing the field confinement in the polymer rails with lower aspect ratio.

Here we present the realization and experimental demonstration of a layered Young interferometer based on a polymer slot waveguide coated with high refractive index materials fabricated by low cost and mass production compatible methods, i.e., nanoimprinting [18] and atomic layer deposition (ALD) [19].

2. Theoretical design

The proposed Young interferometer includes a polymer slot waveguide coated with a high refractive index bilayer of Al2O3/TiO2. Figure 1 is a schematic view of the Young interferometer showing the cross section of the waveguide in the slot region and the tip of the adiabatic ridge to slot waveguide coupler. The input region of the interferometer contains a 2.5 µm ridge waveguide which is tapered to a 1 µm single mode waveguide.

 figure: Fig. 1

Fig. 1 Schematic view of our Young interferometer with the cross section of the waveguide in the slot waveguide and a tip of the adiabatic strip to slot waveguide coupler.

Download Full Size | PDF

The coupled light is evenly split into two arms spaced 50 µm from each other by a 3 dB Y-branch. In order to achieve a high confinement in the slot region as well as maintaining the monomodal behavior, the width and height of the polymer rails were set to 650 nm and 320 nm, respectively. For this configuration, Ormocore, which is an organic-inorganic polymer with a refractive index of 1.54 at λ = 975 nm, was selected for the polymer slot waveguide. This material has low absorption and a high glass transition temperature which allows the use of the ALD technique without material degradation. The polymer waveguide is covered with a 5 nm thick Al2O3 layer of refractive index of 1.63. This thin layer plays a twofold role in our fabrication process. It is a protective layer for polymer rails and an etching stop layer for the TiO2 layer in etching process. High refractive index (n = 2.34) TiO2 coatings only on the vertical sides have a thickness of 90 nm. Al2O3 and TiO2 were chosen mainly due to their transparency in the visible and near infrared wavelengths, and due to the possibility to be deposited by ALD at a relatively low process temperature (120 °C). Original slot width of 340 nm reduces to 150 nm after Al2O3/TiO2 bilayer coating.

The mode profile and field confinement shown in Figs. 2(a)-2(d) were calculated using the Fourier Modal Method (FMM) for TE polarization at the wavelength of 975 nm. This wavelength was selected due to 20 times lower water absorption in comparison to 1550 nm [20]. The thin layer of TiO2 above the polymer rails pushes the modal field towards the over-cladding layer which leads to a reduced field overlap in the slot waveguide. Further increase in TiO2 layer thickness results in leaky modes in both slot and ridge waveguides [Figs. 2(a) and 2(b)] and enables multimodal behavior as well. However, by applying the TiO2 layers only to the vertical sides of the polymer rails the mode is well confined in the slot region [Figs. 2(c) and 2(d)] and the electric field is enhanced in the slot. All the calculations have been done by assuming that water with refractive index of 1.327 is a cover medium. Our calculations show that with addition of the high refractive index layers, the mode is strongly localized in the slot region with the width of 150 nm [Fig. 2(d)], which is 50% wider than typical slot waveguides [8–10].

 figure: Fig. 2

Fig. 2 Simulation results of the amplitude of lateral component of the electric field (EX) of the quasi-TE mode in (a) ridge, (b) slot waveguide with a 90 nm of TiO2 layer all over the waveguides, with down etched TiO2 layer in (c) ridge, (d) slot waveguide, and (e) adiabatic strip to slot waveguide coupler.

Download Full Size | PDF

To enable efficient coupling of light from the ridge waveguide to the slot waveguide, the 1 µm wide ridge waveguide is gradually narrowing to the slot rail width [21]. Alongside the narrowing ridge, another slot rail is introduced with a 50 nm tip gradually expanding to the slot rail width. The length of this adiabatic ridge to slot waveguide coupler is 50 µm. The coupling process in the adiabatic coupler was simulated with the Finite Difference Time Domain (FDTD) method. Figure 2(e) shows the lateral electric field (Ex) propagation along the ridge to slot waveguide coupler.

