Acessibilidade / Reportar erro

Device for Streptavidin Detection Using LSPR and Electrochemical Transductions on the same Platform

Abstract

Research in biosensing currently occupies an important role among biomolecular detection techniques, especially with respect to medical diagnoses. Is this sense, the development of biosensors based on the localized surface plasmon resonance (an optical transducer) or electrochemical transductions have been highlighted. However, optical and electrochemical detection working together enable more versatile detections in biosensing. Thus, we propose a dual sensor, which incorporates these two mechanisms on the same platform. The sensor shared the same transducer that was based on a glass-indium thin oxide (ITO) surface covered with gold nanoparticles. The transducer was prepared by the immobilization of gold nanoparticles on (3-aminopropyl)triethoxysilane-activated glass-ITO. The substrate ability to respond to physical surface variations was assessed. For this, measurements of the gold nanoparticles band redshift (plasmonic transduction) and charge transfer resistance from impedance spectra (electrochemical transduction) were performed. The platform was able to detect streptavidin immobilized on the surface by both mechanisms demonstrating its applicability as a device for biodetection in dual detection mode.

Keywords:
localized surface plasmon resonance; electrochemical impedance spectroscopy; dual transduction; gold nanoparticles; biosensor


Introduction

The need for improved medical and environmental analysis encourages new strategies to determine and quantify biospecies. Quick analysis, low cost, and smaller devices have been evidenced in the scientific research area. Biosensors meet these requirements11 Liu, D.; Luo, Q.; Deng, F.; Li, Z.; Li, B.; Shen, Z.; Anal. Chim. Acta 2017, 971, 26. and therefore have been widely used in different areas, such as analysis of food,22 Ashley, J.; Li, S. F. Y.; Biosens. Bioelectron. 2013, 48, 126.,33 López, M. S.-P.; Redondo-Gómez, E.; López-Ruiz, B.; Talanta 2017, 175, 209. in medical science,44 Ertürk, G.; Özen, H.; Tümer, M. A.; Mattiasson, B.; Denizli, A.; Sens. Actuators, B 2016, 224, 823.

5 Wang, B.; Jing, R.; Qi, H.; Gao, Q.; Zhang, C.; J. Electroanal. Chem. 2016, 781, 212.
-66 Tepeli, Y.; Ülkü, A.; Sens. Actuators, B 2018, 254, 377. environmental science,77 Dali, M.; Zinoubi, K.; Chrouda, A.; Abderrahmane, S.; Cherrad, S.; Jaffrezic-Renault, N.; J. Electroanal. Chem. 2018, 813, 9. and others.88 Faridbod, F.; Gupta, V. K.; Zamani, H. A.; Int. J. Electrochem. 2011, 2011, 352546.,99 Zeng, Y.; Zhu, Z.; Du, D.; Lin, Y.; J. Electroanal. Chem. 2016, 781, 147. Basically, a biosensor has two main elements: a bioreceptor that recognizes specifically the target analyte, and a transductor that converts this interaction into a measurable signal.1010 Sassolas, A.; Blum, L. J.; Leca-Bouvier, B. D.; Biotechnol. Adv. 2012, 30, 489.,1111 Singh, A.; Thapa, K. B.; Kumar, N.; Optik 2015, 126, 244. The nature of the transducers may be diverse, such as electrochemistry, optics, piezoelectric or thermal.1212 Mehrotra, P.; J. Oral Biol. Craniofacial Res. 2016, 6, 153. However, electrochemical and optical transductions are usually more exploited in biosensors and some factors contribute to this fact: electrochemical technique is an older, well established and more studied technology and optical transductions stand out compared to other transducers types for their easy handling and minimized false responses.1313 Braga, J. R. G.; Queiroz, A. A. A.; Ramos, A. C. B.; Int. J. Comput. Digital Syst. 2012, 1, 1.,1414 Senay, R. H.; Ulucan, F.; Kuru, C. İ.; Akgöl, S. In Smart Biosensing Technologies: Past, Present and Future; Gürsan, A. E.; Sezgintürk, M. K., eds.; Hacettepe Üniversitesi: Ankara, Turkey, 2018, p. 25-69.

Optical transduction based on localized surface plasmon resonance (LSPR) is widely used to detect biochemical interactions due to its great performance.1515 Fan, X.; White, I. M.; Shopova, S. I.; Zhu, H.; Suter, J. D.; Sun, Y.; Anal. Chim. Acta 2008, 620, 8.,1616 Al-Ogaidi, I.; Gou, H.; Al-kazaz, A. K. A.; Aguilar, Z. P.; Melconian, A. K.; Zheng, P.; Wu, N.; Anal. Chim. Acta 2014, 811, 76. LSPR is a phenomenon produced by the resonant interaction between an incident electromagnetic wave with electrons in metallic nanoparticles.1717 Mayer, K. M.; Hafner, J. H.; Chem. Rev. 2011, 111, 3828. This phenomenon is strongly dependent on the nanoparticles surrounding dielectric medium, which can be used, e.g., as a probe to monitor metallic surface functionalization.1818 Barnes, W. L.; Dereux, A.; Ebbesen, T. W.; Nature 2003, 424, 824.,1919 Anker, J. N.; Hall, W. P.; Lyandres, O.; Shah, N. C.; Zhao, J.; van Duyne, R. P.; Nat. Mater. 2008, 7, 442. Noble metals such as Au, Ag, Pd, and Pt can support surface plasmons.2020 Sugawa, K.; Tahara, H.; Yamashita, A.; Otsuki, J.; Sagara, T.; Harumoto, T.; Yanagida, S.; ACS Nano 2015, 9, 1895.

