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

Modulation instability and highly sensitive optical fiber biosensor

Open Access Open Access

Abstract

The modulation instability (MI) is one of the interesting nonlinear phenomena which is used as a sensing mechanism. The performance of this mechanism is based on the shift of generated sidebands when the effective refractive index of the sensor is changed. The use of the MI process as a biosensing mechanism has capabilities such as high sensitivity, label-free and selective operation. In this paper, the novel optical fiber structures are proposed as lab-on-a-chip biosensors. In the proposed biosensors, capturing target biomolecules is achieved by immobilizing a capture layer. Consequently, after attaching the target layer the effective refractive index and the total dispersion of the structure are changed. As a result, the MI gain profile is changed and the generated sidebands (Stokes and anti-Stokes) around pump frequency are shifted. We demonstrate that the maximum sensitivity of these proposed sensors is 34.88 nm/nm. Moreover, the nonlinear material which is used in these biosensors is semiconductor doped glass (CS 3-68). This material due to the high nonlinear refractive index is an appropriate candidate to realize the MI process at low pump power. The used laser is a conventional laser 1550 nm with a power of 100 mW. The maximum length of the proposed biosensors is 0.199 cm which can realize the concept of lab-on-a-chip biosensors.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

The optical fiber-based technology is one of the most important approaches to the detection and analysis of different important parameters in biomedical issues. Nowadays, optical fiber-based biosensors are very popular in comparison with other biosensors based on surface plasmon resonance [1,2], ring resonator [3,4], interferometry [5,6], etc. The reasons for this popularity are the undeniable advantages such as high sensitivity, simple configuration, low fabrication cost, low transmission loss [7,8]. Furthermore, the optical fiber biosensors based on Fiber Bragg grating, fiber coupler, and fiber Fabry-Perot cavities are optical sensors in which label-free detection is accessible [911]. However, these fiber-based biosensors in order to be label-free, require some excessive processes such as the creation of grating and etching the cladding layer. Moreover, these biosensors usually cannot be used as lab-on-a-chip biosensors and their sensing mechanism are not all-optical. To cope with these problems and design ultra-high sensitive sensors, the nonlinear features of the optical fiber can be applied.

Nonlinear optical phenomena are one of the sensing mechanisms in different types of sensors [12,13]. Some of these phenomena, such as second harmonic generation (SHG), third-harmonic generation (THG), etc., cause the generation of high-frequency components [14]. Since the generated harmonics are far away from the input spectrum, the sensing mechanism based on the shift of these harmonics requires detectors with different spectral regions. The difference between the spectrum of input source and detector makes the sensing process to be difficult. In order to overcome this problem, the generated sidebands should be in the input spectrum range. One of the nonlinear phenomena which create the generation of sidebands around the pump frequency is modulation instability (MI). The creating agent of the MI process is the interaction between input photons and phonons of media which leads to the exponential growth of weak perturbation such as quantum noise in the dispersive medium [1419]. In this process, sidebands are created when all the involved waves are phase-matched [14,2022]. The generated sidebands are generated during the MI process, are symmetric around the pump frequency because of energy conservation law. These sidebands which have maximum amplitude are referred to as Stokes and anti-Stokes sidebands. The shift of these sidebands could be an appropriate mechanism for sensing. This mechanism is used in the biosensor, strain sensor, refractive index sensor, magnetic field sensor, and temperature sensor [2329].

In the MI-based fiber-optic biosensor, the sensing mechanism is based on the shift of generated sidebands (Stokes and anti-Stokes) after capturing the target biomolecules. By using this method two different biosensors are proposed [26,27]. One of these biosensors is based on the photonic crystal fiber which its length is 50cm and the peak pump power is 500 W. The sensitivity of this sensor is 10.4 nm/nm [26]. The other biosensor is based on a chalcogenide tapered fiber. The length of this sensor is 1.2 cm, the pump power is 500mW, and the sensitivity of this biosensor is 17.8 nm/nm [27]. The sensitivity of MI-based biosensors is defined based on surface sensitivity, which refers to the shift of the generated sidebands per nanometer of the target molecule at the sensor surface [30,31]. The thickness and refractive index of the biolayer have important roles in sensor sensitivity. According to the concept of surface sensitivity, light propagation in the proposed sensors is such a way that it has maximum interaction with the sensor surface and the attached target molecule. Therefore, maximum sensitivity is obtained.