3. Fabrication

The Young interferometer was fabricated by a combination of nanoimprinting and atomic layer deposition, which both are considered as low cost techniques. Since the nanoimprinting process is based on the replication, the quality of the master mold has an important effect on the quality of the final sample. With regard to the small dimensions of the structure, the fabrication of the polymer waveguide started by creating a silicon-HSQ master mold by electron beam lithography. HSQ (Hydrogen silsesquioxane) is a high resolution negative tone resist, which is often used for fabrication of small features (<600 nm) without etching. In addition, it has very small line width fluctuations compared to other e-beam resists [22]. Furthermore, it can be used as a stamp in the nanoimprinting process due to its mechanical properties [23].

The process involves a single patterning step without a plasma etching process for transferring the pattern. Therefore, additional surface roughness, potentially induced from the plasma etching, can be avoided. First, the silicon wafer was spin coated with a single layer of diluted HSQ (DOW corning Fox16) e-beam resist and methyl isobutyl ketone (MIBK) with the ratio of (1:1) to obtain a layer thickness of 320 nm. The resist layer thickness, transferred directly as a height of waveguide, can be controlled by adjusting the spin coating process. The layer was exposed by e-beam lithography (Vistec EBPG 5000 + ES HR) with an area dose of 9000 µC/m2 and developed with water diluted Microdeposit 351 developer. After development, the stamp was hardened for 3 hours at 300 °C. Figure 3 shows scanning electron microscope (SEM) images of the different parts of the Young interferometer on the Si-HSQ stamp: the rail walls are vertical and the surfaces are smooth.

 figure: Fig. 3

Fig. 3 SEM images of a) slot waveguide and b) strip to slot waveguide coupler and c) Y- junction on the Si-HSQ stamp.

Download Full Size | PDF

The fabrication process for the polymer waveguide has been illustrated in Fig. 4. Si-HSQ master stamp had the similar structure to the final imprinted sample. Therefore, in order to obtain the negative structure for imprinting, another mold was replicated from the master stamp. For that purpose, the Si-HSQ stamp was treated with an anti-adhesion layer and replicated by dispensing Ormostamp UV curable resist on a fused silica wafer resulting in the final working stamp (WS). The replicated WS was used for fabrication of the polymer slot waveguide by UV nanoimprinting. For fabrication of polymer waveguide, a thin layer of diluted Ormocore-maT1050 (1:11) was coated on an oxidized silicon wafer (3 µm thermal oxide layer) and baked at 120 °C on a hot plate for 10 min to remove the solvent.

 figure: Fig. 4

Fig. 4 Schematic illustration of fabrication process starting from replication of working stamp (WS) from Si-HSQ master stamp until fabrication of the final ALD-coated polymer waveguide structure.

Download Full Size | PDF

The final polymer structure was fabricated by pressing the polymer WS stamp against the Ormocore layer with a thickness of 470 nm with 9 bar pressure and curing for 60 s in Eitre® Obducat nanoimprinting equipment.

Figure 5(a) shows a SEM image of the cross section of the polymer slot waveguide without high refractive index layers: the aspect ratio in the polymeric rails and slot waveguide is less than one. This facilitates the fabrication of the structure by nanoimprinting. Figure 5(a) also illustrates the fact that because of chosen relaxed dimensions, the polymer rails are in upright position without a rounding error on top of the rails, which would decrease the field confinement in the slot and lead to decreased sensitivity of targeted biosensor [9]. Figure 5(c) depicts the adiabatic coupler with the tip width of 50 nm, which means the structure with an aspect ratio of 6 was fabricated successfully. It should be noted that according to the design of the structure, the reduced aspect ratio allows fabrication of details with the aforementioned resolution by simple nanoimprinting (the aspect ratio in the tip of the coupler part with a similar width in a low refractive index slot waveguide, without applying the TiO2 layers, would be about 16 which would make the fabrication of the details with this resolution nearly impossible). Figure 5(e) is the SEM image of the 3 dB Y- junction. By comparing Fig. 3 to Figs. 5(a), 5(c), and 5(e), it can be seen that the structure has been precisely replicated. After replication of the polymer waveguide, the structure was coated with a 5 nm Al2O3 layer at 120 °C following a 90 nm of TiO2 layer at the same temperature. Applying the ALD process with this temperature simultaneously hardens the polymer structure as well.

 figure: Fig. 5

Fig. 5 SEM images of (a) cross section of a polymer slot waveguide, (b) cross section of ALD- coated polymer slot waveguide after down etching of the top TiO2 layer, (c) strip to slot waveguide coupler on polymer, (d) down etched ALD-coated strip to slot waveguide coupler and (e) Y- junction on polymer waveguide, and (f) Y- junction in down etched ALD-coated polymer waveguide.