21 Zhang, N.; Han, C.; Xu, Y. J.; Foley, J. J.; Zhang, D.; Codrington, J.; Gray, S. K.; Sun, Y.; Nat. Photonics 2016, 10, 473.
-2222 Rycenga, M.; Cobley, C. M.; Zeng, J.; Li, W.; Moran, C. H.; Zhang, Q.; Qin, D.; Xia, Y.; Chem. Rev. 2011, 111, 3669. Copper has also been explored.2323 Liu, P.; Wang, H.; Li, X.; Rui, M.; Zeng, H.; RSC Adv. 2015, 5, 79738. However, Ag and Au exhibit a more intensified LSPR effect and are therefore normally employed.2020 Sugawa, K.; Tahara, H.; Yamashita, A.; Otsuki, J.; Sagara, T.; Harumoto, T.; Yanagida, S.; ACS Nano 2015, 9, 1895.,2121 Zhang, N.; Han, C.; Xu, Y. J.; Foley, J. J.; Zhang, D.; Codrington, J.; Gray, S. K.; Sun, Y.; Nat. Photonics 2016, 10, 473.,2424 Kunwar, S.; Sui, M.; Pandey, P.; Gu, Z.; Pandit, S.; Lee, J.; Sci. Rep. 2019, 9, 1329. Gold is usually the best choice for many applications as it exhibits greater inertia in the face of oxidative processes.2525 Santos, J. F. L.; Santos, M. J. L.; Thesing, A.; Tavares, F.; Griep, J.; Rodrigues, M. R. F. R.; Quim. Nova 2016, 39, 1098.

The electrochemical biosensors are summed up in four measuring techniques: potentiometric, amperometric, and conductometric/impedimetric.2626 Queirós, R. B.; Noronha, J. P.; Marques, P. V. S.; Sales, M. G. F.; Procedia Eng. 2012, 47, 758. The amperometric method is characterized by the application of a potential to the working electrode versus the reference electrode, as a result, electric currents of redox reactions are observed due to electron transfer between the electroactive analyte and the electrode; these measurements are limited by mass transport from species to the electrode. The potentiometric method is characterized by verification of the electrochemical potential observed in the electrochemical cell, which may be further classified into three types, ion-selective electrodes (IES), coated wire electrodes (CWES) and field-effect transistors (FETS).2727 Stradiotto, N. R.; Yamanaka, H.; Zanoni, M. V. B.; J. Braz. Chem. Soc. 2003, 14, 159.,2828 Ronkainen, N. J.; Halsall, H. B.; Heineman, W. R.; Chem. Soc. Rev. 2010, 39, 1747. Condutimetric/impedimetric method monitors the ability of an electroactive species to transfer electrons to the working electrode, normally under a modified electrode, such as the formation of non-conductive films. The electrochemical impedance spectroscopy (EIS) has been a good choice for biosensing applications. The EIS has some advantages as a transduction technique due to its fast detection of the target protein at the electrode-solution interface.55 Wang, B.; Jing, R.; Qi, H.; Gao, Q.; Zhang, C.; J. Electroanal. Chem. 2016, 781, 212.,2929 Daniels, J. S.; Pourmand, N.; Electroanalysis 2007, 19, 1239.,3030 Qi, H.; Shangguan, L.; Li, C.; Li, X.; Gao, Q.; Zhang, C.; Biosens. Bioelectron. 2013, 39, 324. Further, nanomaterials such as carbon nanotubes,3131 Gupta, S.; Murthy, C. N.; Prabha, C. R.; Int. J. Biol. Macromol. 2018, 108, 687. magnetic nanoparticles,3232 Freitas, M.; Viswanathan, S.; Nouws, H. P. A.; Oliveira, M. B. P. P.; Delerue-Matos, C.; Biosens. Bioelectron. 2014, 51, 195. and metallic nanostructures3333 Xia, N.; Wang, X.; Yu, J.; Wu, Y.; Cheng, S.; Xing, Y.; Liu, L.; Sens. Actuators, B 2017, 239, 834. can be used to produce electrochemical biosensors with improved properties due to the size-enhanced characteristics.3434 Tian, L.; Qian, K.; Qi, J.; Liu, Q.; Yao, C.; Song, W.; Wang, Y.; Biosens. Bioelectron. 2018, 99, 564.

The transduction step can be improved by the use of a dual-type detection using different transducers, which results in a more reliable biosensor. The elaboration of a sensor with optical and electrochemical transductions can incorporate many advantages in the device such as broader dynamic range (dynamic ranges can be combined), to obtain more complete information from the sample, and responses self-validation. In addition, the individual characteristics of each detection mode are maintained. The LSPR sensors show detection free of molecule labeling, in vivo applications, suitable for multiplexing detection, and no influence of magnetic, and electric fields on the measurements. On the other hand, the electrochemical ones do not suffer influence from ambient light variation, there is no concern with chromophore photobleaching, and shows a greater dynamic range.3535 Eltzov, E.; Cosnier, S.; Marks, R. S.; Expert Rev. Mol. Diagn. 2011, 11, 533. The incorporation of two transduction mechanisms will make the device more versatile because the two techniques can complement each other. Thus, both transductions can be used to detect the same analyte to obtain more complete data or the technique more compatible with the analysis and sample conditions can be selected to guarantee more reliable answers.

