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

High-performance Bloch surface wave biosensor based on a prism-coupled porous silicon composite structure for the detection of hemoglobin

Open Access Open Access

Abstract

Biosensors have various potential applications in biomedical research and clinical diagnostic, especially in detection of biomolecules in highly diluted solutions. In this study, a high-performance Bloch surface wave biosensor was constructed for the detection of hemoglobin. The procedure consisted of designing a porous silicon-based Kretschmann configuration to ensure excitation of the Bloch surface wave. The performance of the resulting sensor was then optimized by adjusting the buffer layer parameters based on the impedance matching method. The results showed an increase in the quality factor and figure of merit of the biosensor as a function of the decrease in thickness and refractive index of the buffer layer. The combination of the two optimization methods resulted in the quality factor and figure of merit of the optimized biosensor reaching as high as Q = 6967.4 and FOM = 11050RIU−1, respectively. In sum, the designed biosensor with high performance looks promising for future detection of hemoglobin.

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

1. Introduction

Sensors have gradually become the backbone of modern information technology in terms of technology development and applications [1]. So far, sensors based on optical signals have widely been used in medical diagnosis, environmental monitoring, food safety, defense industry, and other fields [2]. In particular, optical sensors are considered important devices for bio-sensing applications owing to their inherent advantages, such as strong anti-interference ability, remote sensing, and rapid response time [36]. In this category, optical biosensors collect changes in the information of test substances by monitoring the variations in certain characteristics of light in interaction with the test object.

Recently, rapid progress has been made in optical sensing and detection technology based on the surface wave. Surface plasmon wave (SPW), surface-enhanced Raman, surface-enhanced fluorescence, and other technologies have received increasing attention as typical representative technologies [710]. The most widely used in commercial applications are SPW-based biosensors, which are very sensitive to the refractive index changes at the sensor surface [9,10]. However, the loss and absorption of the metal film broaden the SPW resonance mode, thereby limiting the detection sensitivity and preventing the detection of small molecules. Thus, one key to achieving higher sensitivity is by narrowing the SPW resonance, which can be done by micro-ring resonator using the whispering-gallery-mode to yield ultrahigh quality factor (Q) values and extremely narrow resonances, thereby high resolution [11]. The disadvantages of this kind of sensor lie in probing light closely limited by the resonant, where only a very small part of the optical mode interacts with the analysis solution to yield relatively low sensitivities. The other has to do with the ultrahigh Q resonator with thermal instability and the difficulty in achieving experimental data [1113]. In such processes, Bloch surface wave (BSW) excited by a moderate-Q one-dimensional photonic crystal (1DPC) structure is a new bio-sensing mechanism [14], offering multiple design freedoms to modulate the Q value and electromagnetic field distributions. The BSW mode originates from the violation of a uniform waveguide system caused by edge effects, leading to a structure with a buffer layer on top of a 1DPC. In other words, an effective Fabry-Perot resonator forms in the buffer layer by the photonic band gap and the total internal reflection boundary [15,16]. The excited BSW mode strongly localizes at the interface between the biosensor and bio-solution. The strong field enhancement renders BSW greatly sensitive to changes in bio-solution. In reflected spectra of biosensors, BSW manifests itself as resonant depressions. Thus, the behaviors of biomolecules in bio-solution can be clarified by looking at shifts in the resonance spectra either of angular, wavelength, phase, or intensity domain [1722]. Variations in bio-solution concentration can induce slight changes in its refractive index, leading to shifts in BSW mode. The sensitivity of biosensors can be defined by the ratio between the BSW mode shift and the refractive index change. In addition, the sensing performances of biosensors can be quantified by their Q values and figure of merit (FOM) [2325].

Porous silicon (P-Si) is a new type of nanomaterial with a unique geometric structure and material properties, making it useful in biomedical sensing, optoelectronics, drug delivery, and energy conversion [2629]. Besides, P-Si is inexpensive and readily fabricated through the electrochemical etching of bulk silicon [30]. Moreover, the application of moderate current densities would tune the porosity of P-Si films during electrochemical etching, thereby modifying the refractive index. Also, films with different refractive indices can create 1DPC structures [3136]. Hence, P-Si-based 1DPC coupled with prism has been used to excite BSW and yield excellent sensing signals [3744].

As part of the standard complete blood count, the determination of total hemoglobin concentration along with the hematocrit and the mean erythrocyte hemoglobin are important factors. Independent optical methods can be used to evaluate these blood parameters and help develop quality control for blood analysis. This is called blood optics, which are principal for bio-photonic and clinical applications, both in therapy and diagnostics [45]. Blood cells include red blood cells, white blood cells, and platelets. Hemoglobin (Hb) is an important constituent of blood and its accurate measurement in the blood can help detect various diseases like anemia, diabetes, and leukemia [46,47]. Meanwhile, the blood refractive index depends on Hb concentration, with volume fraction routinely used to determine and monitor Hb [48]. For the detection of Hb and other biological molecules, low concentration markers are key requirements for early detection of diseases. This requires continuous efforts to develop highly sensitive techniques. For example, Pader et al. [49] demonstrated the application potential of a BSW-based sensor for the detection of protein aggregation $(40\deg /RIU)$. Goyal et al. [48] proposed a BSW based method for the detection of Hb with sensitivity $(69\deg /RIU)$. Rodriguez et al. [50] reported a size-selective grating coupled P-Si BSW biosensor with high sensitivity $(72\deg /RIU)$.