Download Full Size | PDF

Al2O3 and TiO2 layers were grown from TMA (trimethylaluminum) / H2O and TiCl4 (titanium tetrachloride) / H2O precursors, respectively, by the ALD equipment (Beneq TFS 200). By using a fluorine based plasma etching, the TiO2 layer was further removed from the top of the polymeric rails as well as from the bottom of the slot region. As it can be seen in Figs. 5(b) and 5(d), the layers remain smooth also after the etching process. The gap in the beginning of the Y-junction in polymer [Fig. 5(e)] is reduced to zero after ALD coating [Fig. 5(f)]. The Al2O3 layer appears to be a perfect etching stop layer for TiO2 in fluorine based etching and it also prevents any damage or surface roughness on the polymer rails due to etching. By applying this thin etching stop layer of Al2O3, the etching time can be adjusted with relaxed tolerances to achieve a repeatable fabricated structure similar to the design.

It is important to note that ALD is a unique deposition method which can provide a conformal coating with a precise thickness control. The coating thickness can be controlled by the thickness of the grown layer in each cycle, which are 0.069 nm and 0.042 nm for TiO2 and Al2O3 in this work, respectively. Combined with a selective etching between the two layers, the ALD method enables the deposition of TiO2 on vertical sidewalls of the slot waveguide leading to high confinement of light in the slot. Moreover, the stamp fabrication and pattern replication by the nanoimprint method can be carried out with a low aspect ratio as observable in Fig. 5.

A thick over-cladding of Ormocomp was spin coated on the sample to protect the waveguides from any damage when mounting a fluidic cell on the sample. A window with the length of 5 mm was opened on the sample by UV masking. This window was used for characterizing the sample for a sensor application.

4. Performance of Young interferometer sensor

Light from a wavelength stabilized laser (QFBGLD-980-5, Photonics LLC) emitting at 975 nm was coupled into a 2.5 µm wide input waveguide from tapered polarization maintaining fiber forming a spot about 2.5 µm. A TE polarized state was confirmed with an external polarizer prior to the measurement. The modal intensity distributions and interference patterns were imaged through a 40x microscope objective onto the camera (UI-3240CP-NIR-GL, IDS Imaging Development Systems GmbH). The flow cell was mounted on top of the chip and analyte solutions were pumped on the sample with a syringe pump (Nexus 3000, Chemyx Inc.) at the rate of 100 µL/min. Figure 6(a) shows the modal output of the two arms of the Young interferometer. The phase difference between the two arms due to the presence of the slot waveguide leads to an interferogram which can be observed when the objective was defocused and moved 1.135 mm away from the focus point [Fig. 6(b)].

 figure: Fig. 6

Fig. 6 (a) Output intensity profiles of the ALD-coated Young interferometer of the reference arm (right) and the sensing arm, including an 800 µm long slot waveguide section (left), and (b) interference pattern captured after applying water as a cover medium in the sensing window.

Download Full Size | PDF

In order to investigate the performance of the fabricated Young interferometer for detecting the refractive index changes of aqueous media, ethanol-water solution with different mass concentrations of 0.00178%, 0.0167%, and 0.178% were prepared. In a sensor device, with a Young interferometer as a transducer, the aqueous homogeneous refractive index sensing, which is calculated as Sh=neffnc., can be measured by measuring the phase shift Δφ in the interference pattern resulting from the change in refractive index of the cover medium Δnc as

Δφ=ΔφSΔφr=2π(Δneff)totalλL=2π[(Δneff)s(Δneff)r]λL.
ΔφS, Δφr, (Δneff)sand (Δneff)r represent the phase difference and effective refractive index change in the slot and ridge waveguides, respectively. Since in this sample, both arms are exposed to the variation of the cover medium within the sensing window, the measured phase shift is the difference between the phase shifts of the two arms. L is the interaction length, which in this device is the length of the slot waveguide, 800 µm. λ is the wavelength which is 975 nm in this work.

