Open access peer-reviewed chapter

Biological Sensing Using Infrared SPR Devices Based on ZnO

Written By

Hiroaki Matsui

Submitted: 03 February 2022 Reviewed: 18 March 2022 Published: 10 June 2022

DOI: 10.5772/intechopen.104562

From the Edited Volume

Biosignal Processing

Edited by Vahid Asadpour and Selcan Karakuş

Chapter metrics overview

82 Chapter Downloads

View Full Metrics

Abstract

Biological detection based on surface plasmon resonances (SPRs) on metallic Ga-doped zinc oxide (ZnO: Ga) film surfaces is introduced as one of the interesting functionalities of ZnO. SPRs on ZnO: Ga films (ZnO-SPRs) have attracted much attention as alternative plasmonic materials in the infrared (IR) range. This chapter focuses on the structure and optical properties of ZnO-SPR with different layer structure from experimental and theoretical approaches. First, the plasmonic properties of single ZnO: Ga films excited by Kretschmann-type SPRs were investigated. Second, an insulator–metal–insulator structure with a ZnO: Ga film applied as a metal layer is introduced. Finally, hybrid layer structures with the capping of thin dielectric layers to ZnO-SPR (dielectric-assisted ZnO-SPR) were fabricated to enhance SPR properties in the IR range. The biological sensing on ZnO-SPR is experimentally demonstrated by measuring biological interactions. This work provides new insights for fabricating biological sensing platforms on ZnO materials.

Keywords

  • oxide semiconductor
  • surface plasmon
  • near-infrared
  • biological detection
  • zinc oxide
  • heterointerface

1. Introduction

The impurity dopants in zinc oxide (ZnO) produce interesting optical, electrical and magnetic functionalities. For example, photoluminescence behaviors are observed by doping with rare-earth and transition-metal atoms [1, 2, 3, 4]. Ferromagnetic properties are found by the incorporation of transition-metal atoms into the host [5, 6, 7]. Above all, the donor dopants, such as Ga and Al ions, lead to the metallic conductivity of ZnO which has been widely used as transparent electrodes [8, 9]. The metallic properties of ZnO are also prospective candidates for alternatively plasmonic materials in the infrared (IR) range [10, 11, 12, 13]. Unlike classical plasmonic materials such as Ag and Au metals, plasmonic resonances can be controlled by changing electron density in ZnO. Additionally, the electronic structure of ZnO is composed of 4 s and 2p orbitals, providing no inter-band transition such as those shown by Ag and Au metals [14]. The inter-band transition of ZnO only shows in the ultra-violet range. This band system produces a low optical loss in the IR range. Thus far, different geometries such as dots, wires, and films have been chosen to study surface plasmon resonances (SPRs) [15, 16, 17]. These SPR-related studies have shown the unprecedented capabilities of ZnO for use as alternatives to metals in IR applications.

It is known that ZnO is one of the interesting materials in biological sensing.

Optical and electrical techniques detect biomolecular interactions on ZnO film surfaces. The piezoelectric responses of ZnO have been applied to biosensors based on surface acoustic waves [18]. Electrochemical impedance is needed to use transparent electrodes based on ZnO [19]. These interesting properties of ZnO as biocompatible materials have generated much attention as solid oxide substrates for highly sensitive biosensing platforms. Our group has investigated layer samples of ZnO to improve the optical technology associated with SPRs. In particular, layer samples could readily be employed in biological sensing based on propagation-type SPRs. The benefits of layer samples with large surface areas render film platforms more attractive for industrial applications. Recently, SP waves, which are excited by traverse magnetic polarization, have been produced in ZnO-based optical fibers as biochemical sensing probes [20, 21]. Surface-enhanced infrared absorptions have also been confirmed on ZnO film surfaces using a prism-based SPR method [22, 23]. These previous studies indicate the potential of ZnO-based SPR (ZnO-SPR) biosensors and have motivated the study of biological sensing.

In this chapter, we discuss the structural and optical properties of ZnO-SPR from experimental and theoretical approaches. We first outline single layers’ SPR properties based on ZnO: Ga, which is referred to as a single sample. The single sample is excited by Kretschmann-type SPRs using attenuated total reflection (ATR) optics. Second, the utilization of an insulator–metal–insulator (IMI) structure to a ZnO-SPR (IMI sample) is introduced in relation to the long-range SP mode. Finally, we fabricate hybrid layer structures with the capping of thin dielectric layers to ZnO-SPR (dielectric-assisted ZnO-SPR; hybrid sample) to enhance SPR properties in the IR range. Schematic pictures of single, IMI and hybrid samples are shown in Figure 1. Each sample’s detection sensitivity is discussed from the viewpoint of electric-field distribution (E-field), propagation and penetration depths of SP waves. Finally, biological sensing on ZnO-SPR is experimentally demonstrated by measuring biological interactions between biotin and streptavidin. This study introduces SPR devices based on ZnO for biological sensing.

Figure 1.