The sensors development that combines optical and electrochemical transduction has been explored for at least a decade and has found interesting applications. One of the first works found in the literature reports an amperometric and chemiluminescent sensor based on poly(methyl methacrylate)-coated optic fiber modified with poly(pyrrole-benzophenone). The device was used for detection of both urea and hydrogen peroxide using the same platform.3636 Atias, D.; Abu-Rabeah, K.; Herrmann, S.; Frenkel, J.; Tavor, D.; Cosnier, S.; Marks, R. S.; Biosens. Bioelectron. 2009, 24, 3683. In another similar work,3737 Konry, T.; Heyman, Y.; Cosnier, S.; Gorgy, K.; Marks, R. S.; Electrochim. Acta 2008, 53, 5128. an indium tin oxide (ITO)-coated optical fiber and also activated by poly(pyrrole-benzophenone) was used to immobilize and detect cholera toxin by optical (chemiluminescence) and electrochemical (amperometric) transductions. More recently, detection of avidin using a platform based on Au-coated silicon nanowires deposited on glass substrate has been shown. Surface enhanced Raman scattering (SERS) and EIS were both used for protein detection.3838 Convertino, A.; Mussi, V.; Maiolo, L.; Sci. Rep. 2016, 6, 25099. SERS has still been used in conjunction with voltammetric measurements to quantify the interaction between anti-cancer drugs and cancer cell deoxyribonucleic acid (DNA) fragment.3939 Ilkhani, H.; Hughes, T.; Li, J.; Zhong, C. J.; Hepel, M.; Biosens. Bioelectron. 2016, 80, 257.

SERS generated in nanostructures surfaces is extremely sensitive and enables substances detection at a single molecule level. However, this technique has some practical disadvantages such as the difficulty in interpreting spectra (Raman signals appear in SERS spectra and contaminants traces may contribute with peaks), relatively expensive equipment (Raman), and the Raman signal enhanced on the surface is quite critical and dependent on many aspects. Therefore, plasmonic techniques can be successfully applied in cases where an ultrasensitive analysis is not required. In this sense, a sensor that operates with EIS and the attenuated total reflectance effect (traditional Kretschmann plasmonic configuration) has been reported. Both techniques were used for clenbuterol dual detection using a graphene-coated gold surface as transducer.4040 Yan, F.; Zhang, Y.; Zhang, S.; Zhao, J.; Liu, S.; He, L.; Feng, X.; Zhang, H.; Zhang, Z.; Microchim. Acta 2015, 182, 855. Lazar et al.4141 Lazar, J.; Rosencrantz, R. R.; Elling, L.; Schnakenberg, U.; Anal. Chem. 2016, 88, 9590. showed the usage of gold-coated nanoporous polycarbonate membranes to also obtain a resourceful dual detection microfluidic device. Simultaneous signals of EIS and LSPR on the metallic nanoporous platform were evaluated in flow under various treatments performed on the membrane.

Gold nanoparticles (AuNPs) have been employed in the development of sensors that operate with optical and electrochemical transductions. AuNPs exhibit interesting optical properties (from the LSPR effect) that can be exploited. In addition, they are useful for improving the performance of electrodes in electrochemical measurements due to their good conductivity and high surface area by volume ratio. Cheng et al.4242 Cheng, X. R.; Hau, B. Y. H.; Endo, T.; Kerman, K.; Biosens. Bioelectron. 2014, 53, 513. reported a glass/ITO-based sensor covered with electrodeposited gold nanocolloids to detect DNA hybridization. A potentiostat was used to evaluate electrochemically (EIS) the glass/ITO/AuNPS surface. However, a complex light reflection system was used to evaluate the LSPR responses of the AuNPs during the transducer surface modifications. Comparatively, in other more recent work,4343 Li, N.; Zhang, D.; Zhang, Q.; Lu, Y.; Jiang, J.; Liu, G. L.; Liu, Q.; Sens. Actuators, B 2016, 231, 349. gold nanocups filled with AuNPs on glass substrate were used for dual detection of heavy metal ions. In this case, the optical evaluation of the LSPR effect was simpler and based on the light absorption. On the other hand, a less sensitive electrochemical technique (based on voltammetry) was employed.

Whereas the combination of two transduction methods in the same device becomes an interesting approach, we propose here the development of a simpler biosensing platform that explores the LSPR and EIS transduction techniques. The same transducer was shared by both detection methods and it was based on AuNPs coated-glass/ITO substrate. The AuNPs were obtained by conventional chemical reduction and chemically adsorbed on aminated ITO surfaces. The nanoparticles morphology (size and shape) is better controlled by chemical synthesis than by electrodeposition. Furthermore, the LSPR measurements were performed in conventional absorption mode and the EIS ones were obtained by the common potentiostat. The transducer surface was functionalized with cysteamine and thiolated biotin to detect streptavidin using a label-free sensing mode.

Experimental

Materials

Tetrachloroauric acid (HAuCl4, ca. 30% m/v), sodium citrate (99%), (3-aminopropyl)triethoxysilane (APTES, 99%), D-(+)-glucose (99.5%), cysteamine (95%), biotin-sulfo-NHS (BNHS, 90%), streptavidin (65%), and the redox couple [Fe(CN)6]3-/4- (99%) were purchased from Sigma-Aldrich, St. Louis, USA. The phosphate buffered saline (PBS) was obtained with 0.0820 g mL-1 NaCL (99.5%, Sigma-Aldrich, St. Louis, USA), 0.0105 g mL-1 Na2HPO4 (98%, Anidrol, Diadema, Brazil) and 0.0355 g mL-1NaH2PO4 (98%, Anidrol, Diadema, Brazil). Potassium chloride (99%, Synth, Diadema, Brazil), ITO-coated glass slide (1.0 × 2.0 × 0.1 cm, resistance 7 Ω sq-1), and Milli-Q® (Maringá, Brazil) water were also used.