The refractive index variation of the cover medium Δnc is the refractive index difference between ethanol-water solutions with different mass concentrations CEth and DI water which can be calculated as [24]

Δnc=0.0006×C.Eth
After applying the analyte on the sensor, the phase shift can be tracked by observing the spatial displacement of the interference fringes. The interferogram image was captured with the camera in 3 s intervals and by using a discrete Fourier transform, the phase shift in every image was extracted. The processing can be described as follow. The Fourier transform of the image was calculated and the peak corresponding to the spatial frequency of the fringes in the image was extracted. The phase shift was calculated as the argument of the complex image [25]. Figure 7 shows the phase shift as a function of time when applying the ethanol-water solution with different concentrations in the sensing window and flushing it again with DI water. In order to deduce the induced phase shift from the mechanical and/or temperature variations during the measurement, the phase shift graphs were baseline corrected by monitoring the values for one minute before applying the sensing analyte in the sensing window. The smallest phase shift of 0.007 ± 0.003 rad was measured for an ethanol-water solution with a concentration of 0.00178% which corresponds to Δnc=1.068×106.

 figure: Fig. 7

Fig. 7 Measured phase shift after applying an ethanol-water solution with the concentration of a) 0.00178%, b) 0.0167%, and (c) 0. 178%. Red lines are the results after filtering.

Download Full Size | PDF

In the present ALD-coated polymer slot waveguide interferometer, the sensitivity of both arms was increased by applying high refractive index materials, which affects the total result of the device. According to the Eq. (1), the operation of the device for sensing application can be significantly improved by covering the reference arm. For instance, by covering the reference arm, the theoretical phase shift of the device for the ethanol-water solution with a concentration of 0.178%, is increased from 0.07 to 1.3 rad in this device [Fig. 8(a)]. Figure 8(a) also compares the total phase shift of the device against the refractive index change of the cover medium for both experimental and simulated results done by both FMM and OptiFDTD commercial software. The nonlinearity of the experimental results can be due to accumulation of ethanol on the surface of the device in the sensing window due to presence of the TiO2 layers. In order to confirm this assumption, we simulated the slot and ridge waveguide with a 3 nm of the additional layer with refractive index of the 1.5 on the sides of the rails, mimicking the adsorption of ethanol by the TiO2 layers. According to our simulations with FMM, for the ethanol-water solution with the concentration of 0.178% the phase shift decreases to 0.049 rad.

 figure: Fig. 8

Fig. 8 a) The comparison of the experimental and the theoretical values of phase shift calculated by FMM and Opti FDTD. The theoretical phase shift for the same device with the covered reference arm was calculated by FMM, and b) effective refractive index change of the ALD-coated polymer slot waveguide interferometer against the change in the refractive index of the cover medium. The slope of the linear fitting gives the homogeneous sensitivity Sh of the device based on the measured values.

Download Full Size | PDF

In order to calculate the homogeneous sensitivity Sh, the total effective index change of the device (Δneff)total was calculated by replacing Δφ with the measured values and assuming L = 800 µm and λ = 975 nm in Eq. (1). The homogeneous sensitivity of the device is the slope of the linear fitting for the calculated values of (Δneff)total versus Δnc related to different concentrations of the ethanol-water solution [Fig. 8(b)]. Therefore, the homogeneous sensitivity Sh of the device is obtained as 0.051.

5. Conclusion

A Young interferometer based on ALD-coated nanoimprint replicated slot waveguide structure was fabricated. Despite its prior complicated design, the fabrication remains reliable, low cost and repeatable. Applying the thin layer of Al2O3 avoids any roughness or damage of the polymer structure after the etching process. The thin vertical TiO2 layers on sides of the polymer slot waveguide increase the confinement in the slot waveguide with a small footprint and a low aspect ratio. Such a wide slot gives more space for infiltration of liquids and the relaxed dimensions simplify the fabrication of the smallest details by nanoimprinting. The device was characterized at 975 nm wavelength and the feasibility of its functionality for aqueous refractive index sensing at the same wavelength was experimentally demonstrated. A refractive index change of 1 × 10−6 RIU, with a sensing length of 800 µm, was measured. The sensitivity of the device was measured to be 0.051. It was shown that the fabricated low cost structure can be applied for bulk homogeneous sensing.