Schematic figures of SPR sensing platforms of (a) single, (b) IMI and (c) hybrid samples.

Advertisement

2. SPRs properties of single ZnO layer surfaces

This section deals with the correlation between SPR phenomena and evanescent fields on metallic ZnO: Ga film surfaces. The fabrication of ZnO: Ga film is introduced as follows [13]. ZnO: Ga films with 5% Ga content were grown on BK-7 glass substrates at 260°C using the pulsed laser ablation (PLD) method. ArF excimer laser pulses (193 nm, 3 Hz, and 1 J/cm2) were irradiated on a ZnO: Ga ceramic target in an O2 atmosphere of 10−4 Pa. The film thickness (t) was controlled within 41 to 180 nm range. A BK-7 glass substrate with a refractive index of 1.517 was applied using an ATR system with a BK-7 optical prism. SPR reflectance (Rp/Rs) was measured using a Kretschmann-type ATR system connected to a Fourier-transform NIR spectrometer, where Rp and Rs indicate the p- and s-polarized reflection lights, respectively. The spectral intensity of SPR reflectance was acquired in the wavenumber range of 7000–4000 cm−1. For theoretical calculations, SPR reflectance spectra and their electric-field (E-field) depths were simulated by Fresnel relations of an N-multilayer model to calculate reflection coefficients under s- and p-polarized configurations [13, 24].

Figure 2(a) shows SPR reflectance spectra of a 174 nm-thick film as a function of incident angle (θ) from 62o to 72o in 2o increments. The peak position gradually shifted to higher wavenumbers with increasing incident angle, also confirmed by calculated SPR reflectance spectra (Figure 2(b)). The SPR behavior was not observed in the film samples with small thicknesses below 100 nm. The screened bulk plasmon resonances appeared in place of SPRs. Figure 2(c) shows a dispersion curve of a 164 nm-thick film sample, revealing consistency between the experimental (black circles) and calculated (black line) data, which were derived from the single SPR mode at the water-ZnO: Ga film interface. The result of E-field depth also evidenced this result. Figure 2(d) shows a depth profile of a mean-square evanescent field <Ezz2 > at 4500 cm−1 of the p-polarized component along the z-direction. Here, the x and y directions are parallel to the film, while the z direction is normal to the film. The film sample had a < Ezz2 > value of 7.3 with a penetration depth (δW) of 200 nm into the water medium, where δW was determined at the depth at which the field decays by a factor of 1/e. The ZnO: Ga film provided a narrow field depth in the IR range, which was essentially different from Au film-based SPRs [25].

Figure 2.

(a) Experimental and (b) calculated SPR reflectance spectra of a 174 nm-thick ZnO: Ga film. Water was selected as the dielectric medium. The incident angle of light (θ) was changed from 62o to 72o. (c) Experimental and calculated dispersion curves of SPRs. A dotted line indicates a light line in a water medium. (d) a depth-dependent mean square evanescent field at calculated at an SPR peak position of θ = 72.

The SPR performance was evaluated by examining the bulk sensitivity using a change in the refractive index of glucose-water mixed solution. Figure 3(a) shows the SPR reflectance spectra at θ = 75o for a 107 nm-thick ZnO: Ga film using different glucose contents (0, 1 and 5 g/dL) [24]. The peak position systematically shifted to lower wavenumbers with increasing glucose content in water, showing a peak shift (Δv) of 22 cm−1 at 1 g/dL. Furthermore, we experimentally and theoretically evaluated the sensitivity (Sexp and Scal) at a wavenumber of 4500 cm−1. The detection sensitivity was expressed by the following expression:

Figure 3.

(a) SPR reflectance spectra of the single film measured with varying glucose content in water from 0, 1 and 5 g/dL. (b) Correlation between peak shift (Δv) and glucose content in water for the single film. The change in refractive index (n) of a mixed solution consisting of glucose and water is also described in the upper horizontal axis. Black line and dots indicate calculated and experiment data (Figure 4(b) and (c) of [24]). Copyright by the American Institute of Physics.

S=Δv/ΔnE1

where Δn indicates the per unit change in the refractive index. The experimental Sexp was determined to be 8300 cm−1/RIU (RIU: refractive index unit), which was similar to the calculated value (8600 cm−1/RIU) (Figure 3(b)).

This section introduced the SPR responses of single ZnO: Ga film surfaces in the IR range. We could clearly observe the SPR spectra and their field distributions. The bulk sensitivity was evaluated using the glucose-water solution to estimate sensing performance. However, the bulk sensitivity requires improvement for real-time monitoring of biological interactions.