The performances of biosensors can be optimized by raising the Q values and the sensitivity of the probing light. However, such studies rely on numerical parameter sweeping, which provides little physical insights and sometimes cannot fully reflect the influence of system parameters [49,50]. As a result, the design of high-performance biosensors becomes a tedious task. The impedance matching method has been adopted to study this kind of problem theoretically, but the influences of the buffer layer have not yet been fully examined.

In this work, a high-performance Bloch surface wave biosensor was constructed for the detection of Hb. The performances of the resulting sensor were then optimized by adjusting the buffer layer parameters based on the impedance matching method. The results showed an increase in the Q and FOM of the biosensor as a function of the decrease in thickness and refractive index of the buffer layer. The combination of the two optimization methods resulted in the Q and FOM of the optimized biosensor reaching as high as $Q = 6967.4$ and $FOM = 11050RI{U^{ - 1}}$, respectively. In sum, the designed biosensor with high performance looks promising for future detection of Hb.

2. Theory of the optical biosensor

The design scheme of the BSW biosensor based on the Kretschmann configuration is shown in Fig. 1. Here, the prism was made of BK7 $({n_{pri}} = 1.457)$ and the considered 1DPC structure consisted of $|{(BA)^N}|C|$. The mediums A and $B$ corresponded to 50% P-Si and 80% P-Si, respectively. At a working wavelength much larger than the diameter of P-Si pores, the refractive index of the entire P-Si layer can be calculated by the Bruggeman dielectric constant approximation theory [34]:

$$({1 - \rho } )\frac{{\textrm{n}_{si}^2 - \textrm{n}_{eff}^2}}{{\textrm{n}_{si}^2 + 2\textrm{n}_{eff}^2}} + \rho \frac{{\textrm{n}_0^2 - \textrm{n}_{eff}^2}}{{\textrm{n}_0^2 + 2\textrm{n}_{eff}^2}} = 0$$
where ${n_0}\; $ represents the refractive index of vacuum, $\rho \; $ is the porosity of porous silicon, and ${n_{Si\; }}$ refers to the refractive index of silicon. According to Eq. (1), the refractive indices of A and $B$ were determined ${n_A} = 2.15$ and ${n_B} = 1.32$, respectively.

 figure: Fig. 1.

Fig. 1. Schematic diagram of the BSW biosensor.

Download Full Size | PDF

The thicknesses of mediums A and $B$ were set to ${d_A} = 151nm$ and ${d_B} = 246nm$, respectively. A buffer layer C made of $ZnS$ was inserted between the bio-solution and 1DPC. The $ZnS$ thickness was ${d_c} = 230nm$, and the refractive index was ${n_c} = 2.28$. The flat and hard surfaces of the biosensor were very convenient for manufacturing and cleaning, thereby reducing the cost of sensors. To balance the requirements of a high-Q reflected beam and compact configuration, the period number of 1DPC was set at $N = 6$. The probing light was assumed as TE wave, and the incident angle $\theta$ consisted of the angle between the probing light and the normal of the biosensor. Although TE wave was considered as probing light, the TM wave also can be used as the probing light. For the 1DPC structure, both TM wave and TE wave can generate the photonic bandgap necessary to excite BSW. Note that BSW excited by TM wave and TE wave corresponded to different system parameters.

The theoretical analysis revealed forward- and backward-decaying propagation patterns of the excited BSW near the surface [51,52]. The surface impedance at each side of the interface was defined as the ratio of the tangential electric and magnetic fields. The forward-decaying mode can be defined by Eq. (2):

$$Z_S^ +{=} \frac{{{\xi _C}({\xi _{Hb}} + {\xi _C}) - {\xi _C}{e^{2i{k_C}{d_C}}}({\xi _C} - {\xi _{Hb}})}}{{({\xi _{Hb}} + {\xi _C}) - {e^{2i{k_C}{d_C}}}({\xi _C} - {\xi _{Hb}})}}$$

The surface Bloch impedance of the backward-decaying mode can be determined according to Eq. (3):

$$Z_S^ -{=} \frac{{\xi _ + ^{BSW}({\xi _{pri}} - \xi _ - ^{BSW}) + \xi _ - ^{BSW}{e^{ - 2imK\Lambda }}(\xi _ + ^{BSW} - {\xi _{pri}})}}{{({\xi _{pri}} - \xi _ - ^{BSW}) + {e^{ - 2imK\Lambda }}(\xi _ + ^{BSW} - {\xi _{pri}})}}$$
where the normalized wave impedance for TE wave is ${\xi _i} = \omega {\mu _o}/{k_i}(i = pri,C,Hb)$. The normal components of the wave vector of medium $i = pri,A,B,C,Hb$ are ascribed to ${k_i} = \sqrt {({n_i}{k_0}) - {\beta ^2}}$. Here, $\beta = {n_{pri}}{k_0}\sin \theta$ refers to the propagation constant. $\theta$ and ${k_0}$ are the incidence angle and vacuum wave vector, respectively.