Acknowledgments

This work was supported by Finnish Funding Agency for Technology and Innovation TEKES project (No. 70008/13), Tekes FiDiPro project NP-NANO (No. 40315/13), and Academy of Finland grant (No. 284907).

References and links

1. H. Ma, A. K. Y. Jen, and L. R. Dalton, “Polymer-based optical waveguides: Materials, Processing, and Devices,” Adv. Mater. 14(19), 1339–1365 (2002). [CrossRef]  

2. A. L. Pyayt, “Guiding light in electro-optic polymers,” Polymers (Basel) 3(4), 1591–1599 (2011). [CrossRef]  

3. K. S. Lee, H. L. T. Lee, and R. J. Ram, “Polymer waveguide backplanes for optical sensor interfaces in microfluidics,” Lab Chip 7(11), 1539–1545 (2007). [CrossRef]   [PubMed]  

4. I. M. Soganci, A. La Porta, and B. J. Offrein, “Flip-chip optical couplers with scalable I/O count for silicon photonics,” Opt. Express 21(13), 16075–16085 (2013). [CrossRef]   [PubMed]  

5. R. Himmelhuber, R. A. Norwood, Y. Enami, and N. Peyghambarian, “Sol-gel material-anabled electro-optic polymer modulators,” Sensors (Basel) 15(8), 18239–18255 (2015). [CrossRef]   [PubMed]  

6. C. Liguda, G. Bottger, A. Kuligk, R. Blum, M. Eichb, H. Roth, J. Kunert, W. Morgenroth, H. Elsner, and H. G. Meyer, “Polymer photonic crystal slab waveguides,” Appl. Phys. Lett. 78(17), 2434–2436 (2001). [CrossRef]  

7. 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]  

8. P. A. Anderson, B. S. Schmidt, and M. Lipson, “High confinement in silicon slot waveguides with sharp bends,” Opt. Express 14(20), 9197–9202 (2006). [CrossRef]   [PubMed]  

9. F. Dell’Olio and V. M. Passaro, “Optical sensing by optimized silicon slot waveguides,” Opt. Express 15(8), 4977–4993 (2007). [CrossRef]   [PubMed]  

10. L. Vivien, D. Marris-Morini, A. Griol, K. B. Gylfason, D. Hill, J. Lvarez, H. Sohlström, J. Hurtado, D. Bouville, and E. Cassan, “Vertical multiple-slot waveguide ring resonators in silicon nitride,” Opt. Express 16(22), 17237–17242 (2008). [CrossRef]   [PubMed]  

11. C. A. Barrios, B. Sánchez, K. B. Gylfason, A. Griol, H. Sohlström, M. Holgado, and R. Casquel, “Demonstration of slot-waveguide structures on silicon nitride / silicon oxide platform,” Opt. Express 15(11), 6846–6856 (2007). [CrossRef]   [PubMed]  

12. M. Hiltunen, J. Hiltunen, P. Stenberg, S. Aikio, L. Kurki, P. Vahimaa, and P. Karioja, “Polymeric slot waveguide interferometer for sensor applications,” Opt. Express 22(6), 7229–7237 (2014). [CrossRef]   [PubMed]  

13. P. Bettotti, A. Pitanti, E. Rigo, F. De Leonardis, V. M. N. Passaro, and L. Pavesi, “Modeling of slot waveguide sensors based on polymeric materials,” Sensors (Basel) 11(12), 7327–7340 (2011). [CrossRef]   [PubMed]  

14. M. Hiltunen, J. Hiltunen, P. Stenberg, J. Petäjä, E. Heinonen, P. Vahimaa, and P. Karioja, “Polymeric slot waveguide at visible wavelength,” Opt. Lett. 37(21), 4449–4451 (2012). [CrossRef]   [PubMed]  

15. Z. Cui, Nanofabrication: Principles, Capabilities and Limits (Springer 2009).

16. D. Barbero, M. Saifullah, P. Hoffmann, H. Mathieu, D. Anderson, G. Jones, M. Well, and U. Steiner, “High-resolution nanoimprinting with a robust and reusable polymer mold,” Adv. Funct. Mater. 17(14), 2419–2425 (2007). [CrossRef]  

17. L. Ahmadi, J. Tervo, J. Saarinen, and S. Honkanen, “Enhanced sensitivity in polymer slot waveguides by atomic layer deposited bilayer coatings,” Appl. Opt. 52(33), 8089–8094 (2013). [CrossRef]   [PubMed]  