Apparatus

The optical performance was evaluated using absorption spectroscopy in the visible spectral region from 450 to 900 nm (USB2000+, Ocean Optics coupled with a tungsten light source). The morphological characterization of the nanoparticles was performed using transmission electron microscopy (TEM, Jeol, JEM 1400), and the average size calculated using the software Image-Pro Plus® v. 4.5.0.29.4444 Image-Pro Plus, v. 4.5.0.29; Media Cybernetics, Silver Spring, USA, 2002. The ITO/AuNPs substrates were evaluated by scanning electron microscopy (SEM, Quanta 250, FEI) and atomic force microscopy (AFM, Shimadzu, SPM-9700), using dynamic mode. The data were analyzed using Gwyddion software version 2.51.4545 Necas, D.; Klapetek, P.; Cent. Eur. J. Phys. 2012, 10, 181. The electrochemical measurements were performed using an Autolab potentiostat/galvanostat (PGSTAT 302N, Metrohm) using a three-electrode cell with platinum sheet (2.25 cm2) as the counter-electrode, an Ag/AgCl (saturated KCl) as reference and ITO or ITO/AuNPs-coated glass slides as the working electrode (0.51 cm2).

Gold nanoparticles synthesis

The AuNPs were prepared using the Turkevich methodology.4646 Turkevich, J.; Stevenson, P. C.; Hillier, J.; Discuss. Faraday Soc. 1951, 11, 55. Briefly, in a three-necked bottle, HAuCl4 aqueous solution (100 mL, 0.25 mmol L-1) was stirred and heated to boil. Then, sodium citrate (2.5 mL, 1% m/v) was added to the solution under vigorous stirring at 100 °C. After 40 min, the mixture was rapidly cooled in an ice-bath to stop the reaction. The suspension was stored in a refrigerator.

Substrates preparation

First, ITO/glass slides (2.0 × 1.0 × 0.1 cm) were cleaned in a water bath with soap, washed with deionized water and then ultrasonicated in Milli-Q® water, ethanol and acetone for 10 min each followed by drying in N2 (99.99%, White Martins, Londrina, Brazil). The hydrophilization of the ITO surface was performed in a plasma cleaner (Harrick Plasma, PDC-002) for 60 s. Subsequently, the substrates were immersed in an APTES ethanolic solution (4% v/v) at 50 °C overnight followed by washing with ethanol, water and dried by N2 blow. Then, the substrates were left in an oven for 1 h at 100 °C. Finally, they were cooled to room temperature and vertically immersed in AuNPs solution under stirring for 2 h, washed with deionized water, and dried with N2 again.

The AuNPs deposition and electroactive surface area were evaluated by cyclic voltammetry using an electrochemical probe ([Fe(CN)6]3-/4-, 5.0 mmol L-1) with reversible reaction, and 0.5 mol L-1 KCl in the range of -0.2 to 0.8 V and 10 to 150 mV s-1 (scan rate). The plasmonic sensitivity was studied by absorbance spectroscopy changing the surrounding medium of the AuNPs using glucose solutions. The plasmonic bulk sensitivity was obtained by the slope of the wavelength shift against the refractive index (RI) of the glucose solution. The RI of the glucose was measured using a portable refractometer (3850 PAL-RI, Atago).

Optical and electrochemical bioassay

At first, ITO/AuNPs substrates were washed with deionized water and immersed in an aqueous solution of cysteamine (50 µmol L-1) for 6 h. Subsequently, the slides were washed in deionized water and immersed for 12 h in BNHS (0.1 mg mL-1). Finally, the substrates were immersed in a 0.5 mg mL-1 PBS solution (pH 7.45) of streptavidin for 4 h.4747 Monteiro, J. P.; de Oliveira, J. H.; Radovanovic, E.; Brolo, A. G.; Girotto, E. M.; Plasmonics 2016, 11, 45.,4848 Cabral Jr., B. P.; de Oliveira, J. H.; Buzzetti, P. H. M.; Fressatti, M. G. V.; Monteiro, J. P.; Amorin, L. H. C.; Urbano, A.; Radovanovic, E.; Girotto, E. M.; Mater. Lett. 2018, 227, 243.

The EIS was applied to determine the charge transfer resistance before and after the functionalization. The measurements were performed in PBS solution (pH 7.45) using the potential of -100 mV, frequency range of 1 kHz to 10 mHz, and 15 mV of amplitude. The three-electrode electrochemical cell was thermostated at 25 °C and was used ITO/AuNPs as the working electrode. Furthermore, using the same substrate rinsed lightly with PBS, the plasmonic band redshift was monitored after each functionalization step after the EIS measurement. For this, the ITO/AuNPs substrate was inserted into a cuvette and the LSPR spectra were evaluated in PBS solution using a spectrometer.

Results and Discussion

Gold nanoparticles synthesis

After the reaction, the solution showed a red-wine color indicating the AuNPs formation. Figure 1a shows the absorbance spectra of the solution with λmax = 506.1 nm, which is related to LSPR of gold nanoparticles.4949 Ding, L.; Gao, Y.; Di, J.; Biosens. Bioelectron. 2016, 83, 9. The single absorption band with symmetric shape indicates a highly homogeneous synthesis of gold nanospheres, which can be proved with AuNPs images of TEM (Figure 1b). The narrow size distribution obtained in the histogram (Figure 1b inset) shows a good homogeneity with respect to the diameter. The micrograph presented spherical nanoparticles with a calculated average diameter of 13.0 ± 0.3 nm.

Figure 1
(a) Absorption spectra of AuNPs solution and (b) transmission electron microscopy image with a size distribution histogram.

ITO/AuNPs substrates characterization

Typical SEM images of the AuNPs assembled onto ITO substrates are displayed in Figure 2b. It is possible to notice a highly covered surface due to the strong interaction between the AuNPs and the aminated surface of the substrate, overlapping the AuNPs-AuNPs interactions. The substrate presented an absorption band with λmax = 533.5 nm as illustrated in Figure 2a. A redshift of ca. 27 nm was observed in relation to the absorption maximum for the AuNPs suspension. This may be justified due to the significant decrease in interparticle distance during anchoring. It was also noticed that the AuNPs remained roughly spherical in shape. This is also supported because a single and well-defined plasmon band was still observed. In addition, the nanoparticles maintained good homogeneity of diameter distribution, since the full width at half maximum (FWHM) was of 88.9 nm that is close to the obtained AuNPs in suspension (FWHM = 63.1 nm). The agglomeration process caused the band broadening.5050 Chen, K. J.; Lu, C. J.; Talanta 2010, 81, 1670.