Advertisement

3. IMI structures for sensing performance

The evanescent depth of SPR plays an essential role in determining the sensing performance of plasmonic devices [26]. Above all, an IMI structure can control the evanescent field by changing the thickness of the metal layer [27]. This IMI structure is applied to slightly asymmetric dielectric media [28]. The slightly asymmetric IMI structure can provide benefits for aqueous sensing solutions. To date, no attention has been paid to the SPR property based on a ZnO-related IMI structure. The employment of IMI geometry to SPR excitations on ZnO: Ga films remains unclear. In this section, we introduce asymmetric IMI structures consisting of water, ZnO: Ga, and a cytop polymer. In particular, water-based IMI structures are one of the interesting sensing platforms in SPR applications. The benefits of SPR are confirmed from the bulk sensitivity based on index changes. We discuss the sensing performance of ZnO-SPR based on the proof-of-concept of an IMI structure.

The fabrication of IMI samples was performed as follows. The refractive index of a cytop polymer is kept close to water, which can excite SPRs using IMI structures with ZnO: Ga films. A Cytop polymer (perfluropolymer) film (1.8 mm-thickness) was deposited on BK-7 glass substrates using a polymer content of 9% in fluoride solvent by a spin coating method (2500 rpm for 50s). The coated polymer films were annealed at 220°C for 2 h in the air to evaporate the solvent. ZnO: Ga films were fabricated on polymer/glass substrates using PLD at room temperature. ArF excimer laser pulses (193 nm, 5 Hz, and 1 J/cm2) were focused on ZnO: Ga targets in an O2 flow of 10−4 Pa [29].

The IMI sample SPR reflectance spectra are shown in Figure 4(a). The peak positions gradually moved within the 4000 to 6000 cm−1 range with an increasing incident angle from 60.5o to 64o in 2o increments [29]. The SPR spectra were observed even at a film thickness of 22 nm. The SPR peak dependence on the incident angle of light for the IMI sample was higher than that for the single ZnO film. In addition, the IMI sample showed narrower spectral features than those of the single ZnO films. These behaviors were also confirmed by theoretical SPR spectra (Figure 4(b)) [29].

Figure 4.

(a) Experimental and (b) calculated SPR reflectance spectra of the ZnO IMI sample with a film thickness (t) of 22 nm. Water was selected as the dielectric medium. (c) Experimental and calculated dispersion curves of the ZnO IMI sample (t = 22 nm). A dot line indicates light line in water medium. Long-range and short-range SP modes are represented in the figure. (d) a depth-dependent mean square evanescent field at calculated at 4500 cm−1 for the IMI sample (Figures 4(d) and 5(c) of [29]). Copyright by the American Institute of Physics.

The SP mode of an IMI structure is separated into two types of plasmon branches comprising the short-range and long-range SP modes. The dispersion curves of both SP modes can be described using the Maxwell relations in a planar structure [30]:

ε2vε1v+γ2vγ1v.ε0vε1v+γ0vγ1v.e2γ1ωt=ε2vε1v-γ2vγ1v.ε0vε1v+γ0vγ1vE2
γi2=k2-òivc2 (i = 0, 1, 2)E3

where εi(v) (i = 0, 1, 2) represents the dielectric function in each layer, i.e., water, ZnO: Ga, and cytop polymer layers, respectively. Figure 4(c) shows a dispersion curve of the IMI sample, revealing two plasmon branches of short-range and long-range SP modes. The experimental data were similar for the long-range SP mode, which was due to the phase-matching of wave vectors of SPRs at the water–ZnO: Ga–cytop polymer interfaces. An SP wave of the long-range mode is expected to show a longer propagation distance than that of the single-mode. The following equation can express the propagation distance (Lprop): Lprop = 1/2 × Im[kx] [31], which represents the length from the launch point where the evanescent field power decays by a factor of 1/e. The Lprop value of the IMI sample was approximately 10 μm, which was longer than that of the single film (Lprop = 3 μm). The difference in SPR response between the IMI sample and single films was related to the propagation distance.

Figure 4(d) shows a depth profile of a mean-square evanescent field at 4500 cm−1 of the p-polarized component [29]. The value of <Ezz2 > was estimated to be 18.0, which was higher than that of the single film. Besides, the δw value in the water medium was in the vicinity of 1 μm. The penetration depth clearly expanded when using the IMI sample, relating to the long propagation distance and the high angle dependence of SPR reflectance spectra.

Figure 5(a) shows the SPR reflectance spectra (θ = 58.25o) of the IMI sample, revealing a large Δv of 82 cm−1 at 1 g/dL, which was higher than that of the single film [29]. Additionally, the sensitivity was measured at 4500 cm−1, which resulted in a large Sexp value of 30,530 cm−1/RIU. This value was close to the theoretical estimation (Scal = 33,000 cm−1/RIU). The enhanced sensitivity was attributed to the evanescent field depth and longer propagation distance of the SPR, which was compared with the single film (Figure 5(b)) [29].

Figure 5.

(a) SPR reflectance spectra of the IMI sample measured with varying glucose content in water from 0, 1, and 5 g/dL. (b) Correlation between peak shift (Δv) and glucose content in water for the IMI sample. The change in refractive index (n) of a mixed solution consisting of glucose and water is also described in the upper horizontal axis. Black line and dots indicate calculated and experiment data (Figure 4(b) and (c) of [29]). Copyright by the American Institute of Physics.