18. L. J. Guo, “Nanoimprinting: Methods and materials requirements,” Adv. Mater. 19(4), 495–513 (2007). [CrossRef]  

19. R. W. Johnson, A. Hultqvist, and S. F. Bent, “A brief review of atomic layer deposition: from fundamentals to applications,” Mater. Today 17(5), 236–246 (2014). [CrossRef]  

20. K. F. Palmer and D. Williams, “Optical properties of water in the near infrared,” J. Opt. Soc. Am. 64(8), 1107–1110 (1974). [CrossRef]  

21. A. Säynätjoki, L. Karvonen, T. Alasaarela, X. Tu, T. Y. Liow, M. Hiltunen, A. Tervonen, G. Q. Lo, and S. Honkanen, “Low-loss silicon slot waveguides and couplers fabricated with optical lithography and atomic layer deposition,” Opt. Express 19(27), 26275–26282 (2011). [CrossRef]   [PubMed]  

22. H. Namatsu, T. Yamaguchi, M. Nagase, K. Yamazaki, and K. Kurihara, “Nano-patterning of a hydrogen-silsesquioxane resist with reduced linewidth fluctuations,” Microelectron. Eng. 41–42, 331–334 (1998). [CrossRef]  

23. M. Haffner, A. Heeren, M. Fleischer, D. P. Kern, G. Schmidth, and L. W. Molenkamp, “Simple high resolution nanoimprint lithography,” Microelectron. Eng. 84(5-8), 937–939 (2007). [CrossRef]  

24. D. R. Lide, CRC Handbook of Chemistry and Physics (CRC, 2005).

25. X. Fan, Advanced Photonic Structures for Biological and Chemical Detection (Springer, 2009).

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 (8)

Fig. 1
Fig. 1 Schematic view of our Young interferometer with the cross section of the waveguide in the slot waveguide and a tip of the adiabatic strip to slot waveguide coupler.
Fig. 2
Fig. 2 Simulation results of the amplitude of lateral component of the electric field (EX) of the quasi-TE mode in (a) ridge, (b) slot waveguide with a 90 nm of TiO2 layer all over the waveguides, with down etched TiO2 layer in (c) ridge, (d) slot waveguide, and (e) adiabatic strip to slot waveguide coupler.
Fig. 3
Fig. 3 SEM images of a) slot waveguide and b) strip to slot waveguide coupler and c) Y- junction on the Si-HSQ stamp.
Fig. 4
Fig. 4 Schematic illustration of fabrication process starting from replication of working stamp (WS) from Si-HSQ master stamp until fabrication of the final ALD-coated polymer waveguide structure.
Fig. 5
Fig. 5 SEM images of (a) cross section of a polymer slot waveguide, (b) cross section of ALD- coated polymer slot waveguide after down etching of the top TiO2 layer, (c) strip to slot waveguide coupler on polymer, (d) down etched ALD-coated strip to slot waveguide coupler and (e) Y- junction on polymer waveguide, and (f) Y- junction in down etched ALD-coated polymer waveguide.
Fig. 6
Fig. 6 (a) Output intensity profiles of the ALD-coated Young interferometer of the reference arm (right) and the sensing arm, including an 800 µm long slot waveguide section (left), and (b) interference pattern captured after applying water as a cover medium in the sensing window.
Fig. 7
Fig. 7 Measured phase shift after applying an ethanol-water solution with the concentration of a) 0.00178%, b) 0.0167%, and (c) 0. 178%. Red lines are the results after filtering.
Fig. 8
Fig. 8 a) The comparison of the experimental and the theoretical values of phase shift calculated by FMM and Opti FDTD. The theoretical phase shift for the same device with the covered reference arm was calculated by FMM, and b) effective refractive index change of the ALD-coated polymer slot waveguide interferometer against the change in the refractive index of the cover medium. The slope of the linear fitting gives the homogeneous sensitivity S h of the device based on the measured values.

Equations (2)

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

Δφ=Δ φ S Δ φ r =2π ( Δ n eff ) total λ L=2π [ ( Δ n eff ) s ( Δ n eff ) r ] λ L.
Δ n c =0.0006×C . Eth
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.