Figure 2
(a) Absorption spectrum and (b) scanning electron microscopy image of the AuNPs-coated glass-ITO substrates.

Evaluation of LSPR and electrochemical responses

Figure 3 shows the visible absorption spectra of ITO/AuNPs in different glucose solutions. The plasmonic band presented a redshift as the refractive index increased as shown in the inset of Figure 3a. The sensitivity was obtained by the slope of λmax versus glucose solutions RI (Figure 3b) curve. The ITO/AuNPs presented a sensitivity of 88.4 nm RIU-1, value comparable to that of self-assembled monolayers (SAMs) of gold nanoparticles on silicon substrates.1717 Mayer, K. M.; Hafner, J. H.; Chem. Rev. 2011, 111, 3828.,5151 Chen, H.; Kou, X.; Yang, Z.; Ni, W.; Wang, J.; Langmuir 2008, 24, 5233. The bulk sensitivity is an important parameter, since it measures the plasmonic wave disturbance due to dielectric environment changes.1717 Mayer, K. M.; Hafner, J. H.; Chem. Rev. 2011, 111, 3828. This can be used to determine the substrate applicability in the monitoring of surface-anchored biomolecules.5252 Jeon, J.; Uthaman, S.; Lee, J.; Hwang, H.; Kim, G.; Yoo, P. J.; Hammock, B. D.; Kim, C. S.; Park, Y. S.; Park, I. K.; Sens. Actuators, B 2018, 266, 710.,5353 Li, X.; Jiang, L.; Zhan, Q.; Qian, J.; He, S.; Colloids Surf., A 2009, 332, 172. The sensitivity curves exhibited a coefficient of determination (R2) close to 0.9483, which represents that ca. 95% of the λmax shift is explained by the correlation. A linear behavior is usually sought because the disturbances in plasmonic wave are integrally converted in outputs signals in this fitting type.

Figure 3
(a) Absorption spectra of glass-ITO/AuNPs substrate immersed in different glucose solutions with a closer view of the maximum absorption and (b) its linear fit with RI sensitivity of 88.4 nm RIU-1.

The electrochemical test indicated an increase in the electroactive surface area and conductivity when the AuNPs were anchored onto the glass-ITO.5454 Sonuç Karaboğa, M. N.; Şimşek, Ç. S.; Sezgintürk, M. K.; Biosens. Bioelectron. 2016, 84, 22.,5555 Khan, M. Z. H.; J. Electron. Mater. 2017, 46, 2254. Figure 4 shows the voltammograms of glass-ITO and glass-ITO/AuNPs substrates in different scanning rates. The Randles Sevick equation5656 Brett, A. M. O.; Brett, C. M. A.; Electroquimica: Príncipios, Métodos e Aplicações; Livraria Almedina: Coimbra, Portugal, 1996. has been used to calculate the electroactive area from the voltammograms results. Using this equation for the glass-ITO and glass-ITO/AuNPs linear curves (insets in Figures 4a and 4b), it was calculated an electroactive area of 0.81 ± 0.04 and 0.89 ± 0.08 cm2, respectively. Based on these results, the glass-ITO/AuNPs presented an average electroactive area 17.7% higher than glass-ITO. This shows that the active regions of the electrode were enhanced mainly because of the roughness and conductivity enhance. The AFM topographical analysis of glass-ITO and glass-ITO/AuNPs are shown in Figure 5, respectively. The difference between topographies is clear. The AuNPs presence is noticed because the surface roughness shows a more symmetrical and more uniform distribution. Furthermore, the maximum Z range of the treated images shows a maximum height of 41.9 nm with AuNPs, which is 11 nm higher than pure ITO. The height increasing is close to the average radius for nanoparticles in aqueous suspension (13 nm). The analysis by Gwyddion presented a root mean square roughness (RMS) of 3.70 nm to glass-ITO and 4.35 nm to glass-ITO/AuNPs. The RMS ratio presented an increase of 17% due to the anchoring of AuNPs on ITO surface.5757 Slimi, H.; Barhoumi, A.; Waldhoff, N.; Duponchel, B.; Poupin, C.; Cousin, R.; Leroy, G.; Guermazi, S.; Superlattices Microstruct. 2018, 120, 670.,5858 Silva, A. T. B.; Coelho, A. G.; Lopes, L. C. S.; Martins, M. V. A.; Crespilho, F. N.; Merkoçi, A.; da Silva, W. C.; J. Braz. Chem. Soc. 2013, 24, 1237. In fact, this increase is probably confirming the enhancement of the electroactive area.

Figure 4
Cyclic voltammograms of K3[Fe(CN)6] (5.0 mmol L-1) in KCl (0.5 mol L-1) to the (a) glass-ITO and (b) glass-ITO/AuNPs substrates. Scanning rates were from 10 to 150 mV s-1 (25 °C) and the anodic peak at 380 mV was evaluated. Insets show the current of the anodic peak versus square root of the scan rate.

Figure 5
Atomic force microscopy images for the (a) glass-ITO and (b) glass-ITO/AuNPs substrates.