One of the most crucial factors in the fiber-based biosensors is obtaining the lab-on-a-chip concept [3235]. In order to achieve this concept in MI-based biosensors, the length of the sensor should be extremely short. To realize this aim, the high nonlinear refractive index materials are selected in order to manufacture biosensors.

In this paper, given all the recent activities in the MI-based sensors field, the importance of DNA detection [36,37], and as well as the advances in optical fiber design and fabrications [3843], we move towards proposing new structures of optical fiber in order to achieve highly sensitive, label-free, and lab-on-a-chip biosensors. In the proposed MI-based biosensors, after capturing the target biomolecules, Stokes and anti-Stokes sidebands are shifted more than the other MI-based biosensors and, as a result, the proposed sensors are more sensitive. Furthermore, by using the highly nonlinear material in proposed structures, we obtained short-length sensors at low pump power.

This paper is organized as follows: In section 2 theoretical modeling of sensing mechanism and MI process are presented. In section 2 new structures of MI-based optical fiber biosensors are proposed. Section 3 consists of the simulation results and also discussions about the advantages of the proposed structure. Finally, the paper ends with a conclusion.

2. Mechanism and structure of the sensing

The sensing mechanism of the proposed biosensors is based on the shift of Stokes or anti-Stokes sidebands. These sidebands which are generated during the MI process, are symmetric around the pump frequency because of energy conservation law. These sidebands depend on the effective refractive index and total dispersion which consists of material and waveguide dispersion. Therefore, Stokes or anti-Stokes sidebars can be shifted due to changes in total dispersion. These changes in the dispersion can be done through the presence of additional material in propagating medium. This material could be a biolayer that is attached to the sensor. Also, due to the importance of DNA detection in the biological world [36,37], we have considered DNA as the target biolayer.

2.1 MI process

The MI process is similar to degenerate four-wave mixing (FWM), which is the basic principle for detection mechanism in nonlinear optical fiber biosensors [26,27]. The MI is due to the interaction between anomalous group velocity dispersion and the Kerr nonlinearity effect. The effect can be explained using linear stability analysis performed on analytical solutions to the nonlinear Schrödinger equation with considering second-order dispersion [14,44,45]:

$$\frac{{\partial A}}{{\partial z}} + i\frac{{{\beta _2}}}{2}\frac{{{\partial ^2}A}}{{\partial {t^2}}} - i\gamma {|A |^2}A = 0$$
where $A$ is the slowly varying amplitude of the electric field envelope, $\beta {}_2$ is the second-order dispersion coefficient, $\gamma ({{\omega_0}} )= {{{n_2}{\omega _0}} / {c{A_{eff}}}}$ is the nonlinearity parameter, c is the speed of light in vacuum, ${\omega _0}$ is the central frequency of pump, $A{}_{eff}$ is effective-area and $n{}_2$ is the nonlinear refractive index. By solving this equation using the perturbation theory, can be obtained MI gain relation. However, anomalous dispersion (${\beta _2}\langle 0$) is necessary for realizing the MI process [14,22,29,30,46]. In this case (${\beta _2}\langle 0$), sidebands around the pump signal are generated and grow exponentially. Here the gain spectrum is given as:
$$g(\Omega )= |{{\beta_2}\Omega } |\sqrt {\frac{1}{{2|{{\beta_2}} |{L_{NL}}}} - \frac{{{\Omega ^2}}}{{16}}} ,\quad \;{L_{NL}} = \frac{1}{{\gamma {P_0}}}.$$
${L_{NL}}$ is nonlinear length, ${P_0}$ is the peak power of the pump. This equation indicates that the first-order dispersion ${\beta _1}$ has no effect on the MI gain profile. The frequencies with the maximum gain in the MI output spectrum are known as Stokes and anti-Stokes frequencies which are critical parameters in the sensing process. The Stokes and anti-Stokes frequencies are considered as ${\omega _s}$ and ${\omega _{as}}$ in Eq. (3), respectively.
$${\omega _s} = {\omega _0} - {\Omega _{\max }}\quad ,\quad {\omega _{as}} = {\omega _0} + {\Omega _{\max }}\quad ,\quad {\Omega _{\max }} = \sqrt {\frac{{4\gamma {P_0}}}{{|{{\beta_2}} |}}}$$
where ${\Omega _{\max }}$ is the frequency in which the MI gain is maximum. According to Eq. (2), variations in the effective refractive index cause a change in the MI gain profile. As a result, Stokes and anti-Stokes sidebands are shifted. The variations in the effective refractive index may be occurred through attaching the biolayer. Hence the changing in the profile of MI gain can be used for biosensing.