In this section, we investigated the SPR properties and sensing performance of the IMI sample. The SPR spectra of the IMI sample displayed narrower features than those of the single film. This result provided the extended evanescent field depth and the long propagation distance. Consequently, the sensitivity of the IMI sample was markedly enhanced compared to that of the single film.

Advertisement

4. Dielectric-assisted ZnO-SPR

In Sections 2 and 3, we introduced the SPR properties of the single film and IMI sample. The single film showed broad SPR resonances and weak evanescent fields. The use of an IMI structure increased the evanescent field because of the long-range SP mode. However, the IMI sample produced a large sensing volume on the sample surface due to the long penetration depth, where it is not easy to detect small changes in the refractive index near the sample surfaces. Therefore, there is a need to find new strategies for circumventing the structural limitations of ZnO-SPRs in terms of penetration depth and evanescent field. Consequently, we propose a new structural concept based on dielectric-assisted SPRs to overcome some difficulties associated with ZnO-SPRs for real-time monitoring of biological interactions.

When sufficiently thick dielectric layers are placed on Ag and Au-metal layer surfaces, waveguide modes are supported in addition to the conventional SP mode [32, 33]. However, the sensing performance of a waveguide is comparable to that of a classical SPR. However, the Ag and Au-based SPR devices with thin dielectric layers (e.g., Al2O3 and SiO2) show SP waves along the thin dielectric layer surfaces with a thickness insufficient to support a waveguide mode. The SP waves of dielectric-assisted SPRs sufficiently penetrate the analyte region, unlike the waveguide mode. Therefore, the introduction of thin dielectric layers on top of Ag and Au metallic films has been reported to improve the sensing performance in the visible range [34].

In this section, we report on the capping of thin dielectric layers to a ZnO-SPR sensing platform to enhance the detection sensitivity in the IR range. Here, we define dielectric-assisted ZnO-SPR devices as “hybrid samples”. The insertion of dielectric layers changes the E-field distribution and penetration depth, providing enhanced detection sensitivity. The improvement in sensing performance is discussed from the viewpoint of penetration depth and E-field of the SP waves. Finally, we introduce the sensing capabilities of the hybrid samples by measuring the biological interactions between biotin and streptavidin.

The fabrication of hybrid samples was conducted as follows. Cytop polymer films (2.2 μm thickness) were deposited on BK-7 glass substrates. Metallic ZnO: Ga films with a thickness of 22 nm were fabricated on the polymer-coated substrates using a PLD method at room temperature (RT). Fabrication conditions of the polymer and ZnO: Ga films were the same as those of the IMI sample. In this study, we selected Ga2O3 as a dielectric layer. 190 m-thick Ga2O3 dielectric layers were deposited over ZnO: Ga layers on the polymer/glass substrates at RT in an O2 flow of 10−4 Pa using the PLD method [35].

Figure 6(a) shows a hybrid sample’s experimental SPR reflectance spectra with a Ga2O3 thickness of 190 nm [35]. The peak positions moved within the 4000–5500 cm−1range as the incident angle increased in 2.5o increments. The dependence of the SPR peak on the incident angle of light for the hybrid sample was higher than that for the IMI sample. The hybrid sample exhibited narrower spectral features than those of the IMI sample. These SPR behaviors were further reproduced by theoretical SPR spectra (Figure 6(b)) [35]. The difference in SPR response between the hybrid and IMI samples is attributed to the propagation length. An experimental Lprop of approximately 14 mm was obtained for the hybrid sample, which was greater than that of the IMI sample (Lprop = 10 μm). Figure 6(c) shows the dispersion curves of the hybrid and IMI samples [35]. The dispersion curve of the IMI sample was close to the light line in free space due to the long-range SP mode. However, the dispersion curve of the hybrid sample shifted to higher SP wave vectors by insertion of a Ga2O3 layer on top of a ZnO: Ga layer surface, leading to changes in the optical properties of the SPR reflectance spectra. Besides, the use of Ga2O3 dielectric layer was effective in reducing the penetration depth in the water medium. Figure 6(d) shows a depth profile of a mean-square evanescent field at 4500 cm−1 of the p-polarized component [35]. The value of <Ezz2 > was increased up to 34.5, which was higher than that of the IMI sample. Additionally, the δw value in the water medium was suppressed to 400 nm. The reduced penetration depth was related to an increase in SP wave vector by inserting the dielectric layer to the IMI sample.

Figure 6.

(a) Experimental and (b) calculated SPR reflectance spectra of the hybrid sample with a Ga2O3 thickness of 190 nm. Water was selected as the dielectric medium. (c) Experimental and calculated dispersion curves of the IMI sample and hybrid sample. Black open and closed circles indicate experimental data. Dotted and straight lines represent calculation data. (d) a depth-dependent mean square evanescent field at calculated at 4500 cm−1 for the hybrid sample (Figures 4(c), 6(d) and (h) of [35]). Copyright by the American Chemical Society.

