Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2808 2022-07-20 10:00:39 |
2 No change Meta information modification 2808 2022-07-20 21:28:54 | |
3 Reference format revised. Meta information modification 2808 2022-07-21 04:10:44 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Pourmadadi, M.;  Dinani, H.S.;  Tabar, F.S.;  Khassi, K.;  Janfaza, S.;  Tasnim, N.;  Hoorfar, M.; Soleimani Dinani, H. Applications of Graphene in Biosensors for Cancer Detection. Encyclopedia. Available online: https://encyclopedia.pub/entry/25319 (accessed on 18 May 2024).
Pourmadadi M,  Dinani HS,  Tabar FS,  Khassi K,  Janfaza S,  Tasnim N, et al. Applications of Graphene in Biosensors for Cancer Detection. Encyclopedia. Available at: https://encyclopedia.pub/entry/25319. Accessed May 18, 2024.
Pourmadadi, Mehrab, Homayoon Soleimani Dinani, Fatemeh Saeidi Tabar, Kajal Khassi, Sajjad Janfaza, Nishat Tasnim, Mina Hoorfar, Homayoon Soleimani Dinani. "Applications of Graphene in Biosensors for Cancer Detection" Encyclopedia, https://encyclopedia.pub/entry/25319 (accessed May 18, 2024).
Pourmadadi, M.,  Dinani, H.S.,  Tabar, F.S.,  Khassi, K.,  Janfaza, S.,  Tasnim, N.,  Hoorfar, M., & Soleimani Dinani, H. (2022, July 20). Applications of Graphene in Biosensors for Cancer Detection. In Encyclopedia. https://encyclopedia.pub/entry/25319
Pourmadadi, Mehrab, et al. "Applications of Graphene in Biosensors for Cancer Detection." Encyclopedia. Web. 20 July, 2022.
Applications of Graphene in Biosensors for Cancer Detection
Edit

Biosensors are a very promising tool for the possibility of sensitive, specific, and non-invasive diagnosis for early detection of cancer. Effective, accurate methods of cancer detection and clinical diagnosis are urgently needed. Biosensors are devices that are designed to detect a specific biological analyte by essentially converting a biological entity (i.e., protein, DNA, RNA) into an electrical signal that can be detected and analyzed. The ultimate goal of biosensors is to detect signals specific to each disease and cancer, and recently nanoparticles have been widely used in the design of biosensors. Graphene and its derivatives are nanoparticles with unique properties and have many applications in nanobiosensors.

graphene cancer biomarkers nanobiosensor biosensing method diagnostic applications

1. Introduction

Biosensor technology is a growing field to satisfy the need for sensitive and rapid detection problems [1]. In addition to the use of commercial SPR-based biosensors using the Kerchman plasma excitation scheme [2], in recent years, the use of optical, electrochemical, piezoelectric, and other types of biosensors has shown successful results in low concentrations of biomarkers for the detection of cancers including breast cancer, lung cancer, and prostate cancer [3]. In fact, with the proliferation of biosensors capable of detecting cancer early, a major revolution has taken place in the field of cancer diagnosis [4]. Biosensors are a set of components that record a physical, chemical, or biological change and convert it into a measurable marking [5]. The main parts of biosensors include the bioreceptor, transducer, and read-out system [6]. The bioreceptor reacts with the target substance and is usually composed of an antibody, aptamer, cell, or enzyme. Another essential part is the transducer, which must convert the biological signal into a measurable signal. The transducer can be electrochemical, optical, or mechanical [7]. Finally, the read-out system displays the final response and reads the measured signal [6][7]. Electrochemical biosensors provide high sensitivity and selectivity for vital biomarkers that are responsible for vital molecular events in tumor formation and progression [8]. In fact, in addition to differentiating tumor cells from normal cells, these sensors lead to the targeted diagnosis of localized tumor cells and circulating tumor cells. Electrochemical sensors are suitable tools for cell counting, cell classification, and the detection of tumor cells. In the diagnosis of tumor cells, electrochemical biosensors achieve not only high sensitivity (limit of detection of 10 tumor cells in a 250 μL sample) and high specificity but also diagnose duplicate antigens at the tumor cell surface and successfully prevent false-positive results [9].