2.2 Proposed structures for biosensor

We investigated different geometric shapes for the proposed structure. Three cases of the best shapes are shown in Fig. 1. These structures consist of four and five cores which are connected together through holders within a cladding. The space between cores and cladding is filled by an aqueous solution during sensing. The use of an aqueous environment increases the interaction of light and the target layer. ${t_{c - layer}}$ represents capture layer that is immobilized on the outer of cores. This layer is used to catch the target biomolecules (DNA). These target molecules are illustrated in Fig. 1 by ${t_{b - layer}}$.

 figure: Fig. 1.

Fig. 1. The proposed setup and structures for MI-based optical fiber biosensors. The dimensions of the biosensor structures (diameters of central and lateral cores, the width of holders, the space between central and lateral cores, the space between lateral cores and cladding) are listed in Table 1. ${t_{c - layer}}$ (=40 nm) and ${t_{b - layer}}$ (=5 nm) are thicknesses of capture layer and target layer.

Download Full Size | PDF

The material of the structures is CS 3-68 which is a conventional glass doped by CdS, Se, and S. This material is selected because of its high nonlinear properties. The nonlinear refractive index (${n_2}$) of CS 3-68 is equal to $2.3 \times {10^{ - 10}}{\raise0.5ex\hbox{$\scriptstyle {c{m^2}}$}/\lower0.25ex\hbox{$\scriptstyle W$}}$[47,48]. This value is about ${10^3}({10^6})$ times higher than the nonlinear refractive index of chalcogenide glasses (silica) [49]. Using this material causes the length of the structure to be reduced and also lowers the required pump power. The shorter length of the sensor makes it suitable to be integrated and consequently applicable in lab-on-a-chip.

3. Simulation and discussion

The dimensions of the biosensor structures which are illustrated in Fig. 1 are given in Table 1.

Tables Icon

Table 1. Dimensions of the proposed structures.

Due to the advances in the design and fabrication of optical fibers [3843], the dimensions of these parameters are considered in such a way that they can be fabricated.

The modal analysis is performed for the biosensor using the finite difference eigenmode (FDE) solver. The analysis is done in the optical C-band around the wavelength of 1.55$\mu m$. The sources which can generate the optical power around this wavelength are easily available. The pump power which is used in the simulation is 100mW. It is five times less than previous research [27]. Regarding the minimum dimension in the structures is 5nm, the mesh step is selected 2nm in order to accurate solving. The refractive indexes of capture and target layers are considered to be 1.45 [26,27,50]. The simulation results of propagation mode profiles in three different situations are illustrated in Fig. 2. Figure 2(a), Fig. 2(d) and Fig. 2(g) demonstrate the mode profiles without capture and target layers in three proposed structures. Considering the target biomolecule to be a simple double-stranded DNA, the capture layer like poly-L-lysine should have at least 40nm thickness [26,50]. The mode profiles of three proposed structures after adding the capture layer are illustrated in Fig. 2(b), Fig. 2(e) and Fig. 2(h). After capturing the target biomolecule with the thickness of 5nm, the mode profiles of the three proposed structures are shown in Fig. 2(c), Fig. 2(f) and Fig. 2(i) [26]. Comparison between the propagation mode profiles shows significant changes. Especially after capturing the target biomolecule.

 figure: Fig. 2.