Bioassay of streptavidin dual detection

The first step to the biofunctionalization was to prepare a cysteamine layer on the AuNPs, as a result of the strong interaction between the Au surface and thiol groups of the cysteamine ensuring aminated nanoparticles.5959 Mandenius, C. F.; Welin, S.; Danielsson, B.; Lundström, I.; Mosbach, K.; Anal. Biochem. 1984, 137, 106. Secondly, the BNHS molecule bonds in the cysteamine -NH2 group by the removal of N-hydroxysuccinimide group. Then, it was possible to detect the streptavidin using the glass-ITO/AuNPs substrate due to the huge host-guest interaction with biotin (affinity constant (Ka) ca. 1015 mol L-1).6060 Ozawa, M.; Ozawa, T.; Nishio, M.; Ueda, K.; J. Mol. Graphics Modell. 2017, 75, 117.,6161 Zhang, K.; Liu, G.; Goldys, E. M.; Biosens. Bioelectron. 2018, 102, 80.

The plasmonic wave disturb due to the functionalization on AuNPs allowed to monitor the streptavidin presence by the LSPR band redshifts (Figure 6a).6262 Sepúlveda, B.; Angelomé, P. C.; Lechuga, L. M.; Liz-Marzán, L. M.; Nano Today 2009, 4, 244. The average λmax (Figure 6a) after each modification was expressed in a bar chart, Figure 6b. The measures were performed in triplicate. The redshift caused by streptavidin was 4.2 ± 0.2 in relation to the biotin-covered AuNPs λmax. The higher disturbed in the plasmonic band for streptavidin molecules occurred because its show higher molecular weight (protein).5353 Li, X.; Jiang, L.; Zhan, Q.; Qian, J.; He, S.; Colloids Surf., A 2009, 332, 172.

Figure 6
(a) Absorption spectra and (b) average λmax for each functionalization steps of the biotest assay.

Regarding the electrochemical transduction, it should be emphasized that the non-conductive species adsorption in working electrodes surfaces makes it less conductive and makes the redox process difficult. EIS technique can monitor these events using an electrochemical probe.55 Wang, B.; Jing, R.; Qi, H.; Gao, Q.; Zhang, C.; J. Electroanal. Chem. 2016, 781, 212. At -100 mV, the dissolved O2 is involved in redox reactions and can be used as a probe to evaluate the decrease in conductivity of the electrode. Basically, EIS describe the system response due to a sinusoidal oscillation in the potential. The signal is measured by the impedance spectra, usually Nyquist graphs, where the semicircle diameter indicates the charge transfer resistance (Rct). This parameter is a quantitative response that measures the capacity of an electroactive specie (O2 in this case) to share electrons on the electrode surface.4242 Cheng, X. R.; Hau, B. Y. H.; Endo, T.; Kerman, K.; Biosens. Bioelectron. 2014, 53, 513.,6363 Girotto, E. M.; de Paoli, M.-A.; Quim. Nova 1999, 22, 358.

Figure 7a shows the Nyquist diagram for the electrodes under each modification step (AuNPs, cysteamine, biotin, streptavidin). The Rct values are given in graph bars (Figure 7b). Glass-ITO presented a Rct of 907.7 Ω cm-2 while glass-ITO/AuNPs gave a 758.2 Ω cm-2 resistance. The presence of AuNPs enhanced the conductivity at the electrodes surface and provided a larger electroactive area. The cysteamine layer raised the Rct to 809.1 Ω cm-2 due to the steric hindrance caused by the compact non-conductive layer.5454 Sonuç Karaboğa, M. N.; Şimşek, Ç. S.; Sezgintürk, M. K.; Biosens. Bioelectron. 2016, 84, 22.,6464 Liu, S. Q.; Ju, H. X.; Anal. Biochem. 2002, 307, 110. The posterior incorporation of biotin and streptavidin layer resulted in higher Rct values of 830.1 and 1028.8 Ω cm-2, respectively. The higher Rct obtained by streptavidin is related to the high-volume non-conductive protein making the charge transfer more difficult.6565 Pei, R.; Cheng, Z.; Wang, E.; Yang, X.; Biosens. Bioelectron. 2001, 16, 355.,6666 Rajesh; Sharma, V.; Mishra, S. K.; Biradar, A. M.; Mater. Chem. Phys. 2012, 132, 22. It was also clear the possibility of detecting streptavidin molecules by the increase in the Rct.

Figure 7
(a) Nyquist diagram in PBS (pH = 7.45) and (b) bar graph to charge transfer resistance for each working electrode functionalization steps. Amplitude used was 15 mV and frequency range was between 1 kHz and 10 mHz at −100 mV.

Conclusions

The AuNPs were morphologically and spectroscopically characterized and prepared as a self-assembled monolayer on glass-ITO. The substrate glass-ITO had a plasmonic sensitivity of 88.4 nm RIU-1, which shows that the substrate was able for LSPR transduction. Furthermore, the assembly of AuNPs monolayer on glass-ITO promoted the desired increase in the electroactive area of the substrate. The dual detection method presented a reliable detection of the streptavidin using a few steps functionalization onto gold nanoparticles. A redshift of ca. 4 nm in the LSPR band and an increase of ca. 200 Ω cm-2 in the Rct were observed when streptavidin molecules were assembled on biotin monolayer. Both transduction techniques were able to monitor streptavidin using a shared transducer based on glass-ITO/AuNPs. It should also be noted that this device can provide real-time optical transduction, which is an interesting feature for the biosensing area.

Acknowledgments

We thank the Microscopy Central Complex Research Support (COMCAP) of the State University of Maringá (UEM)-Brazil for equipment support and Brazilian Agency CAPES for the financial support. This work was also supported by CNPq (Brazil) through Inomat-INCT, National Institute for Complex Functional Materials (CNPq 573644/2008-0).