In electrochemical biosensors, a redox reaction takes place between the bioreceptor and the target molecule [10]. The reaction in the electrochemical transducer requires a reference, counter, and working electrode, the working electrode acts as a transducer [11], and tumor antigens are used as biomarkers for cancer diagnosis in biosensors [12]. The ultimate goal of biosensors is to detect signals specific to each disease and cancer, and recently nanoparticles have been widely used in the design of biosensors. These nanoparticles can react with the analyte in the sample and increase the detection range [13][14]. Nanostructured materials as guiding elements in biosensors that increase the stability of bioreceptor loading at the electrode surface by providing a more electroactive surface using functional groups. There are various purposes for using nanoparticles to make nanobiosensors, including increasing the surface to stabilize biomaterials and thus increasing sensitivity, catalyzing the process, large dynamic range, the possibility of reacting at low potentials, and helping the quick transfer of electrons from the active center of the electrode surface reaction in the electrochemical nanobiosensors [15]. Nanoparticles can be easily fabricated using chemical methods and react with the analyte in the sample, and their inherent properties, such as optical or magnetic properties, can be used [16][17]. Graphene and its derivatives are nanoparticles with unique properties and have many applications in nanobiosensors [18]. Graphene is a two-dimensional (2D) sheet of carbon atoms in a hexagonal (honeycomb) configuration that is an allotrope of carbon [19][20]. Graphene sheets are formed by placing carbon atoms side by side, and in a graphene sheet, each carbon atom is bonded to three other carbon atoms [21]. These three bonds are on the same sheet, and the angles between them are equal to 120°. The length of the C-C bond in graphene is about 0.142 nm [22]. The inherent strength of graphene layers results from these bonds, known as covalent bonds [23]. While the carbon bonds are sp2 hybridized, and the σ C-C bond is the strongest bond in materials, the pi (π) bond is responsible for the electron conduction of graphene [24]. With these unique structural properties, graphene has been shown to have special properties and has attracted much interest in scientific research [25]. Graphene transfers heat better than any alternative material and is an excellent electrical conductor with unique optical properties. The derivatives of graphene such as Graphene Oxide (GO), Reduced Graphene Oxide (RGO), and Graphene Quantum Dots (GQDs) have entered the field of graphene research and nanobiosensors because of their unique properties [25]. GO is a derivative of graphene that is obtained by graphite using oxidizing materials [26]. GO is one typical 2D structured and oxygenated planer molecular material. RGO is a reduced form of GO that contains a π-conjugated system. The only major difference between GO and RGO is the number of oxygen atoms present and their conductivity [27]. RGO has low electrical conductivity due to the disruption of the main structure of graphene, but it has a high ability to be functionalized with a variety of chemical and biological molecules. GQDs consist of one or a few layers of graphite and are smaller than 100 nm [28].

2. Graphene-Based Materials and Their Properties

The physical properties of carbon materials are related to their hybridization stats (sp, sp2, sp3) [29][30][31][32]. Graphene with sp2 hybridization is a fine and zero bandgap semiconductor, but the sp3 hybridization diamond is a hard insulator. Graphene is highly thermally conductive, chemically stable, and flexible [33]. One of the superior properties of graphene is that its charge carriers treat as massless particles, and they can move with little scattering in an ambient condition [34]. The charge transport of graphene and its electronic properties are due to its great electronic band structure [35]. Especially among all nanomaterials, graphene has a wide surface area (2630 m2g−1) [24][25] and is available for direct interaction with many biomolecules [36]. It can be wielded with structural defects using low-cost fabrication methods by chemical modification [37]. The sensitivity of the electrical resistance of graphene to the adsorption makes it highly useful for highly sensitive sensing applications [33].
Other unique properties of graphene, including its optical, magnetic, and high elasticity, make it a suitable monolayer structure for preparing several graphene-based nanocomposites. Graphene has one of the highest tensile strengths of all materials and a high Young’s modulus, which is the relation between stress and strain, that gives graphene its great mechanical properties [38][39]. Using graphene in several sensing applications based on an electrochemical read-out has been offered in various chemical and biological sensors [40]. Graphene derivatives, such as GO and RGO, have been consumed for the fabrication of a diverse range of graphene-based nanocomposites in biosensors by a mixture of metal and biomolecules with enhanced sensitivity [33][41]. Because of the great surface-to-volume ratio and functional chemical groups, GO has broad capacities for the adsorption of biomolecules [42]. GO is composed of graphene layers with active oxygen-containing functional groups on its surface, such as hydroxyl, epoxy, and carboxyl. GO has unique physicochemical properties such as its small size between 20 nm–100 nm, conductivity, and optical and electronic properties [43][44]. Furthermore, GO is hydrophilic and water soluble, whereas graphene is hydrophobic and does not dissolve easily in water [25]. GO has a significantly disrupted sp2 carbon network and shows many defects, and functional groups make its insulators [24]. GO is not conductive and has reduced mechanical properties compared to graphene. Accordingly, to improve the conductivity of GO, it is necessary to convert GO into RGO. Graphene nanomaterials can be successfully functionalized through non-covalent or covalent interactions. Typical covalent reactions include oxidation, reduction, radicals, and nucleophilic/electrophilic additives [45].