Fig. 2. Propagation mode profiles of the proposed structures. (a), (d), (g): without any additional layers. (b), (e), (h): after adding capture layer. (c), (f), (i): after capturing the target biomolecule by capture layer.

Download Full Size | PDF

The dispersions curves ($D = {{2\pi c{\beta _2}} / {{\lambda ^2}}}$) and second-order dispersions curves (${\beta _2}$) of the proposed structures are shown in Fig. 3. Figure 3(a), Fig. 3(b) and Fig. 3(c) demonstrate dispersions curves of proposed structures. The second-order dispersions curves of proposed structures are illustrated in Fig. 3(d), Fig. 3(e) and Fig. 3(f), respectively. These results are obtained by frequency analysis in the wavelength range 1-1.7$\mu m$.

 figure: Fig. 3.

Fig. 3. (a),(d): Dispersion and second-order dispersions curves of the structure 1. (b),(e): dispersion and second-order dispersions curves of the structure 2. (c),(f): dispersion and second-order dispersions curves of the structure 3. i: without any additional layers, ii: after adding capture layer, iii: after capturing the target biomolecule by capture layer.

Download Full Size | PDF

According to the results of Fig. 2, it is expected that the distribution of mode shape and optical power in additional cores leads to high dispersion. This fact is clearly visible in Fig. 3 and Table 2. As an instance, in structure 3 at the wavelength of 1550 nm, the dispersion value is modified from +40.53 to +18.49 after capturing the target biomolecule and causes the foremost of power to be guided in the central core.

Tables Icon

Table 2. The optical specifications of the proposed structures at $\lambda = 1550nm$

In order to obtain parameters of the MI process as the sensing mechanism in the proposed structures, the values of Table 2 are applied in Eq. (2) and Eq. (3). Consequently, the maximum gain frequency, the Stokes, and anti-Stokes frequencies are calculated and given in Table 3. As the length of the sensor is sufficiently short, the propagation loss is negligible and as a result, this loss does not affect the MI gain.

Tables Icon

Table 3. The Values of the MI parameters in the proposed structures.

The obtained MI gain profiles for the proposed structures are shown in Fig. 4. The capture of the target layer changes the total dispersion and effective refractive index of the structures. As a result, the MI gain profiles change, and the stokes and anti-stokes sidebands shift to the lower and higher frequencies, respectively. This frequency shift of the sidebands is the sensing mechanism principle. The maximum MI gain that occurs in stokes and anti-stokes sidebands, is obtained from Eq. (2), ${g_{\max }} = g({{\Omega _{\max }}} )= \gamma {P_0}$. After capturing the target layer, the effective area of the proposed structures is increased (based on Table 2) and consequently, the maximum MI gain is slightly decreased. The MI gain profiles and their variations for the proposed structures are demonstrated in Fig. 4(a), Fig. 4(b), and Fig. 4(c). Based on the MI gain profiles, the pump signal, stokes, and anti-stokes wavelengths are highlighted in Fig. 4(d) Fig. 4(e), and Fig. 4(f).

 figure: Fig. 4.

Fig. 4. (a), (b), (c): MI gain profiles of the proposed structures. (d), (e), (f): Stokes, anti-Stokes, and pump peak of the proposed structures. i: without any additional layers, ii: after adding capture layer, iii: after capturing the target biomolecule by capture layer.

Download Full Size | PDF

The shift of the Stokes sideband is greater than the anti-Stokes sideband in the wavelength domain. As an instance, in structure 3 with regard to pump wavelength (1550 nm), the Stokes and anti-Stokes wavelengths are 1864.15 nm and 1326.51 nm, respectively. After capturing the target layer the Stokes and anti-Stokes wavelengths shift to 2038.57 nm and 1250.38 nm, respectively. For this reason, the Stokes sideband is used to express the sensitivity. As mentioned, the sensitivity of the biosensor is defined as the shift of the Stokes sidebands per thickness of target biomolecules [26,27]. Hence, the sensitivity is defined in Eq. (4) which ${\lambda _{s2}}$, ${\lambda _{s1}}$ and t Stokes wavelength after immobilizing the capture layer, Stokes wavelength after capturing the target layer, and thickness of the target layer, respectively.