Figure 7(a) shows the SPR reflectance spectra of the hybrid sample [35]. The dip peaks shifted to lower wavenumbers with increasing glucose concentration. A wavenumber shift of 19 cm−1 at 1 g/dL was observed from the SPR reflectance spectra taken at θ = 63o. The Sexp and Scal of the hybrid sample were 18,700 and 19,200 cm−1/RIU, respectively. These values were smaller than those of the IMI sample [35]. (Figure 7(b)). This was due to the reduced penetration depth of the hybrid sample. Furthermore, there is a need to evaluate the sensitivity (S) and spectral linewidth (FWHM: full-width half-maximum) when considering the performance of a biosensor. The figure-of-merit (FoM) is defined by the following relation:

Figure 7.

(a) SPR reflectance spectra of the hybrid sample measured with varying glucose content in water (0, 1, 1, 2 and 4 g/dL). (b) Correlation between peak shift (Δv) and glucose content in water for the hybrid sample. The change in refractive index (n) of a mixed solution consisting of glucose and water is also described in the upper horizontal axis. Black line and dots indicate calculated and experiment data (Figure 8 of [35]). Copyright by the American Chemical Society.

FoM=S/FWHME4

This relation is used to quantify the general performance of a biosensor [36]. This normalization allows for comparison with other sensing platforms. The hybrid sample provided a higher FoM value (51.0 RIU−1) than did the IMI sample (36.1 RIU−1). The enhanced FoM was attributed to the narrowing of the spectral linewidth resulting from the insertion of the Ga2O3 layer.

Advertisement

5. Real-time monitoring of biological interactions

We evaluated the SPR sensing performance of the hybrid sample using the biological interactions between biotin-PEG-DPPE and streptavidin. PEG and DPPE indicate poly(ethylene glycol) and 1, 2 dipalmitoyl-sn-glycerol-3-phosphatidylethanolamine, respectively. The high binding affinity and irreversible binding of the molecular pair of biotin-streptavidin is a powerful tool to measure changes in the SPR reflectance, which has medical applications such as antigen–antibody reactions and allergic reactions. We conducted surface modifications before the biological experiments using self-assembled monolayer (SAM) formation.

The Ga2O3 layer surface of the hybrid sample was chemically modified using a SAM of n-octadecylphosphonic acid [C18H37PO(OH)2:ODPA] to form a CH3-terminated SAM (CH3-SAM). This CH3-SAM is commonly used to obtain hydrophobic surfaces because of the strong hydrogen bonding acid–base character of the –PO(OH)2 group [37]. The hybrid samples were immersed in ODPA (5 mM in ethanol) at RT for 48 h after O2 plasma irradiation. The surface states and chemical composition of the SAM-coated sample were investigated using X-ray photoemission spectroscopy (XPS). Figure 8(a) shows the typical Ga(3d), P(2p), C(1 s), and O(1 s) peaks for different surface treatments [35]. The P(2p) peak of the ODPA-coated sample was observed at 136 eV, which was not obtained for the ethanol or O2 plasma-treated samples. This result clarified the formation of CH3-SAM on the sample surface. Immersion of the hybrid samples in a toluene solution of ODPA reduced the peak intensities related to Ga(3d) and Ga(3p) assigned to Ga–O within 20 min owing to the adsorption of ODPA (Figure 8(b)) [35]. The ratio of the surface carbon concentration to the phosphorous concentration was 20 ± 1, which was similar to 18, the value expected from the molecular formula of ODPA (Figure 8(c)). These XPS results revealed the formation of ODPA on the hybrid samples [35].

Figure 8.

(a) XPS core-level spectra of Ga(3d), P(2p), C(1 s) and O(1 s)-related peaks following different surface modifications. The XPS spectra were taken after surface treatments of ODPA in ethanol, ethanol only and O2 plasma. (b) Ga(3d)- and C(1 s)-related peak intensities as a function of OPDA treatment time. (c) P(2p)-related peak and intensity ratio of C(1 s) and P(2p) as a function of OPDA treatment time (Figure 10 of [35]). Copyright by the American Chemical Society.

Biological interactions were conducted as follows [35]. The hybrid and IMI samples were washed with ethanol several times and then dried using nitrogen gas. Both samples were first exposed to a phosphate-buffered saline (PBS) solution with pH = 7.4 [PBS (1) process], until a stable SPR signal was obtained. Next, both samples were exposed to a solution of biotin-PEG-DPPE (100 μg/mL in PBS) for 35 min, followed by washing with PBS for 5 min [PBS (2) process]. After the sample surfaces were treated with biotin-PEG-DPPE, bovine serum albumin (BSA) was introduced to confirm non-specific protein adsorption to the sample surfaces [BSA process], followed by another PBS wash [PBS (3) process]. PBS solutions with different streptavidin concentrations (0–10 μg/dl in PBS) were then introduced [Streptavidin process]. The streptavidin concentrations were used to monitor the specific biological bonding of biotin and avidin [38].