3. Graphene-Based Biosensors

3.1. Graphene-Based FET Biosensors

Over the last few decades, attempts have been made to utilize semiconductor field-effect transistors (FETs) in biological and chemical sensors due to their well-characterized behavior and ease of use of portable devices. In addition, the electrical measurement of bimolecular interactions is particularly desirable due to its adequacy for low-cost mobile sensors that could be used in the area by non-technical individuals [46]. FET-based biosensors are excellent candidates for several label-free transducers. They have gained much more attention in recent years for their considerable advantages, such as high scalability, ultra-sensitivity, rapid real-time analysis, inherent amplification, reduced power needs, direct electrical read-out, and low-cost bulk development, as opposed to surface plasmon resonance, microcantilever sensors, fluorescence instruments, and other approaches [4][47].
An FET-based biosensor includes three electrodes—the source, the drain, and the gate—such that the section between the drain and the source functions as a biological detection component that interacts with the target analyte/biomolecules and senses their presence, concentration, and electrical activity. The biosensor then converts the biological information directly into a detectable signal [48]. Subsequently, based on the implementation, the signal acquired can be illustrated, amplified, stored, and analyzed or sent to the data center for additional processes [49]. The function of the FET-based biosensor can be encapsulated as: (1) an alteration in the concentration of the sample solution leads to a difference in the charge close to the interface of the sensor; (2) this shift of charge causes a decrease in the effective gate voltage; (3) this variation in the effective gate voltage contributes to a change in the current flow of the drain [50].
Due to the linear dispersion relation of the charge carriers in graphene, they have a very small effective mass and excellent carrier mobilities [51]. Furthermore, graphene-based nanomaterials have high thermal conductivity, as well as availability and low cost. The charge transfer nature of graphene enables a special channel material that replaces silicon and other common semiconductors. In comparison to other semiconductors, graphene does not require impurity doping to conduct electricity. It demonstrates self-doping phenomena that cause the carrier type and concentration to be regulated with the aid of an external electrical field. The alternating ability between GFET’s n- and p-channel varies from other FET technologies [52][53][54].
As a conductive channel substrate, many findings have been obtained through research on the contact resistance between graphene and metal to enhance the efficiency of the field-effect transistor. In addition, the need for high carrier mobility and luxurious functional groups has prompted a great deal of effort, such as the heating-assisted spray system for manufacturing mass production of RGO sheets and the interlayered quantum dots for elevating carrier mobility. Biocompatibility, ease of functionalization, and biocatalysis in the oxygen reduction reaction (ORR) are other outstanding features of graphene-based nanomaterials that are useful in the fabrication of GFETs for ultrasensitive and low-noise cancer detection [54][55][56].
It is important to maintain the stability of the antibody-modified surface during FET measurements. However, unlike other measurement methods, FET measurements use an electrical application on the gate area that can damage the surface. Electrical application may repel antibodies that are immobilized on the surface. Thus, preserving antibody-modified FET properties during measurements improves the reliability of FET antigen detection. Hidshima et al. evaluated three types of FET-modified antibodies to measure stability. The change in FET properties was determined three times in a row by calculating the change in the amount of threshold voltage change (Cv). Modified Cv FET with electrically activated antibodies showed good stability [57].

3.2. Graphene-Based Surface Plasmon Resonance (SPR) Biosensors

SPR-based technologies have proven to be one of the most effective tools for real-time tracking of molecular dynamics, alongside quantitative measurement of numerous biomarkers such as proteins, DNA, entire cells, etc. [58][59][60]. One of the most critical aspects of the SPR biosensors is the surface of the sensor, as it plays a crucial role in the overall efficiency of the sensor. Many experiments have also concentrated on the use of smart sensing layers for more adjustable implementations [61][62][63]. SPR sensing is a powerful, label-free method for investigating noncovalent molecular interactions as a non-invasive method, and in the last two decades, SPR has been widely used in the study of noncovalent protein–DNA, protein–cell, DNA–RNA, DNA–DNA, protein–protein, etc., interaction experiments [64].
SPR biosensors have the potential for extensive implementation in biomarker diagnostics due to the very high surface plasmons’ sensitivity to the alternation in the reflective index (RI) of the dielectric medium. Interaction between the immobilized receptors on the metal surface and the analyte molecules causes a variation in the sensed medium’s refractive index (dielectric), which leads to an alteration in the propagation constant of the surface plasmons. This phenomenon affects the resonance condition of surface plasmons with specific surface plasmon waves (SPW) that have interactions with the incident p-polarized light of the same propagation constant. The energy of light photon transfers to the surface plasmons at the resonance angle, and the reflectance of the light significantly decreasing, results in forming a sharp dip in the SPR curve (reflectance with respect to the incident angle) [65][66]. The measured interacting optical wave can be used to fabricate various types of SPR biosensors with angular, intensity, phase, or wavelength modulation [67].
Phase detection based on coherence is usually very sensitive to mechanical movements in optical components and ambient noise, which leads to phase measurement errors. The stability of the SPR biosensor in the time axis is also required to monitor the biological response over time.