The positive and negative directions of Bloch impedances of the 1DPC in Eq. (3) can be expressed by Eq. (4):

$$\xi _ \pm ^{BSW} = \frac{{ \mp 2{\xi _A}{\xi _B}\sin (K\Lambda ) - i(\xi _A^2 - \xi _B^2)\sin ({k_A}{d_A})\sin ({k_B}{d_B})}}{{(\xi _A^{} - \xi _B^{})\sin ({k_A}{d_A} - {k_B}{d_B}) + (\xi _A^{} + \xi _B^{})\sin ({k_A}{d_A} + {k_B}{d_B})}}$$

According to the dispersion relationship, the Bloch wave vector of the 1DPC can be written by Eq. (5):

$$K\textrm{ = }\frac{1}{\Lambda }\arccos [\cos ({k_A}{d_A})\cos ({k_B}{d_B}) - \frac{1}{2}(\frac{{{k_A}}}{{{k_B}}} + \frac{{{k_B}}}{{{k_A}}})\sin ({k_A}{d_A})\sin ({k_B}{d_B})]$$
where $\Lambda $ is the thickness of the unit cell $(\Lambda = {d_A} + {d_B})$.

The surface impedance mismatch function of forwarding- and backward-decaying modes can be defined by Eq. (6):

$${Z_S} = |{Z_S^ +{-} Z_S^ - } |$$

The BSW would be excited when the impedance matching ${Z_S} = 0$. However, the changes in Hb molecules would destroy the balance, leading to impedance mismatch as a function of external variables ${n_{Hb}}$, $\theta$, and $\lambda$ for a given configuration.

$${Z_S} = \frac{{\partial {Z_S}}}{{\partial {n_{Hb}}}}\Delta {n_{Hb}} + \frac{{\partial {Z_S}}}{{\partial \lambda }}\Delta \lambda + \frac{{\partial {Z_S}}}{{\partial \theta }}\Delta \theta$$

Here, the Hb concentration and the refractive index of the bio-solution to be measured would change slightly. Namely, the first term on the right side of Eq. (7) would induce an impedance mismatch. To ensure excited BSW, the incident angle or probing wavelength was modified. The shift in resonant spectra of the incident angle domain required fixing the wavelength $\Delta \lambda = 0$. Thus, the incident angle sensitivity can be defined according to Eq. (8):

$${S_\theta } = \left|{\frac{{\Delta \theta }}{{\Delta {\textrm{n}_{H\textrm{b}}}}}} \right|= \left|{\frac{{\partial {Z_S}/\partial {n_H}}}{{\partial {Z_S}/\partial \theta }}} \right|$$

The Q can be used to describe the sharpness of the resonant peak, employed to express the relationship between the total resonant electromagnetic wave energy and the energy consumed by attenuation. This can be expressed by Eq. (9) in the spectrum calculation:

$$Q = \frac{\theta }{{FWHM}}$$
where FWHM represents the full width at half maximum of the resonance dip.

FOM can be used as an important parameter to measure the ability of the biosensor to respond to any changes in the resonance angle. FOM may be given as a ratio between the sensitivity S (deg /RIU) and $FWHM$:

$$FOM = \frac{S}{{FWHM}}$$

For further evaluation of the performance of the biosensor, the limit of detection (LOD) of the biosensor can be calculated as follows [53]:

$$LOD = \frac{{\Delta {n_{H\textrm{b}}}}}{{\Delta \theta }} \cdot \sigma$$
where $\sigma$ is the resolution of the instrument set as 0.005 deg.

3. Results and discussion

The performance of the resulting BSW biosensor in the detection of the Hb concentration was analyzed. A biosensor optimization scheme was built from the perspective of impedance matching. This was achieved by combining the optimization methods of the buffer layer thickness and refractive index to study the Q and FOM of the biosensor. To this end, the impedance mismatch function of the biosensor based on Eq. (6) was first plotted (red dashed line in Fig. 2(a)). The reflection spectrum of the biosensor was then calculated with the transfer matrix method (black solid line in Fig. 2(a)) [54]. As can be seen, the impedance matching and the reflection dip both occurred at $\theta \textrm{ = }72.311\deg$. Thus, the impedance matching condition can be used as the criterion to determine the BSW excitation, helping to build the BSW excitation system. In Fig. 2(b), the electric field intensity distribution inside the biosensor was given at the reflection dip $(\theta \textrm{ = }72.311\deg$ and $\lambda = 1060nm)$ [55]. The electric field of BSW looked strongly confined at the surface between the biosensor and the bio-solution, and an evanescent wave appeared in the bio-solution. These characteristics revealed BSW as excellent sensing signals for the detection of small changes in the bio-solution.

 figure: Fig. 2.