$$S = \frac{{{\lambda _{s2}}\, - \,{\lambda _{s1}}\,}}{t}.$$

In order to realize the MI process and lab-on-a-chip concept simultaneously, the length of the sensor should be at least the nonlinear length of the proposed structures. The sensitivity and length of the proposed biosensors are given in Table 4. At best (structure 3), the proposed biosensor is about 2 times more sensitive than the previous MI-based biosensor. Furthermore, in this case, the length of the proposed biosensor is more compatible than the previous works with lab-on-a-chip technology [27].

Tables Icon

Table 4. The sensitivity and length of the proposed biosensors.

4. Conclusion

In this paper, MI-based fiber biosensors are introduced whose detection is label-free and selective. The sensing mechanism is based on shifting the Stokes sideband in the MI process after capturing the target layer (DNA). In order to achieve a highly sensitive biosensor, the three structures are presented. The material of the proposed structures is considered to be CS 3-68 glass which has high nonlinear refractive. As a result, the pump power required (in order to occur MI process) and the length of the proposed biosensors are reduced. The reduction of the biosensor length realizes the Lab-on-a-chip concept. Another advantage of the proposed structures is that high sensitivity can be achieved using conventional low-power laser sources. In the best case, we achieved a sensitivity of 34.88 nm/nm and a sensor length of 0.199 cm when the pump power is 100mW at wavelength 1550 nm.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. X. Fan, I. M. White, S. I. Shopova, H. Zhu, J. D. Suter, and Y. Sun, “Sensitive optical biosensors for unlabeled targets: A review,” Anal. Chim. Acta 620(1-2), 8–26 (2008). [CrossRef]  

2. R.-B. Hwang, “A theoretical design of evanescent wave biosensors based on gate-controlled graphene surface plasmon resonance,” Sci. Rep. 11(1), 1–10 (2021). [CrossRef]  

3. K. De Vos, I. Bartolozzi, E. Schacht, P. Bienstman, and R. Baets, “Silicon-on-Insulator microring resonator for sensitive and label-free biosensing,” Opt. Express 15(12), 7610–7615 (2007). [CrossRef]  

4. Y. R. Bawankar and A. Singh, “Microring Resonators Based Applications in Silicon Photonics-A Review,” in 2021 5th Conference on Information and Communication Technology (CICT), 2021, pp. 1–6.

5. A. Psarouli, A. Botsialas, A. Salapatas, G. Stefanitsis, D. Nikita, G. Jobst, N. Chaniotakis, D. Goustouridis, E. Makarona, P. S. Petroua, I. Raptis, K. Misiakos, and S. E. Kakabakos, “Fast label-free detection of C-reactive protein using broad-band Mach-Zehnder interferometers integrated on silicon chips,” Talanta 165, 458–465 (2017). [CrossRef]  

6. X. Li, N. Chen, X. Zhou, P. Gong, S. Wang, Y. Zhang, and Y. Zhao, “A review of specialty fiber biosensors based on interferometer configuration,” J. Biophotonics 14(6), e202100068 (2021). [CrossRef]  

7. M. E. Bosch, A. J. R. Sánchez, F. S. Rojas, C. B. Ojeda, “Recent development in optical fiber biosensors,” Sensors 7(6), 797–859 (2007). [CrossRef]  

8. V. Naresh and N. Lee, “A review on biosensors and recent development of nanostructured materials-enabled biosensors,” Sensors 21(4), 1109 (2021). [CrossRef]  

9. M. Hernaez, C. R. Zamarreño, S. Melendi-Espina, L. R. Bird, A. G. Mayes, F. J. Arregui, “Optical fibre sensors using graphene-based materials: A review,” Sensors 17(12), 155 (2017). [CrossRef]  