3.3. Graphene-Based Fluorescent Biosensors

Biomolecular imaging and biomarker detection can be performed using fluorescent nanomaterials and labels, a highly sensitive and selective method with adequate spatiotemporal resolution and low cost for application [68]. Numerous fluorimetric diagnostic tests and fluorescent-based biosensors focused on biocatalyst behavior have been documented using organic dyes, inorganic semiconductor quantum dots (QDs), and carbon nanomaterials as fluorimetric indicators [69][70]. The operation of these biosensors is based on the fluorescence phenomenon that happens when a fluorophore or fluorescently labeled molecule absorbs the corresponding electromagnetic radiation. According to the signal generating technique in fluorescent biosensors, they are categorized into four types, including FRET (Förster Resonance Energy Transfer), FLIM (Fluorescence Lifetime Imaging), FCS (Fluorescence Correlation Spectroscopy), and FI (change in fluorescence intensity) [71]. The main advantages of fluorescence-based biosensors are extreme sensitivity, they are minimally invasive or non-invasive, they have the ability to utilize fluorescence intensity and fluorescence lifetime, and they provide the structure and microenvironment of molecules [72][73][74].

3.4. Graphene-Based Electrochemical Biosensors

Electrochemical sensing essentially needs a reference electrode, a counter (auxiliary electrode), and a working electrode, sometimes termed as the sensing or redox electrode. The reference electrode, usually produced from Ag/AgCl, is held at length from the reaction site to preserve a specified and consistent potential [75]. The working electrode acts as a transduction component in the biochemical reaction, while the counter electrode links the electrolytic solution so that the current can be delivered to the working electrode. These electrodes ought to be conductive and have chemical stability [76]. According to the Institute of Clinical and Laboratory Standards (CLSI; EP05-A3, EP24-A2, EP25-A), a change factor (CV) of less than 10% is required for reproducibility, accuracy, and stability. Therefore, further improvements in biosensors, especially in the case of electrodes and intermediates, are two components that determine the reproducibility, accuracy, and stability of all electrochemical biosensors [77]. Platinum, gold, carbon (e.g., graphite), and silicone derivatives are thus widely utilized based on the analyte [76].

3.5. Graphene-Based Surface-Enhanced Raman Scattering (SERS) Biosensors

In Raman scattering, photons inelastically lose (Stokes) or gain (anti-Stokes) energy because of molecular vibrational events and represent information about the molecular structure enabling in situ and real-time detection [78]. Surface-enhanced Raman scattering (SERS) is a subset of Raman scattering, a widely used sensing technique in which when the molecules are adsorbed on corrugated metal surfaces such as silver or gold nanoparticles, inelastic light scattering by molecules is greatly improved [79][80]. By way of plasmonic nanostructures, it provides a million-fold improvement, making the efficiency of detection down to the level of single molecules. Two different pathways, namely electromagnetic enhancement and chemical enhancement, accomplish SERS enhancement [81].

3.6. Graphene-Based Electrochemiluminescent Biosensors

Electroluminescence or electrogenerated chemiluminescence (ECL) is the shared area of electrochemistry and chemiluminescence (CL). In this mechanism, by applying a potential on the electrode’s surface, the electrochemical energy is being converted to radiative energy [82][83][84]. This phenomenon can happen by utilizing species that undergo electron transfer reaction to form an excited state and, after that, produce light when molecules return to the ground state [85][86]. Therefore, ECL does not need external light sources, so problems of light scattering inherent in photoluminescence (PL) can be avoided [84]. In the ECL method, the electrochemical reactions take place through the interplay of the luminophore and a co-reactant molecule by applying only one single potential step. Two approaches to the co-reactant ECL mechanism are available. The first one is the oxidative reaction, in which a potent reducing radical agent is produced by oxidation of the co-reactant in a homogeneous follow-up reaction. This radical can reduce the oxidized luminophore, so the luminophore becomes excited and emits light. The second one is reductive oxidation, in which the reduction takes place straightforwardly [87].