Fig. 2. (a) The variation in impedance mismatch function with the incident angle (red dashed line) and the reflection spectrum (black solid line) at incident wavelength $\lambda = 1060nm$. (b) The electric field distribution inside the biosensor at $\theta \textrm{ = }72.311\deg$ and $\lambda = 1060nm$.

Download Full Size | PDF

A displacement of the BSW reflectance spectrum as a function of the concentration of the Hb solution is provided in Fig. 3. By considering the laser sources and detectors, the incident wavelength was fixed as $\lambda = 1060nm$. The refractive index of Hb solution (${n_{Hb}}$) varied with Hb concentration (${c_{Hb}}$) and the incident wavelength. This can be calculated by the model function shown in Eq. (11) [56]:

$${n_{Hb}}(\lambda ,{c_{Hb}}) = {n_{{H_2}O}}(\lambda )[\alpha (\lambda ){c_{Hb}} + 1]$$

 figure: Fig. 3.

Fig. 3. The shifts in BSW reflectance spectra as a function of Hb concentration. The corresponding resonance angles were determined as ${\theta _1} = 72.320\deg ({c_{Hb}} = 10g/dl)$, ${\theta _2} = 72.414\deg ({c_{Hb}} = 11g/dl)$, ${\theta _3} = 72.578\deg ({c_{Hb}} = 12g/dl)$ and ${\theta _4} = 72.736\deg ({c_{Hb}} = 13g/dl)$, respectively.

Download Full Size | PDF

In Eqs. (8)–(11), the sensitivity, FWHM, Q, FOM, and LOD were set to ${S_\theta } = 52.49\deg /RIU$, $FWHM = 0.04\deg$, $Q = 1812$, $FOM = 1312.25RI{U^{ - 1}}$, and $LOD = 9.5\ast {10^{ - 5}}RIU$, respectively. These values were the average of four resonance dips, with errors within the allowable measurement range.

Since the buffer layer parameters significantly impact the impedance matching, the biosensor performances were optimized in terms of buffer layer parameters. To this end, the effect of the buffer layer thickness on the biosensor performance was first studied and the results are given in Fig. 4. Here, the variations in the sensitivity, FWHM, Q, and FOM of the biosensor with buffer layer thickness were all calculated. As shown in Fig. 4(a), the sensitivity of the biosensor increased as a function of the decrease in buffer layer thickness, while FWHM showed a declining trend. In Fig. 4(b), the Q and FOM of the biosensor also increased with the decrease in thickness of the buffer layer, attributed to the definition of Q and FOM. Fortunately, the performance of the biosensor can be doubled by reducing the thickness of the buffer layer from 240 nm to 215 nm. This process was very effective in improving the performance of the biosensor.

 figure: Fig. 4.

Fig. 4. The effect of the buffer layer thickness on the biosensor performance.

Download Full Size | PDF

The effect of the buffer layer refractive index on the biosensor performance was also considered. The variations in sensitivity, FWHM, Q, and FOM of the biosensor with buffer layer refractive index were all calculated and the results are gathered in Fig. 5. The sensitivity of the biosensor increased almost linearly as the buffer layer refractive index decreased (Fig. 5(a)). Also, the changes in FWHM were still not significant. In Fig. 5(b), the Q and FOM of the biosensor incremented with the decline in the refractive index of the buffer layer. At refractive indices, less than 2.27, the Q and FOM of the biosensor increased significantly, mainly due to the smaller FWHM in Fig. 5(a). Therefore, changes in the refractive index of the buffer layer can also improve the performance of biosensors.

 figure: Fig. 5.

Fig. 5. The effect of the buffer layer refractive index on the biosensor performance.

Download Full Size | PDF

Combining these two optimization methods in Figs. 4 and 5 resulted in buffer layer thickness of the optimized biosensor of ${d_c} = 215nm$ and buffer layer refract index of ${n_c} = 2.26$. The reflection spectrum of the optimized sensor is shown in Fig. 6 (the incident wavelength is 1060 nm). As mentioned earlier, the Hb concentration was supposed to evenly shift from $10g/dl$ to $13g/dl$ by a step of $1g/dl$. Hence, the shifts in the BSW reflectance spectra were calculated as a function of Hb concentration. The corresponding resonance angles were estimated to ${\theta _5} = 69.368\deg$ $({c_{Hb}} = 10g/dl)$, ${\theta _6} = 69.525\deg$ $({c_{Hb}} = 11g/dl)$, ${\theta _7} = 69.771\deg$ $({c_{Hb}} = 12g/dl)$, and ${\theta _8} = 70.031\deg ({c_{Hb}} = 13g/dl)$, respectively. The FWHM of the reflection dips and angle shifts between the neighboring reflection dips were recorded as $FWHM = 0.01\deg$ and $\Delta \theta = 0.221\deg$, respectively. By considering the slight variations in bio-solution refractive index $\Delta {n_{Hb}} = 0.002$, the biosensor sensitivity was calculated as ${S_\theta } = 110.5\deg /RIU$ and LOD was determined as $4.52\ast {10^{ - 5}}RIU$. Importantly, the Q and FOM of the optimized biosensor reached as high as $Q = 6967.4$ and $FOM = 11050RI{U^{ - 1}}$, respectively. Such Q and FOM were much larger than those of similarly reported biosensors [53,57].

 figure: Fig. 6.