10. H. Tazawa, T. Kanie, and M. Katayama, “Fiber-optic coupler based refractive index sensor and its application to biosensing,” Appl. Phys. Lett. 91(11), 113901 (2007). [CrossRef]  

11. T. Wei, Y. Han, H.-L. Tsai, and H. Xiao, “Miniaturized fiber inline Fabry-Perot interferometer fabricated with a femtosecond laser,” Opt. Lett. 33(6), 536–538 (2008). [CrossRef]  

12. C. B. De Araújo, A. S. L. Gomes, and G. Boudebs, “Techniques for nonlinear optical characterization of materials: a review,” Rep. Prog. Phys. 79(3), 036401 (2016). [CrossRef]  

13. I. Floris, J. M. Adam, P. A. Calderón, and S. Sales, “Fiber optic shape sensors: A comprehensive review,” Opt. Lasers Eng. 139, 106508 (2021). [CrossRef]  

14. G. P. Agrawal, Nonlinear Fiber Optics (Academic Press, New York, USA). Chapter, 2007.

15. G. P. Agrawal, “Modulation instability in erbium-doped fiber amplifiers,” IEEE Photonics Technol. Lett. 4(6), 562–564 (1992). [CrossRef]  

16. A. Hasegawa, “Generation of a train of soliton pulses by induced modulational instability in optical fibers,” Opt. Lett. 9(7), 288–290 (1984). [CrossRef]  

17. G. P. Agrawal, P. L. Baldeck, and R. R. Alfano, “Modulation instability induced by cross-phase modulation in optical fibers,” Phys. Rev. A 39(7), 3406–3413 (1989). [CrossRef]  

18. G. Genty, F. Dias, B. Kibler, N. Akhmediev, and J. M. Dudley, “Akhmediev Breather dynamics and the nonlinear modulation instability spectrum,” in Nonlinear Optics and Applications IV, 2010. vol. 7728, p. 772812.

19. A. Govindarajan, B. A. Malomed, A. Mahalingam, and A. Uthayakumar, “Modulational instability in linearly coupled asymmetric dual-core fibers,” Appl. Sci. 7(7), 645 (2017). [CrossRef]  

20. V. E. Zakharov and A. A. Gelash, “Nonlinear stage of modulation instability,” Phys. Rev. Lett. 111(5), 054101 (2013). [CrossRef]  

21. V. E. Zakharov and L. A. Ostrovsky, “Modulation instability: the beginning,” Phys. D 238(5), 540–548 (2009). [CrossRef]  

22. S. A. Madani, M. Bahrami, and A. Rostami, “Multi-clad optical fiber design for ultra-wideband modulation instability,” Phys. Scr.2022.

23. B. Gu, W. Yuan, M. H. Frosz, A. P. Zhang, S. He, and O. Bang, “Nonlinear fiber-optic strain sensor based on four-wave mixing in microstructured optical fiber,” Opt. Lett. 37(5), 794–796 (2012). [CrossRef]  

24. N. Nagarajan, R. Yamunadevi, R. V. J. Raja, and G. J. Raj, “Ultra sensitive nonlinear fiber optics-based refractive index sensor using degenerate four wave mixing technique in photonic crystal fiber,” IEEE Sens. J. 18(16), 6607–6614 (2018). [CrossRef]  

25. N. Nallusamy, R. V. J. Raja, and G. J. Raj, “Highly sensitive nonlinear temperature sensor based on modulational instability technique in liquid infiltrated photonic crystal fiber,” IEEE Sens. J. 17(12), 3720–3727 (2017). [CrossRef]  

26. J. R. Ott, M. Heuck, C. Agger, P. D. Rasmussen, and O. Bang, “Label-free and selective nonlinear fiber-optical biosensing,” Opt. Express 16(25), 20834–20847 (2008). [CrossRef]  

27. C. Markos and O. Bang, “Nonlinear label-free biosensing with high sensitivity using As 2 S 3 chalcogenide tapered fiber,” J. Lightwave Technol. 33(13), 2892–2898 (2015). [CrossRef]  