References

  1. Lazcka, O.; Del Campo, F.J.; Munoz, F.X. Pathogen detection: A perspective of traditional methods and biosensors. Biosens. Bioelectron. 2007, 22, 1205–1217.
  2. Tetyana, P.; Shumbula, P.M.; Njengele-Tetyana, Z. Biosensors: Design, Development and Applications. In Nanopores; IntechOpen: London, UK, 2021.
  3. Alharthi, S.D.; Bijukumar, D.; Prasad, S.; Khan, A.M.; Mathew, M.T. Evolution in Biosensors for Cancers Biomarkers Detection: A Review. J. Bio-Tribo-Corros. 2021, 7, 42.
  4. Mehrotra, P. Biosensors and their applications—A review. J. Oral Biol. Craniofacial Res. 2016, 6, 153–159.
  5. Pourmadadi, M.; Yazdian, F.; Hojjati, S.; Khosravi-Darani, K. Detection of Microorganisms Using Graphene-Based Nanobiosensors. Food Technol. Biotechnol. 2021, 59, 496–506.
  6. Tothill, I.E. Biosensors for cancer markers diagnosis. Semin. Cell Dev. Biol. 2009, 20, 55–62.
  7. Ferrari, M. BioMEMS and Biomedical Nanotechnology: Volume IV: Biomolecular Sensing, Processing and Analysis; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2007; Volume 4.
  8. Tabish, T.A.; Hayat, H.; Abbas, A.; Narayan, R.J. Graphene quantum dot-based electrochemical biosensing for early cancer detection. Curr. Opin. Electrochem. 2021, 30, 100786.
  9. Zhang, Z.; Li, Q.; Du, X.; Liu, M. Application of electrochemical biosensors in tumor cell detection. Thorac. Cancer 2020, 11, 840–850.
  10. Pourmadadi, M.; Shayeh, J.S.; Arjmand, S.; Omidi, M.; Fatemi, F. An electrochemical sandwich immunosensor of vascular endothelial growth factor based on reduced graphene oxide/gold nanoparticle composites. Microchem. J. 2020, 159, 105476.
  11. Pourmadadi, M.; Shayeh, J.S.; Omidi, M.; Yazdian, F.; Alebouyeh, M.; Tayebi, L. A glassy carbon electrode modified with reduced graphene oxide and gold nanoparticles for electrochemical aptasensing of lipopolysaccharides from Escherichia coli bacteria. Microchim. Acta 2019, 186, 787.
  12. Gunawardana, C.G.; Diamandis, E.P. High throughput proteomic strategies for identifying tumour-associated antigens. Cancer Lett. 2007, 249, 110–119.
  13. Martinelli, C.; Pucci, C.; Ciofani, G. Nanostructured carriers as innovative tools for cancer diagnosis and therapy. APL Bioeng. 2019, 3, 011502.
  14. Nair, P.R.; Alam, M.A. Screening-limited response of nanobiosensors. Nano Lett. 2008, 8, 1281–1285.
  15. Madrid, R.E.; Chehín, R.; Chen, T.-H.; Guiseppi-Elie, A. Biosensors and Nanobiosensors. In Further Understanding of the Human Machine: The Road to Bioengineering; World Scientific: Singapore, 2017; pp. 391–462.
  16. Bellan, L.M.; Wu, D.; Langer, R.S. Current trends in nanobiosensor technology. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2011, 3, 229–246.
  17. Ghiasi, B.; Mehdipour, G.; Safari, N.; Behboudi, H.; Hashemi, M.; Omidi, M.; Sefidbakht, Y.; Yadegari, A.; Hamblin, M.R. Theranostic applications of stimulus-responsive systems based on carbon dots. Int. J. Polym. Mater. Polym. Biomater. 2019, 70, 117–130.
  18. Nikoleli, G.-P.; Nikolelis, D.P.; Siontorou, C.; Karapetis, S.; Bratakou, S.; Tzamtzis, N. Nanobiosensors Based on Graphene Electrodes: Recent Trends and Future Applications. In Applications of Nanomaterials; Elsevier: Amsterdam, The Netherlands, 2018; pp. 161–177.
  19. Georgakilas, V.; Otyepka, M.; Bourlinos, A.B.; Chandra, V.; Kim, N.; Kemp, K.C.; Hobza, P.; Zboril, R.; Kim, K.S. Functionalization of graphene: Covalent and non-covalent approaches, derivatives and applications. Chem. Rev. 2012, 112, 6156–6214.
  20. Jin, X.; Feng, C.; Ponnamma, D.; Yi, Z.; Parameswaranpillai, J.; Thomas, S.; Salim, N. Review on exploration of graphene in the design and engineering of smart sensors, actuators and soft robotics. Chem. Eng. J. Adv. 2020, 4, 100034.
  21. Heyrovska, R. Atomic structures of graphene, benzene and methane with bond lengths as sums of the single, double and resonance bond radii of carbon. arXiv 2008, arXiv:0804.4086.