Fig. 6. The shifts in BSW reflectance spectra as a function of the Hb concentrations. The corresponding resonance angles are ${\theta _5} = 69.368\deg ({c_{Hb}} = 10g/dl)$, ${\theta _6} = 69.525\deg$ $({c_{Hb}} = 11g/dl)$, ${\theta _7} = 69.771\deg ({c_{Hb}} = 12g/dl)$, and ${\theta _8} = 70.031\deg ({c_{Hb}} = 13g/dl)$, respectively.

Download Full Size | PDF

4. Conclusions

In this paper, a high-performance Bloch surface wave biosensor was successfully constructed for the detection of Hemoglobin concentration. To this end, a porous silicon-based Kretschmann configuration was first designed to ensure excitation of the Bloch surface wave. The performances of the sensor were then optimized by adjusting the buffer layer parameters based on the impedance matching method. The results revealed an increase in the quality factor and figure of merit of the biosensor with the decrease in thickness and refractive index of the buffer layer. By combining the two optimization methods, the quality factor and figure of merit of the optimized biosensor reached as high as $Q = 6967.4$ and $FOM = 11050RI{U^{ - 1}}$, respectively. Overall, the proposed method looks very promising for the fabrication of Bloch surface wave biosensors for the detection of Hemoglobin concentrations.

Funding

Department of Science and Technology of Liaoning Province (22-BS-294); Scientific Research Fund of Liaoning Provincial Education Department (L2020049); National Natural Science Foundation of China (62105132).

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. H. Altug, S. H. Oh, S. A. Maier, and J. Homola, “Advances and applications of nanophotonic biosensors,” Nat. Nanotechnol. 17(1), 5–16 (2022). [CrossRef]  

2. A. Ngo, P. Gandhi, and W. G. Miller, “Frequency that laboratory tests influence medical decisions,” J. Appl. Lab. Med. 1(4), 410–414 (2017). [CrossRef]  

3. W. Bai, H. Yang, Y. Ma, H. Chen, J. Shin, Y. Liu, Q. Yang, I. Kandela, Z. Liu, S. K. Kang, C. Wei, C. R. Haney, A. Brikha, X. Ge, X. Feng, P. V. Braun, Y. Huang, W. Zhou, and J. A. Rogers, “Flexible transient optical waveguides and surface-wave biosensors constructed from monocrystalline silicon,” Adv. Mater. 30(32), 1801584 (2018). [CrossRef]  

4. M. Ouyang, D. Tu, L. Tong, M. Sarwar, A. Bhimaraj, C. Li, G. L. Cote, and D. Di Carlo, “A review of biosensor technologies for blood biomarkers toward monitoring cardiovascular diseases at the point-of-care,” Biosens. Bioelectron. 171, 112621 (2021). [CrossRef]  

5. H. Heidarzadeh, “Analysis and simulation of a plasmonic biosensor for hemoglobin concentration detection using noble metal nano-particles resonances,” Opt. Commun. 459, 124940 (2020). [CrossRef]  

6. C. C. Chang, “Recent advancements in aptamer-based surface plasmon resonance biosensing strategies,” Biosensors 11(7), 233 (2021). [CrossRef]  

7. H. Yan, Y. Zou, S. Chakravarty, C. J. Yang, Z. Wang, N. Tang, D. Fan, and R. T. Chen, “Silicon on-chip bandpass filters for the multiplexing of high sensitivity photonic crystal microcavity biosensors,” Appl. Phys. Lett. 106(12), 121103 (2015). [CrossRef]  

8. K. B. Be, J. Grabska, and C. W. Huck, “Biomolecular and bioanalytical applications of infrared spectroscopy – A review,” Anal. Chim. Acta 1133, 150–177 (2020). [CrossRef]  

9. D. T. Nurrohman and N. F. Chiu, “A Review of Graphene-Based Surface Plasmon Resonance and Surface-Enhanced Raman Scattering Biosensors: Current Status and Future Prospects,” Nanomaterials 11(1), 216 (2021). [CrossRef]  

10. Y. Zhao, R. J. Tong, F. Xia, and Y. Peng, “Current status of optical fiber biosensor based on surface plasmon resonance,” Biosens. Bioelectron. 142, 111505 (2019). [CrossRef]  

11. A. H. Aly, S. Awasthi, M. Mohaseb, Z. Matar, and A. Amin, “MATLAB simulation-based theoretical study for detection of a wide range of pathogens using 1D defective photonic structure,” Crystals 12(2), 220 (2022). [CrossRef]  