References

  • 1
    Liu, D.; Luo, Q.; Deng, F.; Li, Z.; Li, B.; Shen, Z.; Anal. Chim. Acta 2017, 971, 26.
  • 2
    Ashley, J.; Li, S. F. Y.; Biosens. Bioelectron. 2013, 48, 126.
  • 3
    López, M. S.-P.; Redondo-Gómez, E.; López-Ruiz, B.; Talanta 2017, 175, 209.
  • 4
    Ertürk, G.; Özen, H.; Tümer, M. A.; Mattiasson, B.; Denizli, A.; Sens. Actuators, B 2016, 224, 823.
  • 5
    Wang, B.; Jing, R.; Qi, H.; Gao, Q.; Zhang, C.; J. Electroanal. Chem. 2016, 781, 212.
  • 6
    Tepeli, Y.; Ülkü, A.; Sens. Actuators, B 2018, 254, 377.
  • 7
    Dali, M.; Zinoubi, K.; Chrouda, A.; Abderrahmane, S.; Cherrad, S.; Jaffrezic-Renault, N.; J. Electroanal. Chem. 2018, 813, 9.
  • 8
    Faridbod, F.; Gupta, V. K.; Zamani, H. A.; Int. J. Electrochem. 2011, 2011, 352546.
  • 9
    Zeng, Y.; Zhu, Z.; Du, D.; Lin, Y.; J. Electroanal. Chem. 2016, 781, 147.
  • 10
    Sassolas, A.; Blum, L. J.; Leca-Bouvier, B. D.; Biotechnol. Adv. 2012, 30, 489.
  • 11
    Singh, A.; Thapa, K. B.; Kumar, N.; Optik 2015, 126, 244.
  • 12
    Mehrotra, P.; J. Oral Biol. Craniofacial Res. 2016, 6, 153.
  • 13
    Braga, J. R. G.; Queiroz, A. A. A.; Ramos, A. C. B.; Int. J. Comput. Digital Syst. 2012, 1, 1.
  • 14
    Senay, R. H.; Ulucan, F.; Kuru, C. İ.; Akgöl, S. In Smart Biosensing Technologies: Past, Present and Future; Gürsan, A. E.; Sezgintürk, M. K., eds.; Hacettepe Üniversitesi: Ankara, Turkey, 2018, p. 25-69.
  • 15
    Fan, X.; White, I. M.; Shopova, S. I.; Zhu, H.; Suter, J. D.; Sun, Y.; Anal. Chim. Acta 2008, 620, 8.
  • 16
    Al-Ogaidi, I.; Gou, H.; Al-kazaz, A. K. A.; Aguilar, Z. P.; Melconian, A. K.; Zheng, P.; Wu, N.; Anal. Chim. Acta 2014, 811, 76.
  • 17
    Mayer, K. M.; Hafner, J. H.; Chem. Rev. 2011, 111, 3828.
  • 18
    Barnes, W. L.; Dereux, A.; Ebbesen, T. W.; Nature 2003, 424, 824.
  • 19
    Anker, J. N.; Hall, W. P.; Lyandres, O.; Shah, N. C.; Zhao, J.; van Duyne, R. P.; Nat. Mater. 2008, 7, 442.
  • 20
    Sugawa, K.; Tahara, H.; Yamashita, A.; Otsuki, J.; Sagara, T.; Harumoto, T.; Yanagida, S.; ACS Nano 2015, 9, 1895.
  • 21
    Zhang, N.; Han, C.; Xu, Y. J.; Foley, J. J.; Zhang, D.; Codrington, J.; Gray, S. K.; Sun, Y.; Nat. Photonics 2016, 10, 473.
  • 22
    Rycenga, M.; Cobley, C. M.; Zeng, J.; Li, W.; Moran, C. H.; Zhang, Q.; Qin, D.; Xia, Y.; Chem. Rev. 2011, 111, 3669.
  • 23
    Liu, P.; Wang, H.; Li, X.; Rui, M.; Zeng, H.; RSC Adv. 2015, 5, 79738.
  • 24
    Kunwar, S.; Sui, M.; Pandey, P.; Gu, Z.; Pandit, S.; Lee, J.; Sci. Rep. 2019, 9, 1329.
  • 25
    Santos, J. F. L.; Santos, M. J. L.; Thesing, A.; Tavares, F.; Griep, J.; Rodrigues, M. R. F. R.; Quim. Nova 2016, 39, 1098.
  • 26
    Queirós, R. B.; Noronha, J. P.; Marques, P. V. S.; Sales, M. G. F.; Procedia Eng. 2012, 47, 758.
  • 27
    Stradiotto, N. R.; Yamanaka, H.; Zanoni, M. V. B.; J. Braz. Chem. Soc. 2003, 14, 159.
  • 28
    Ronkainen, N. J.; Halsall, H. B.; Heineman, W. R.; Chem. Soc. Rev. 2010, 39, 1747.
  • 29
    Daniels, J. S.; Pourmand, N.; Electroanalysis 2007, 19, 1239.
  • 30
    Qi, H.; Shangguan, L.; Li, C.; Li, X.; Gao, Q.; Zhang, C.; Biosens. Bioelectron. 2013, 39, 324.
  • 31
    Gupta, S.; Murthy, C. N.; Prabha, C. R.; Int. J. Biol. Macromol. 2018, 108, 687.
  • 32
    Freitas, M.; Viswanathan, S.; Nouws, H. P. A.; Oliveira, M. B. P. P.; Delerue-Matos, C.; Biosens. Bioelectron. 2014, 51, 195.
  • 33
    Xia, N.; Wang, X.; Yu, J.; Wu, Y.; Cheng, S.; Xing, Y.; Liu, L.; Sens. Actuators, B 2017, 239, 834.
  • 34
    Tian, L.; Qian, K.; Qi, J.; Liu, Q.; Yao, C.; Song, W.; Wang, Y.; Biosens. Bioelectron. 2018, 99, 564.
  • 35