  22. Vorontsov, A.V.; Tretyakov, E.V. Determination of graphene’s edge energy using hexagonal graphene quantum dots and PM7 method. Phys. Chem. Chem. Phys. 2018, 20, 14740–14752.
  23. Lee, C.; Wei, X.; Kysar, J.W.; Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science 2008, 321, 385–388.
  24. Malhotra, B.D.; Srivastava, S.; Ali, M.A.; Singh, C. Nanomaterial-based biosensors for food toxin detection. Appl. Biochem. Biotechnol. 2014, 174, 880–896.
  25. Banerjee, A.N. Graphene and its derivatives as biomedical materials: Future prospects and challenges. Interface Focus 2018, 8, 20170056.
  26. Donarelli, M.; Ottaviano, L. 2D materials for gas sensing applications: A review on graphene oxide, MoS2, WS2 and phosphorene. Sensors 2018, 18, 3638.
  27. Kalajahi, S.T.; Mofradnia, S.R.; Yazdian, F.; Rasekh, B.; Neshati, J.; Taghavi, L.; Pourmadadi, M.; Haghirosadat, B.F. Inhibition performances of graphene oxide/silver nanostructure for the microbial corrosion: Molecular dynamic simulation study. Environ. Sci. Pollut. Res. 2022, 1–14.
  28. Tian, P.; Tang, L.; Teng, K.; Lau, S. Graphene quantum dots from chemistry to applications. Mater. Today Chem. 2018, 10, 221–258.
  29. Pourmadadi, M.; Nouralishahi, A.; Shalbaf, M.; Shayeh, J.S.; Nouralishahi, A. An electrochemical aptasensor for detection of Prostate specific antigen based on carbon quantum dots-gold nanoparticles. Biotechnol. Appl. Biochem. 2022.
  30. Behboudi, H.; Mehdipour, G.; Safari, N.; Pourmadadi, M.; Saei, A.; Omidi, M.; Tayebi, L.; Rahmandoust, M. Carbon quantum dots in nanobiotechnology. In Nanomaterials for Advanced Biological Applications; Springer: Berlin/Heidelberg, Germany, 2019; pp. 145–179.
  31. Abolghasemzade, S.; Pourmadadi, M.; Rashedi, H.; Yazdian, F.; Kianbakht, S.; Navaei-Nigjeh, M. PVA based nanofiber containing CQDs modified with silica NPs and silk fibroin accelerates wound healing in a rat model. J. Mater. Chem. B 2021, 9, 658–676.
  32. Malmir, S.; Karbalaei, A.; Pourmadadi, M.; Hamedi, J.; Yazdian, F.; Navaee, M. Antibacterial properties of a bacterial cellulose CQD-TiO2 nanocomposite. Carbohydr. Polym. 2020, 234, 115835.
  33. Vermisoglou, E.; Panáček, D.; Jayaramulu, K.; Pykal, M.; Frébort, I.; Kolář, M.; Hajdúch, M.; Zbořil, R.; Otyepka, M. Human virus detection with graphene-based materials. Biosens. Bioelectron. 2020, 166, 112436.
  34. Huang, X.; Yin, Z.; Wu, S.; Qi, X.; He, Q.; Zhang, Q.; Yan, Q.; Boey, F.; Zhang, H. Graphene-based materials: Synthesis, characterization, properties, and applications. Small 2011, 7, 1876–1902.
  35. Suvarnaphaet, P.; Pechprasarn, S. Graphene-based materials for biosensors: A review. Sensors 2017, 17, 2161.
  36. Li, D.; Zhang, W.; Yu, X.; Wang, Z.; Su, Z.; Wei, G. When biomolecules meet graphene: From molecular level interactions to material design and applications. Nanoscale 2016, 8, 19491–19509.
  37. Chua, C.K.; Pumera, M. Chemical reduction of graphene oxide: A synthetic chemistry viewpoint. Chem. Soc. Rev. 2014, 43, 291–312.
  38. Yildiz, G.; Bolton-Warberg, M.; Awaja, F. Graphene and graphene oxide for bio-sensing: General properties and the effects of graphene ripples. Acta Biomater. 2021, 131, 62–79.
  39. Zamani, M.; Pourmadadi, M.; Ebrahimi, S.S.; Yazdian, F.; Shayeh, J.S. A novel labeled and label-free dual electrochemical detection of endotoxin based on aptamer-conjugated magnetic reduced graphene oxide-gold nanocomposite. J. Electroanal. Chem. 2022, 908, 116116.
  40. Wang, Z.; Dai, Z. Carbon nanomaterial-based electrochemical biosensors: An overview. Nanoscale 2015, 7, 6420–6431.
  41. Cheng, C.; Li, S.; Thomas, A.; Kotov, N.A.; Haag, R. Functional graphene nanomaterials based architectures: Biointeractions, fabrications, and emerging biological applications. Chem. Rev. 2017, 117, 1826–1914.