12. A. H. Aly, W. Sabra, and H. A. Elsayed, “Cutoff frequency in metamaterials photonic crystals within Terahertz frequencies,” Int. J. Mod. Phys. B 31(15), 1750123 (2017). [CrossRef]  

13. Y. Guo, J. Y. Ye, C. Divin, B. Huang, T. P. Thomas, J. Baker, R. James, and T. B. Norris, “Real-time biomolecular binding detection using a sensitive photonic crystal biosensor,” Anal. Chem. 82(12), 5211–5218 (2010). [CrossRef]  

14. H. Kaviani and J. Barvestani, “Photonic crystal based biosensor with the irregular defect for detection of blood plasma,” Appl. Surf. Sci. 599(15), 153743 (2022). [CrossRef]  

15. 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]  

16. 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]  

17. A. H. Aly, D. Mohamed, Z. Zaky, Z. Matar, N. Abd El-Gawaad, A. Shalaby, F. Tayeboun, and M. Mohaseb, “Novel biosensor detection of tuberculosis based on photonic band gap materials,” Mater. Res. 24(3), e20200483 (2021). [CrossRef]  

18. A. H. Aly, S.-W. Ryu, H.-T. Hsu, and C.-J. Wu, “THz transmittance in one-dimensional superconducting nanomaterial-dielectric superlattice,” Mater. Chem. Phys. 113(1), 382–384 (2009). [CrossRef]  

19. A. H. Aly, A. A. Ameen, M. Mahmoud, Z. Matar, M. Al-Dossari, and H. A. Elsayed, “Photonic crystal enhanced by metamaterial for measuring electric permittivity in GHz range,” Photonics 8(10), 416 (2021). [CrossRef]  

20. A. H. Aly, D. Mohamed, and M. Mohaseb, “Metamaterial control of hybrid multifunctional High-Tc superconducting photonic crystals for 1D Quasi-periodic structure potential applications,” Mater. Res. 23(3), e20190695 (2020). [CrossRef]  

21. A. H. Aly and D. Mohamed, “The optical properties of metamaterial-superconductor photonic band gap with/without defect layer,” J. Supercond. Novel Magn. 32(7), 1897–1902 (2019). [CrossRef]  

22. K. Chakrabarti, S. Mostufa, and A. K. Paul, “Design and analysis of a position chirped metamaterial photonic crystal array for confinement of light pulses,” J. Opt. 23(11), 115201 (2021). [CrossRef]  

23. D. Yang, S. Kita, F. Liang, C. Wang, H. Tian, Y. Ji, M. Lončar, and Q. Quan, “High sensitivity and high Q-factor nanoslotted parallel quadrabeam photonic crystal cavity for real-time and label-free sensing,” Appl. Phys. Lett. 105(6), 063118 (2014). [CrossRef]  

24. D. Threm, Y. Nazirizadeh, and M. Gerken, “Photonic crystal biosensors towards on-chip integration,” J. Biophotonics 5(8-9), 601–616 (2012). [CrossRef]  

25. H. Qiao, A. H. Soeriyadi, B. Guan, P. J. Reece, and J. J. Gooding, “The analytical performance of a porous silicon Bloch surface wave biosensors as protease biosensor,” Sens. Actuators, B 211, 469–475 (2015). [CrossRef]  

26. M. Ge, X. Fang, J. Rong, and C. Zhou, “Review of porous silicon preparation and its application for lithium-ion battery anodes,” Nanotechnology 24(42), 422001 (2013). [CrossRef]  

27. A. Jane, R. Dronov, A. Hodges, and N. H. Voelcker, “Porous silicon biosensors on the advance,” Trends Biotechnol. 27(4), 230–239 (2009). [CrossRef]  

28. L. Bai, Y. Gao, J. Wang, T. Aili, Z. Jia, X. Lv, X. Huang, and J. Yang, “Detection of β-lactoglobulin by a porous silicon microcavity biosensor based on the angle spectrum,” Sensors 22(5), 1912 (2022). [CrossRef]  

29. S. Khanna, P. Marathey, R. Patel, S. Paneliya, R. Chaudhari, J. Vora, A. Ray, R. Banerjee, and I. Mukhopadhyay, “Unravelling camphor mediated synthesis of TiO2 nanorods over shape memory alloy for efficient energy harvesting,” Appl. Surf. Sci. 541, 148489 (2021). [CrossRef]  

30. V. S.-Y. Lin, K. Motesharei, K.-P. S. Dancil, M. J. Sailor, and M. R. Ghadiri, “A porous silicon-based optical interferometric biosensor,” Science 278(5339), 840–843 (1997). [CrossRef]  

31. S. Sahu, J. Ali, P. P. Yupapin, and G. Singh, “Porous silicon based Bragg-grating resonator for refractive index biosensor,” Photonic Sens. 8(3), 248–254 (2018). [CrossRef]  

32. C. Solanki, R. Bilyalov, J. Poortmans, J.-P. Celis, J. Nijs, and R. Mertens, “Self-standing porous silicon films by one-step anodizing,” J. Electrochem. Soc. 151(5), C307–C314 (2004). [CrossRef]  