28. N. Nallusamy, P. Zu, R. V. J. Raja, N. Arzate, and D. Vigneswaran, “Degenerate four-wave mixing for measurement of magnetic field using a nanoparticles-doped highly nonlinear photonic crystal fiber,” Appl. Opt. 58(2), 333–339 (2019). [CrossRef]  

29. Y. Sun, X. Yan, F. Wang, X. Zhang, S. Li, T. Suzuki, Y. Ohishi, and T. Cheng, “Theoretical investigation of an alcohol-filled tellurite photonic crystal fiber temperature sensor based on four-wave mixing,” Sensors 20(4), 1007 (2020). [CrossRef]  

30. J. Flueckiger, S. Schmidt, V. Donzella, A. Sherwali, D. M. Ratner, L. Chrostowski, and K. C. Cheung, “Sub-wavelength grating for enhanced ring resonator biosensor,” Opt. Express 24(14), 15672–15686 (2016). [CrossRef]  

31. D. Novoa and N. Y. Joly, “Specialty Photonic Crystal Fibers and Their Applications,” Crystals 11(7), 739 (2021). [CrossRef]  

32. A. Ricciardi, A. Crescitelli, P. Vaiano, G. Quero, M. Consales, M. Pisco, E. Esposito, and A. Cusano, “Lab-on-fiber technology: a new vision for chemical and biological sensing,” Analyst 140(24), 8068–8079 (2015). [CrossRef]  

33. G. Luka, A. Ahmadi, H. Najjaran, E. Alocilja, M. DeRosa, K. Wolthers, A. Malki, H. Aziz, A. Althani, and M. Hoorfar, “Microfluidics integrated biosensors: A leading technology towards lab-on-a-chip and sensing applications,” Sensors 15(12), 30011–30031 (2015). [CrossRef]  

34. B. D. Gupta, A. Pathak, and A. M. Shrivastav, “Optical Biomedical Diagnostics Using Lab-on-Fiber Technology: A Review,” Photonics 9(2), 86 (2022). [CrossRef]  

35. P. Vaiano, B. Carotenuto, M. Pisco, A. Ricciardi, G. Quero, M. Consales, A. Crescitelli, E. Esposito, and A. Cusano, “Lab on Fiber Technology for biological sensing applications,” Laser Photonics Rev. 10(6), 922–961 (2016). [CrossRef]  

36. N. A. Hawes, A. E. Fidler, L. A. Tremblay, X. Pochon, B. J. Dunphy, and K. F. Smith, “Understanding the role of DNA methylation in successful biological invasions: a review,” Biol. Invasions 20(9), 2285–2300 (2018). [CrossRef]  

37. Y. Yang, W. Li, and J. Liu, “Review of recent progress on DNA-based biosensors for Pb2+ detection,” Anal. Chim. Acta 1147, 124–143 (2021). [CrossRef]  

38. Y. Li, H. Xin, Y. Zhang, and B. Li, “Optical fiber technologies for nanomanipulation and biodetection: a review,” J. Lightwave Technol. 39(1), 251–262 (2021). [CrossRef]  

39. Z. Liu and H.-Y. Tam, “Fabrication and Sensing Applications of Special Microstructured Optical Fibers,” in Selected Topics on Optical Fiber Technologies and Applications, F. Xu and C. Mou, eds. IntechOpen, Rijeka, 2018. [CrossRef]  

40. J. Troles and L. Brilland, “13 - Chalcogenide microstructured optical fibers for infrared applications,” in Chalcogenide Glasses, J.-L. Adam and X. Zhang, eds. Woodhead Publishing, 2014, 411–437.doi: https://doi.org/10.1533/9780857093561.2.411.