  42. Lawal, A.T. Progress in utilisation of graphene for electrochemical biosensors. Biosens. Bioelectron. 2018, 106, 149–178.
  43. Yang, W.; Deng, X.; Huang, W.; Qing, X.; Shao, Z. The physicochemical properties of graphene nanocomposites influence the anticancer effect. J. Oncol. 2019, 2019, 7254534.
  44. Nanda, S.S.; Papaefthymiou, G.C.; Yi, D.K. Functionalization of graphene oxide and its biomedical applications. Crit. Rev. Solid State Mater. Sci. 2015, 40, 291–315.
  45. Fiorani, A.; Merino, J.P.; Zanut, A.; Criado, A.; Valenti, G.; Prato, M.; Paolucci, F. Advanced carbon nanomaterials for electrochemiluminescent biosensor applications. Curr. Opin. Electrochem. 2019, 16, 66–74.
  46. Pathak, S.; Estrela, P. Field-effect transistors: Current advances and challenges in bringing them to point-of-care. In Nanobiosensors and Nanobioanalyses; Springer: Berlin/Heidelberg, Germany, 2015; pp. 353–371.
  47. Syu, Y.-C.; Hsu, W.-E.; Lin, C.-T. Review—Field-Effect transistor biosensing: Devices and clinical applications. ECS J. Solid State Sci. Technol. 2018, 7, Q3196.
  48. Sze, S.M.; Ng, K.K. Physics of Semiconductor Devices; John Wiley & Sons: New York, NY, USA, 2006.
  49. Kaisti, M. Detection principles of biological and chemical FET sensors. Biosens. Bioelectron. 2017, 98, 437–448.
  50. Wadhera, T.; Kakkar, D.; Wadhwa, G.; Raj, B. Recent advances and progress in development of the field effect transistor biosensor: A review. J. Electron. Mater. 2019, 48, 7635–7646.
  51. Sharma, S.; Prakash, V.; Mehta, S. Graphene/silver nanocomposites-potential electron mediators for proliferation in electrochemical sensing and SERS activity. TrAC Trends Anal. Chem. 2017, 86, 155–171.
  52. Fahim-Al-Fattah, M.; Rahman, M.T.; Islam, M.S.; Bhuiyan, A.G. A Study on Theoretical Performance of Graphene FET using Analytical Approach with Reference to High Cutoff Frequency. Int. J. Nanosci. 2016, 15, 1640001.
  53. Aspermair, P.; Mishyn, V.; Bintinger, J.; Happy, H.; Bagga, K.; Subramanian, P.; Knoll, W.; Boukherroub, R.; Szunerits, S. Reduced graphene oxide–based field effect transistors for the detection of E7 protein of human papillomavirus in saliva. Anal. Bioanal. Chem. 2020, 413, 779–787.
  54. Hu, S.; Jia, Y. Function of Tetra (4-Aminophenyl) Porphyrin in Altering the Electronic Performances of Reduced Graphene Oxide-Based Field Effect Transistor. Molecules 2019, 24, 3960.
  55. Tian, M.; Xu, S.; Zhang, J.; Wang, X.; Li, Z.; Liu, H.; Song, R.; Yu, Z.; Wang, J. RNA detection based on graphene field-effect transistor biosensor. Adv. Condens. Matter Phys. 2018, 2018, 8146765.
  56. Seo, G.; Lee, G.; Kim, M.J.; Baek, S.-H.; Choi, M.; Ku, K.B.; Lee, C.-S.; Jun, S.; Park, D.; Kim, H.G. Rapid detection of COVID-19 causative virus (SARS-CoV-2) in human nasopharyngeal swab specimens using field-effect transistor-based biosensor. ACS Nano 2020, 14, 5135–5142.
  57. Hideshima, S.; Sato, R.; Kuroiwa, S.; Osaka, T. Fabrication of stable antibody-modified field effect transistors using electrical activation of Schiff base cross-linkages for tumor marker detection. Biosens. Bioelectron. 2011, 26, 2419–2425.
  58. Nie, W.; Wang, Q.; Zou, L.; Zheng, Y.; Liu, X.; Yang, X.; Wang, K. Low-fouling surface plasmon resonance sensor for highly sensitive detection of microRNA in a complex matrix based on the DNA tetrahedron. Anal. Chem. 2018, 90, 12584–12591.
  59. Daems, D.; Pfeifer, W.; Rutten, I.; Saccà, B.; Spasic, D.; Lammertyn, J. Three-dimensional DNA origami as programmable anchoring points for bioreceptors in fiber optic surface plasmon resonance biosensing. ACS Appl. Mater. Interfaces 2018, 10, 23539–23547.
  60. Bian, S.; Lu, J.; Delport, F.; Vermeire, S.; Spasic, D.; Lammertyn, J.; Gils, A. Development and validation of an optical biosensor for rapid monitoring of adalimumab in serum of patients with Crohn’s disease. Drug Test. Anal. 2018, 10, 592–596.