33. A. D. R. Pérez, M. de la Mora, J. Benítez, R. Castañeda-Guzmán, J. A. Reyes-Esqueda, and M. Villagrán-Muniz, “In situ study of porous silicon thin films thermal oxidation by pulsed laser photoacoustics,” Semicond. Sci. Technol. 33(8), 085001 (2018). [CrossRef]  

34. I. Suárez, V. Chirvony, D. Hill, and J. Martínez-Pastor, “Simulation of surface-modified porous silicon photonic crystals for biosensing applications,” Photonics Nanostructures: Fundam. Appl. 10(3), 304–311 (2012). [CrossRef]  

35. A. H. Soeriyadi, B. Gupta, P. J. Reece, and J. J. Gooding, “Optimising the enzyme response of a porous silicon photonic crystal via the modular design of enzyme sensitive polymers,” Polym. Chem. 5(7), 2333–2341 (2014). [CrossRef]  

36. J. Shi, A. Rezk, C. Ma, L. Zhang, P. Yang, D. Ge, and S. Zhu, “Excitation of Bloch surface wave using silver nanoparticles for sensitivity enhanced biosensor,” Mater. Res. Express 6(9), 095042 (2019). [CrossRef]  

37. M. Al-Dossari, S. Awasthi, A. Mohamed, N. Abd El-Gawaad, W. Sabra, and A. H. Aly, “Bio-alcohol sensor based on one-dimensional photonic crystals for detection of organic materials in wastewater,” Materials 15(11), 4012 (2022). [CrossRef]  

38. I. S. Amiri, B. K. Paul, K. Ahmed, A. H. Aly, R. Zakaria, P. Yupapin, and D. Vigneswaran, “Tri-core photonic crystal fiber based refractive index dual sensor for salinity and temperature detection,” Microw. Opt. Technol. Lett. 61(3), 847–852 (2019). [CrossRef]  

39. A. Natesan, K. P. Govindasamy, T. R. Gopal, V. Dhasarathan, and A. H. Aly, “Tricore photonic crystal fibre based refractive index sensor for glucose detection,” IET Optoelectron. 13(3), 118–123 (2019). [CrossRef]  

40. A. H. Aly, S. S. A. Ghany, B. Kamal, and D. Vigneswaran, “Theoretical studies of hybrid multifunctional YaBa2Cu3O7 photonic crystals within visible and infra-red regions,” Ceram. Int. 46(1), 365–369 (2020). [CrossRef]  

41. M. Awad and A. H. Aly, “Experimental and theoretical studies of hybrid multifunctional TiO2/TiN/TiO2,” Ceram. Int. 45(15), 19036–19043 (2019). [CrossRef]  

42. A. H. Aly and D. Mohamed, “BSCCO/SrTiO3 one dimensional superconducting photonic crystal for many applications,” J. Supercond. Novel Magn. 28(6), 1699–1703 (2015). [CrossRef]  

43. Z. A. Zaky, A. Panda, P. D. Pukhrambam, and A. H. Aly, “The impact of magnetized cold plasma and its various properties in sensing applications,” Sci. Rep. 12(1), 3754 (2022). [CrossRef]  

44. A. H. Aly, S. Awasthi, A. M. Mohamed, M. Al-Dossari, Z. Matar, M. Mohaseb, N. Abd El-Gawaad, and A. Amin, “1D reconfigurable bistable photonic device composed of phase change material for detection of reproductive female hormones,” Phys. Scr. 96(12), 125533 (2021). [CrossRef]  

45. O. Zhernovaya, O. Sydoruk, V. Tuchin, and A. Douplik, “The refractive index of human hemoglobin in the visible range,” Phys. Med. Biol. 56(13), 4013–4021 (2011). [CrossRef]  

46. E. Beutler and J. Waalen, “The definition of anemia: what is the lower limit of normal of the blood hemoglobin concentration?” Blood 107(5), 1747–1750 (2006). [CrossRef]  

47. A. Keshavarz, S. Zangenehzadeh, and A. Hatef, “Optimization of surface plasmon resonance-based biosensors for monitoring hemoglobin levels in human blood,” Appl. Nanosci. 10(5), 1465–1474 (2020). [CrossRef]  

48. A. K. Goyal and S. Pal, “Design analysis of Bloch surface wave based sensor for haemoglobin concentration measurement,” Appl. Nanosci. 10(9), 3639–3647 (2020). [CrossRef]  

49. V. Paeder, V. Musi, L. Hvozdara, S. Herminjard, and H. Herzig, “Detection of protein aggregation with a Bloch surface wave based sensor,” Sens. Actuators, B 157(1), 260–264 (2011). [CrossRef]  

50. G. A. Rodriguez, J. D. Ryckman, Y. Jiao, and S. M. Weiss, “A size selective porous silicon grating-coupled Bloch surface and sub-surface wave biosensor,” Biosens. Bioelectron. 53, 486–493 (2014). [CrossRef]  