Figure 9(a) shows the SPR reflectance spectra measured after each process [35]. The SPR peak position shifted to small wavenumbers after the surfaces were treated with biotin-PEG-DPPE and streptavidin. The SPR peak shift (Δv) was observed to be 14 cm−1. The SPR peak shift was evaluated by monitoring the reflectance difference (ΔR) at 4250 cm−1 induced by biotin-streptavidin binding (Figure 9(b)) [35]. The ΔR values were changed slightly through surface treatment with biotin-PEG-DPPE. The streptavidin binding to biotin-PEG-DPPE remarkably depended on the streptavidin concentration. However, the IMI sample did not exhibit a significant change in ΔR during the biotin-streptavidin binding (Figure 9(c) and (d)) [35]. The time-dependent ΔR of the hybrid and IMI samples was attributed to the biotin-streptavidin binding. The use of the Ga2O3 layer on top of the ZnO: Ga layer surface successfully allowed for monitoring of the biological interactions. This resulted from the higher E-field and shorter field depth of the hybrid sample compared with those of the IMI sample.

Figure 9.

Experimental SPR reflectance spectra of the hybrid sample (a) and the IMI samples (c) in each process in the biological reaction between biotin-PEG-DPPE and streptavidin of a concentration of 5 μg/mL. Detection of streptavidin at different concentrations using biotin-PEG-DPPE of the hybrid (b) and IMI samples (d). Differential reflectance (DR) was monitored at 4250 and 5250 cm−1 for the hybrid and IMI samples, respectively, (Figure 13 of [35]). Copyright by the American Chemical Society.

The present detection limit of the hybrid samples was approximately 1 μg/ml (15 nM) due to the background noises. Industrial applications are expected to detect bio-molecule concentration at the ng/dl (~ pM) level. The spectral features of the present ZnO-SPRs were markedly influenced by the interface roughness, relating to the sensing activity for biomolecular detection. Recently, high-sensitive SPR detection at the pM levels of small biomolecules (e.g. biotin) using nanophotonic devices such as cavity structures and metamaterials with layer structures have been reported [39, 40, 41]. Recently, biological sensing based on ZnO-related SPR and metamaterials are reported by some papers [42, 43, 44]. Therefore, monitoring biological interactions at the pM level could be realized by enhancing the structural and crystalline quality of ZnO-SPR sensors. The detection sensitivity of ZnO-SPR is expected to improve the employment of new nano-plasmonic structures such as cavities and metamaterials.

Advertisement

6. Conclusion

This chapter reported on the structural and optical properties of ZnO-SPR with different layer structures (single, IMI, and hybrid samples) from experimental and theoretical approaches. First, the SPR properties and sensing performance of the single sample were investigated. Second, the IMI sample was introduced in ZnO-SPR devices to enhance sensing activity. Finally, the hybrid sample with the capping of thin dielectric layers to ZnO-SPR was developed to monitor biological interactions between biotin and streptavidin. This work highlights new insights for fabricating biological sensing platforms on ZnO materials.