    Eltzov, E.; Cosnier, S.; Marks, R. S.; Expert Rev. Mol. Diagn. 2011, 11, 533.
  • 36
    Atias, D.; Abu-Rabeah, K.; Herrmann, S.; Frenkel, J.; Tavor, D.; Cosnier, S.; Marks, R. S.; Biosens. Bioelectron. 2009, 24, 3683.
  • 37
    Konry, T.; Heyman, Y.; Cosnier, S.; Gorgy, K.; Marks, R. S.; Electrochim. Acta 2008, 53, 5128.
  • 38
    Convertino, A.; Mussi, V.; Maiolo, L.; Sci. Rep. 2016, 6, 25099.
  • 39
    Ilkhani, H.; Hughes, T.; Li, J.; Zhong, C. J.; Hepel, M.; Biosens. Bioelectron. 2016, 80, 257.
  • 40
    Yan, F.; Zhang, Y.; Zhang, S.; Zhao, J.; Liu, S.; He, L.; Feng, X.; Zhang, H.; Zhang, Z.; Microchim. Acta 2015, 182, 855.
  • 41
    Lazar, J.; Rosencrantz, R. R.; Elling, L.; Schnakenberg, U.; Anal. Chem. 2016, 88, 9590.
  • 42
    Cheng, X. R.; Hau, B. Y. H.; Endo, T.; Kerman, K.; Biosens. Bioelectron. 2014, 53, 513.
  • 43
    Li, N.; Zhang, D.; Zhang, Q.; Lu, Y.; Jiang, J.; Liu, G. L.; Liu, Q.; Sens. Actuators, B 2016, 231, 349.
  • 44
    Image-Pro Plus, v. 4.5.0.29; Media Cybernetics, Silver Spring, USA, 2002.
  • 45
    Necas, D.; Klapetek, P.; Cent. Eur. J. Phys. 2012, 10, 181.
  • 46
    Turkevich, J.; Stevenson, P. C.; Hillier, J.; Discuss. Faraday Soc. 1951, 11, 55.
  • 47
    Monteiro, J. P.; de Oliveira, J. H.; Radovanovic, E.; Brolo, A. G.; Girotto, E. M.; Plasmonics 2016, 11, 45.
  • 48
    Cabral Jr., B. P.; de Oliveira, J. H.; Buzzetti, P. H. M.; Fressatti, M. G. V.; Monteiro, J. P.; Amorin, L. H. C.; Urbano, A.; Radovanovic, E.; Girotto, E. M.; Mater. Lett. 2018, 227, 243.
  • 49
    Ding, L.; Gao, Y.; Di, J.; Biosens. Bioelectron. 2016, 83, 9.
  • 50
    Chen, K. J.; Lu, C. J.; Talanta 2010, 81, 1670.
  • 51
    Chen, H.; Kou, X.; Yang, Z.; Ni, W.; Wang, J.; Langmuir 2008, 24, 5233.
  • 52
    Jeon, J.; Uthaman, S.; Lee, J.; Hwang, H.; Kim, G.; Yoo, P. J.; Hammock, B. D.; Kim, C. S.; Park, Y. S.; Park, I. K.; Sens. Actuators, B 2018, 266, 710.
  • 53
    Li, X.; Jiang, L.; Zhan, Q.; Qian, J.; He, S.; Colloids Surf., A 2009, 332, 172.
  • 54
    Sonuç Karaboğa, M. N.; Şimşek, Ç. S.; Sezgintürk, M. K.; Biosens. Bioelectron. 2016, 84, 22.
  • 55
    Khan, M. Z. H.; J. Electron. Mater. 2017, 46, 2254.
  • 56
    Brett, A. M. O.; Brett, C. M. A.; Electroquimica: Príncipios, Métodos e Aplicações; Livraria Almedina: Coimbra, Portugal, 1996.
  • 57
    Slimi, H.; Barhoumi, A.; Waldhoff, N.; Duponchel, B.; Poupin, C.; Cousin, R.; Leroy, G.; Guermazi, S.; Superlattices Microstruct. 2018, 120, 670.
  • 58
    Silva, A. T. B.; Coelho, A. G.; Lopes, L. C. S.; Martins, M. V. A.; Crespilho, F. N.; Merkoçi, A.; da Silva, W. C.; J. Braz. Chem. Soc. 2013, 24, 1237.
  • 59
    Mandenius, C. F.; Welin, S.; Danielsson, B.; Lundström, I.; Mosbach, K.; Anal. Biochem. 1984, 137, 106.
  • 60
    Ozawa, M.; Ozawa, T.; Nishio, M.; Ueda, K.; J. Mol. Graphics Modell. 2017, 75, 117.
  • 61
    Zhang, K.; Liu, G.; Goldys, E. M.; Biosens. Bioelectron. 2018, 102, 80.
  • 62
    Sepúlveda, B.; Angelomé, P. C.; Lechuga, L. M.; Liz-Marzán, L. M.; Nano Today 2009, 4, 244.
  • 63
    Girotto, E. M.; de Paoli, M.-A.; Quim. Nova 1999, 22, 358.
  • 64
    Liu, S. Q.; Ju, H. X.; Anal. Biochem. 2002, 307, 110.
  • 65
    Pei, R.; Cheng, Z.; Wang, E.; Yang, X.; Biosens. Bioelectron. 2001, 16, 355.
  • 66
    Rajesh; Sharma, V.; Mishra, S. K.; Biradar, A. M.; Mater. Chem. Phys. 2012, 132, 22.

Publication Dates

  • Publication in this collection
    29 Mar 2021
  • Date of issue
    Apr 2021

History

  • Received
    01 Apr 2020
  • Accepted
    17 Nov 2020
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br