41. D. Pysz, I. Kujawa, R. Stępień, M. Klimczak, A. Filipkowski, M. Franczyk, L. Kociszewski, J. Buźniak, K. Haraśny, and R. Buczyński, “Stack and draw fabrication of soft glass microstructured fiber optics,” Bull. Pol. Acad. Sci.: Tech. Sci. 62(4), 667–682 (2014). [CrossRef]  

42. V. S. Shiryaev, “Chalcogenide glass hollow-core microstructured optical fibers,” Front. Mater. 2, 24 (2015). [CrossRef]  

43. Y. Liang, Z. Yu, L. Li, and T. Xu, “A self-assembled plasmonic optical fiber nanoprobe for label-free biosensing,” Sci. Rep. 9(1), 7379 (2019). [CrossRef]  

44. S. B. Cavalcanti, J. C. Cressoni, H. R. da Cruz, and A. S. Gouveia-Neto, “Modulation instability in the region of minimum group-velocity dispersion of single-mode optical fibers via an extended nonlinear Schrödinger equation,” Phys. Rev. A 43(11), 6162–6165 (1991). [CrossRef]  

45. K. Nithyanandan, R. V. J. Raja, and K. Porsezian, “Modulational instability in a twin-core fiber with the effect of saturable nonlinear response and coupling coefficient dispersion,” Phys. Rev. A 87(4), 043805 (2013). [CrossRef]  

46. J. H. Li, T. T. Sun, Y. Q. Ma, Y. Y. Chen, Z. L. Cao, F. L. Xian, D. Z. Wang, and W. Wang, “Modulation instabilities in nonlinear two-core optical fibers with fourth order dispersion,” Optik 208, 164134 (2020). [CrossRef]  

47. R. W. Boyd, Nonlinear optics. Academic press, 2020.

48. S. K. Apte, B. B. Kale, R. S. Sonawane, S. D. Naik, S. S. Bodhale, and B. K. Das, “Homogeneous growth of CdS/CdSSe nanoparticles in glass matrix,” Mater. Lett. 60(4), 499–503 (2006). [CrossRef]  

49. A. Zakery and S. R. Elliott, “Optical properties and applications of chalcogenide glasses: a review,” J. Non. Cryst. Solids 330(1-3), 1–12 (2003). [CrossRef]  

50. L. Rindorf, J. B. Jensen, M. Dufva, L. H. Pedersen, P. E. Høiby, and O. Bang, “Photonic crystal fiber long-period gratings for biochemical sensing,” Opt. Express 14(18), 8224–8231 (2006). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. The proposed setup and structures for MI-based optical fiber biosensors. The dimensions of the biosensor structures (diameters of central and lateral cores, the width of holders, the space between central and lateral cores, the space between lateral cores and cladding) are listed in Table 1. ${t_{c - layer}}$ (=40 nm) and ${t_{b - layer}}$ (=5 nm) are thicknesses of capture layer and target layer.
Fig. 2.
Fig. 2. Propagation mode profiles of the proposed structures. (a), (d), (g): without any additional layers. (b), (e), (h): after adding capture layer. (c), (f), (i): after capturing the target biomolecule by capture layer.
Fig. 3.
Fig. 3. (a),(d): Dispersion and second-order dispersions curves of the structure 1. (b),(e): dispersion and second-order dispersions curves of the structure 2. (c),(f): dispersion and second-order dispersions curves of the structure 3. i: without any additional layers, ii: after adding capture layer, iii: after capturing the target biomolecule by capture layer.
Fig. 4.
Fig. 4. (a), (b), (c): MI gain profiles of the proposed structures. (d), (e), (f): Stokes, anti-Stokes, and pump peak of the proposed structures. i: without any additional layers, ii: after adding capture layer, iii: after capturing the target biomolecule by capture layer.

Tables (4)

Tables Icon

Table 1. Dimensions of the proposed structures.

Tables Icon

Table 2. The optical specifications of the proposed structures at λ = 1550 n m

Tables Icon

Table 3. The Values of the MI parameters in the proposed structures.

Tables Icon

Table 4. The sensitivity and length of the proposed biosensors.

Equations (4)

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

A z + i β 2 2 2 A t 2 i γ | A | 2 A = 0
g ( Ω ) = | β 2 Ω | 1 2 | β 2 | L N L Ω 2 16 , L N L = 1 γ P 0 .
ω s = ω 0 Ω max , ω a s = ω 0 + Ω max , Ω max = 4 γ P 0 | β 2 |
S = λ s 2 λ s 1 t .
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.