References

  1. 1. Jadwisienczak WM, Lozykoqski HJ, Xu A, Patel B. Visible emission from ZnO doped with rare-earth ions. Journal of Electronic Materials. 2002;31:776-684. DOI: 10.1007/s11664-002-0235-z
  2. 2. Badalawa W, Matsui H, Osone T, Hasuike N, Harima H, Tabata H. Correlation between structural and luminescent properties in Eu3+-doped ZnO epitaxial layers. Journal of Applied Physics. 2011;109:053502. DOI: 10.1063/1.3549633
  3. 3. Panda J, Sasmal I, Nath TK. Magnetic and optical properties of Mn-doped ZnO vertically aligned nanorods synthesized by hydrothermal technique. AIP Advances. 2016;6:035118. DOI: 10.1063/1.4944837
  4. 4. Babikier M, Wang D, Wang J, Li Q , Sun J, Yan Y, et al. Cu-doped ZnO nanorod arrays: The effects of copper precursor and concentration. Nanoscale Research Letters. 2014;9:199. DOI: 10.1186/1556-276X-9-199
  5. 5. Neal JR, Behan AJ, Ibrahim RM, Blythe HJ, Ziese M, Fox AM, et al. Room-temperature magneto-optics of ferromagnetic transition-metal-doped ZnO thin films. Physical Review Letters. 2006;96:197208. DOI: 10.1103/PhysRevLett.96.197208
  6. 6. Matsui H, Tabata H. Lattice, band and spin engineering on Zn1-xCoxO. Journal of Applied Physics. 2013;113:183525
  7. 7. Ali N, Singh B, Khan ZA, Vijava AR, Tarafder K, Ghosh S. Origin of ferromagnetism in Cu-doped ZnO. Scientific Reports. 2019;9:2461. DOI: 10.1038/s41598-019-39660-x
  8. 8. Jin ZC, Hamberg I, Granqvist CG. Optical properties of sputter-deposited ZnO: Al thin films. Journal of Applied Physics. 1988;64:5117. DOI: 10.1063/1.342419
  9. 9. Yamamoto N, Makino H, Osone N, Ujihara A, Ito T, Hokari H, et al. Structural, electrical and optical properties of highly transparent conductive ZnO films. Thin Solid Films. 2012;520:4131. DOI: 10.3169/itej.66.555
  10. 10. Buonsanti R, Llordes A, Aloni S, Helms BA, Milliron DJ. Tunable infrared absorption and visible transparency of colloidal aluminum-doped zinc oxide nanocrystals. Nano Letters. 2011;11:4706. DOI: 10.1021/nl203030f.nnichsen
  11. 11. Sachet E, Losego MD, Guske J, Franzen S, Maria JP. Mid-infrared surface plasmon resonance in znc oxide semiconductor thin films. Applied Physics Letters. 2013;102:051111. DOI: 10.1063/1.4791700
  12. 12. Naik GJ, Kim J, Boltasseva A. Oxides and nitrides as alternative plasmonic materials in the optical range. Optical Materials Express. 2011;1(6):1090-1099. DOI: 10.1364/OME.1.001090
  13. 13. Badalawa W, Matsui H, Ikehata A, Tabata H. Surface plasmon modes guided by ZnO: Ga layers bounded by different dielectrics. Applied Physics Letters. 2011;99:011913. DOI: 10.1063/1.3608313
  14. 14. Sönnichsen C, Franzl F, von Plessen G, Feldmann J. Plasmon resonances in large noble-metal clusters. New Journal of Physics. 2002;4:93. DOI: 10.1088/1367-2630/4/1/393
  15. 15. Kalusniak S, Sadofev S, Hennberger F. ZnO as a tunable metal: New types of surface plasmon polaritons. Physical Review Letters. 2014;112:137401. DOI: 10.1103/PhysRevLett.112.137401
  16. 16. Kim J, Dutta A, Memarzadeh B, Kildishev AV, Mosallaei H, Boltasseva A. Mint: Zinc oxide based plasmonic multilayer resonator: Localized and gap surface plasmon in the infrared. ACS Photonics. 2015;2:1224. DOI: 10.1021/acsphotonics.5b00318
  17. 17. Nader N, Vangala S, Hendrickson JR, Leedy KD, Look DC, Guo J, et al. Investigation of plasmon resonance tunneling through subwavelength hole arrays conductive ZnO films. Journal of Applied Physics. 2015;118:173106. DOI: 10.1063/1.4934875
  18. 18. Cao X, Cao S, Guo H, Li T, Jie Y, Wang N, et al. Piezotronic effect enhanced label-free detection of DNA using a schottky-contacted ZnO nanowire biosensor. ACS Nano. 2016;10:8038-8044. DOI: 10.1021/acsnano.6b04121
  19. 19. Wang C, Huang N, Zhuang H, Jiang X. Enhanced performance of nanocrystalline ZnO DNA biosensors via introducing electrochemical covalent biolinkers. ACS Applied Materials & Interfaces. 2015;7:7605-7612. DOI: 10.1021/acsmi.5b00040
  20. 20. Wang Y, Dong J, Luo Y, Tang J, Lu H, Yu J, et al. Indium tin oxide coated two-mode fiber for enhanced SPR sensor on near-infrared region. IEEE Photonics Journal. 2017;9:4801309. DOI: 10.1109/JPHOT.2017.2757513
  21. 21. Minn K, Anopchenko A, Yang J, Lee HWH. Excitation of epsilon-near-zero resonance in ultra-thin indium tin oxide shell embedded nanostructured optical fiber. Scientific Reports. 2018;8:2342. DOI: 10.1038/s41598-018-19633-2
  22. 22. Martínez J, Ródenas A, Aguiló M, Fernadez T, Solis J, Diaz F. Mid-infrared surface plasmon resonance polariton chemical sensing on fiber-coupled ITO coated glass. Optics Letters. 2016;41:2493-2496. DOI: 10.1364/OL.41.002493