51. J. Ma, X.-B. Kang, and Z.-G. Wang, “Sensing performance optimization of the Bloch surface wave biosensor based on the Bloch impedance-matching method,” Opt. Lett. 43(21), 5375–5378 (2018). [CrossRef]  

52. X. Kang, W. Tan, Z. Wang, and H. Chen, “Optic Tamm states: The Bloch-wave-expansion method,” Phys. Rev. A: At., Mol., Opt. Phys. 79(4), 043832 (2009). [CrossRef]  

53. Y. Li, T. Yang, Z. Pang, G. Du, S. Song, and S. Han, “Phase-sensitive Bloch surface wave sensor based on variable angle spectroscopic ellipsometry,” Opt. Express 22(18), 21403–21410 (2014). [CrossRef]  

54. Z. Y. Li and L. L. Lin, “Photonic band structures solved by a plane-wave-based transfer-matrix method,” Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 67(4), 046607 (2003). [CrossRef]  

55. C. C. Katsidis and D. I. Siapkas, “General transfer-matrix method for optical multilayer systems with coherent, partially coherent, and incoherent interference,” Appl. Opt. 41(19), 3978–3987 (2002). [CrossRef]  

56. M. Friebel and M. Meinke, “Model function to calculate the refractive index of native hemoglobin in the wavelength range of 250-1100 nm dependent on concentration,” Appl. Opt. 45(12), 2838–2842 (2006). [CrossRef]  

57. A. Sinibaldi, R. Rizzo, G. Figliozzi, E. Descrovi, N. Danz, P. Munzert, A. Anopchenko, and F. Michelotti, “A full ellipsometric approach to optical sensing with Bloch surface waves on photonic crystals,” Opt. Express 21(20), 23331–23344 (2013). [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 (6)

Fig. 1.
Fig. 1. Schematic diagram of the BSW biosensor.
Fig. 2.
Fig. 2. (a) The variation in impedance mismatch function with the incident angle (red dashed line) and the reflection spectrum (black solid line) at incident wavelength $\lambda = 1060nm$. (b) The electric field distribution inside the biosensor at $\theta \textrm{ = }72.311\deg$ and $\lambda = 1060nm$.
Fig. 3.
Fig. 3. The shifts in BSW reflectance spectra as a function of Hb concentration. The corresponding resonance angles were determined as ${\theta _1} = 72.320\deg ({c_{Hb}} = 10g/dl)$, ${\theta _2} = 72.414\deg ({c_{Hb}} = 11g/dl)$, ${\theta _3} = 72.578\deg ({c_{Hb}} = 12g/dl)$ and ${\theta _4} = 72.736\deg ({c_{Hb}} = 13g/dl)$, respectively.
Fig. 4.
Fig. 4. The effect of the buffer layer thickness on the biosensor performance.
Fig. 5.
Fig. 5. The effect of the buffer layer refractive index on the biosensor performance.
Fig. 6.
Fig. 6. The shifts in BSW reflectance spectra as a function of the Hb concentrations. The corresponding resonance angles are ${\theta _5} = 69.368\deg ({c_{Hb}} = 10g/dl)$, ${\theta _6} = 69.525\deg$ $({c_{Hb}} = 11g/dl)$, ${\theta _7} = 69.771\deg ({c_{Hb}} = 12g/dl)$, and ${\theta _8} = 70.031\deg ({c_{Hb}} = 13g/dl)$, respectively.

Equations (12)

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

( 1 ρ ) n s i 2 n e f f 2 n s i 2 + 2 n e f f 2 + ρ n 0 2 n e f f 2 n 0 2 + 2 n e f f 2 = 0
Z S + = ξ C ( ξ H b + ξ C ) ξ C e 2 i k C d C ( ξ C ξ H b ) ( ξ H b + ξ C ) e 2 i k C d C ( ξ C ξ H b )
Z S = ξ + B S W ( ξ p r i ξ B S W ) + ξ B S W e 2 i m K Λ ( ξ + B S W ξ p r i ) ( ξ p r i ξ B S W ) + e 2 i m K Λ ( ξ + B S W ξ p r i )
ξ ± B S W = 2 ξ A ξ B sin ( K Λ ) i ( ξ A 2 ξ B 2 ) sin ( k A d A ) sin ( k B d B ) ( ξ A ξ B ) sin ( k A d A k B d B ) + ( ξ A + ξ B ) sin ( k A d A + k B d B )
K  =  1 Λ arccos [ cos ( k A d A ) cos ( k B d B ) 1 2 ( k A k B + k B k A ) sin ( k A d A ) sin ( k B d B ) ]
Z S = | Z S + Z S |
Z S = Z S n H b Δ n H b + Z S λ Δ λ + Z S θ Δ θ
S θ = | Δ θ Δ n H b | = | Z S / n H Z S / θ |
Q = θ F W H M
F O M = S F W H M
L O D = Δ n H b Δ θ σ
n H b ( λ , c H b ) = n H 2 O ( λ ) [ α ( λ ) c H b + 1 ]
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.