  23. 23. Kalusniak S, Sadofev S, Hennberger F. Resonant interaction of molecular vibrations and surface plasmon polaritons: The weak coupling regime. Physical Review B. 2014;90:125423. DOI: 10.1103/PhysRevB.90.125423
  24. 24. Matsui H, Ikehata A, Tabata H. Surface plasmon sensors on ZnO: Ga layer surfaces: Electric field distributions and absorption-sensitivity enhancements. Applied Physics Letters. 2015;106:011905. DOI: 10.1063/1.4905211
  25. 25. Ikehata A, Itoh T, Ozaki Y. Surface plasmon resonance near-infrared spectroscopy. Analytical Chemistry. 2004;76:6461-6469. DOI: 10.1021/ac049003a
  26. 26. Yang F, Sambles JR, Bradberry GW. Long-range surface modes supported by thin films. Physical Review B. 1991;44:5855. DOI: 10.1103/PhysRevB.44.5855
  27. 27. Dionne JA, Swealtlock LA, Atwater HA, Polman P. Planar metal plasmon waveguides: Frequency-dependent dispersion, propagation, localization, and loss beyond the free electron model. Physical Review B. 2005;72:075405. DOI: 10.1103/PhysRevB.72.075405
  28. 28. Wendier L, Haupt R. Long-range surface plasmon - polariton in asymmetric layer structures. Journal of Applied Physics. 1986;59:3289. DOI: 10.1063/1.336884
  29. 29. Matsui H, Ikehata A, Tabata H. Asymmetric plasmon structures on ZnO: Ga for high sensitivity in the infrared range. Applied Physics Letters. 2016;109:191601. DOI: 10.1063/1/4966598
  30. 30. Economou EM. Surface plasmons in thin films. Physical Review B. 1969;182:539. DOI: 10.1104/PhysRev.182.539
  31. 31. Dastmalchi B, Tassin P, Koschny T, Soukoulis CN. A new perspective on plasmonics: Confinement and propagaton length of surface plasmons for different materials and geometries. Advanced Optical Materials. 2015;3:177. DOI: 10.1002/adom.201500446
  32. 32. Hayashi S, Nesterenko DV, Rahmouni A, Sekkat Z. Observation of Fano line shapes arising from coupling between surface plasmon polariton and waveguide modes. Applied Physics Letters. 2016;108:051101. DOI: 10.1063/1.4940984
  33. 33. Abbas A, Linman MJ, Cheng Q. Sensitivity comparison of surface plasmon resonance and plasmon-waveguide resonance biosensors. Sensors and Actuators A. 2011;156:169-175. DOI: 10.1016/j.snb.2011.04.008
  34. 34. Lahav A, Shalabaney A, Abdulhalim I. Surface plasmon sensor with enhanced sensitivity using top nano dielectric layer. Journal of Nanophotonics. 2009;3:031501. DOI: 10.1117/1.3079803
  35. 35. Kuranaga Y, Matsui H, Ikehata Am Shimoda Y, Noiri M, Ho YL, Delaunay JJ, et al. Enhancing detection sensitivity of ZnO-based infrared plasmonic sensors using capped dielectric Ga2O3 layers for real-time monitoring of biological interactions. ACS Applied Bio Materials. 2020;3:6331-6342. DOI: 10.1021/acsabm.0c00792
  36. 36. Otte MA, Sepúlveda B, Ni W, Juste JP, Liz-Marzán LM, Lechuga LM. Identification of the optimal spectra region for plasmonic and nanoplasmonic sensing. ACS Nano. 2010;4:349-357. DOI: 10.1021/nn901024e
  37. 37. Li F, Shishkin E, Mastro MA, Uite JK, Eddy CR, Edgar JH, et al. Photopolymerization of self-assembled monolayers of Diacetylentic Alkylphophonic acids on group-III nitride substrates. Langmuir. 2010;26:10725-10730. DOI: 10.1021/la100273q
  38. 38. Seto H, Yamashita C, Kamba S, Kondo T, Hasegawa M, Matsumoto M, et al. Biotinylation of silicon and nickel surfaces and detection of streptavidin as biosensor. Langmuir. 2013;29:9457-9463. DOI: 10.1021/la401068n
  39. 39. Sreekanth KV, Sreejith S, Mishra A, Chen X, Sun H, Lim CT, et al. Biosensing with the singular phase of an ultrathin metal-dielectric nanophotonic cavity. Nature Communications. 2018;9:369. DOI: 10.1038/s41467-018-02860-6
  40. 40. Shinohara S, Tanaka D, Okamoto K, Tamada K. Colorimetric plasmon sensors with multilayered metallic nanoparticle sheets. Physical Chemistry Chemical Physics. 2015;17:186060-118612. DOI: 10.1039/C5CP02564H
  41. 41. Sreekanth KV, Sreejith S, Alapan Y, Sitti M, Lim CT, Singh R. Microfluidics integrated lithography-free nanophotonic biosensors for the detection of small molecules. Advanced Optical Materials. 2019;7:1801313. DOI: 10.1003/adom.201801313
  42. 42. Guo S, Wu X, Li Z, Tong K. High-sensitivity biosensor-based enhanced SPR by ZnO/MoS2 nanowires array layer with graphene oxide nanosheet. International Journal of Optics. 2020;2020:7342737. DOI: 10.1155/2020/342737
  43. 43. Xu H, Song Y, Zhu P, Zhao W, Liu T, Wang Q , et al. Alcohol sensor based on surface plasmon resonance of ZnO nanoflowers/Au structure. Materials. 2022;15:189. DOI: 10.3390/ma15010189
  44. 44. Mei GS, Menon PS, Hegde G. ZnO for performance enhancement of surface plasmon biosensor: A review. Materials Research Express. 2020;7:012003. DOI: 10.1088/2053-1591/ab66a7

Written By

Hiroaki Matsui

Submitted: 03 February 2022 Reviewed: 18 March 2022 Published: 10 June 2022