  61. Wasserberg, D.; Cabanas-Danés, J.; Prangsma, J.; O’Mahony, S.; Cazade, P.-A.; Tromp, E.; Blum, C.; Thompson, D.; Huskens, J.; Subramaniam, V. Controlling protein surface orientation by strategic placement of oligo-histidine tags. ACS Nano 2017, 11, 9068–9083.
  62. Chen, L.; Hwang, E.; Zhang, J. Fluorescent nanobiosensors for sensing glucose. Sensors 2018, 18, 1440.
  63. Zeng, S.; Hu, S.; Xia, J.; Anderson, T.; Dinh, X.-Q.; Meng, X.-M.; Coquet, P.; Yong, K.-T. Graphene—MoS2 hybrid nanostructures enhanced surface plasmon resonance biosensors. Sens. Actuators B Chem. 2015, 207, 801–810.
  64. Bakhtiar, R. Surface plasmon resonance spectroscopy: A versatile technique in a biochemist’s toolbox. J. Chem. Educ. 2013, 90, 203–209.
  65. Maurya, J.; Prajapati, Y. A comparative study of different metal and prism in the surface plasmon resonance biosensor having MoS2-graphene. Opt. Quantum Electron. 2016, 48, 280.
  66. Nesterenko, D.V.; Sekkat, Z. Resolution estimation of the Au, Ag, Cu, and Al single-and double-layer surface plasmon sensors in the ultraviolet, visible, and infrared regions. Plasmonics 2013, 8, 1585–1595.
  67. Wijaya, E.; Lenaerts, C.; Maricot, S.; Hastanin, J.; Habraken, S.; Vilcot, J.-P.; Boukherroub, R.; Szunerits, S. Surface plasmon resonance-based biosensors: From the development of different SPR structures to novel surface functionalization strategies. Curr. Opin. Solid State Mater. Sci. 2011, 15, 208–224.
  68. Kaczmarski, J.A.; Mitchell, J.A.; Spence, M.A.; Vongsouthi, V.; Jackson, C.J. Structural and evolutionary approaches to the design and optimization of fluorescence-based small molecule biosensors. Curr. Opin. Struct. Biol. 2019, 57, 31–38.
  69. Bai, Y.; Xu, T.; Zhang, X. Graphene-Based Biosensors for Detection of Biomarkers. Micromachines 2020, 11, 60.
  70. Behboodi, H.; Pourmadadi, M.; Omidi, M.; Rahmandoust, M.; Siadat, S.O.R.; Shayeh, J.S. Cu-CDs as dual optical and electrochemical nanosensor for βME detection. Surf. Interfaces 2021, 29, 101710.
  71. Nawrot, W.; Drzozga, K.; Baluta, S.; Cabaj, J.; Malecha, K. A fluorescent biosensors for detection vital body fluids’ agents. Sensors 2018, 18, 2357.
  72. Pickup, J.C.; Hussain, F.; Evans, N.D.; Rolinski, O.J.; Birch, D.J. Fluorescence-based glucose sensors. Biosens. Bioelectron. 2005, 20, 2555–2565.
  73. Klonoff, D.C. Overview of fluorescence glucose sensing: A technology with a bright future. J. Diabetes Sci. Technol. 2012, 6, 1242–1250.
  74. He, X.-P.; Hu, X.-L.; James, T.D.; Yoon, J.; Tian, H. Multiplexed photoluminescent sensors: Towards improved disease diagnostics. Chem. Soc. Rev. 2017, 46, 6687–6696.
  75. Mehdipour, G.; Shabani Shayeh, J.; Omidi, M.; Pour Madadi, M.; Yazdian, F.; Tayebi, L. An electrochemical aptasensor for detection of prostate-specific antigen using reduced graphene gold nanocomposite and Cu/carbon quantum dots. Biotechnol. Appl. Biochem. 2021.
  76. Grieshaber, D.; MacKenzie, R.; Vörös, J.; Reimhult, E. Electrochemical biosensors-sensor principles and architectures. Sensors 2008, 8, 1400–1458.
  77. Chen, L.-C.; Wang, E.; Tai, C.-S.; Chiu, Y.-C.; Li, C.-W.; Lin, Y.-R.; Lee, T.-H.; Huang, C.-W.; Chen, J.-C.; Chen, W.L. Improving the reproducibility, accuracy, and stability of an electrochemical biosensor platform for point-of-care use. Biosens. Bioelectron. 2020, 155, 112111.
  78. Li, P.; Long, F.; Chen, W.; Chen, J.; Chu, P.K.; Wang, H. Fundamentals and applications of surface-enhanced Raman spectroscopy–based biosensors. Curr. Opin. Biomed. Eng. 2020, 13, 51–59.
  79. Zhao, Y.; Li, X.; Liu, Y.; Zhang, L.; Wang, F.; Lu, Y. High performance surface-enhanced Raman scattering sensing based on Au nanoparticle-monolayer graphene-Ag nanostar array hybrid system. Sens. Actuators B Chem. 2017, 247, 850–857.
  80. Langer, J.; de Aberasturi, J.D.; Aizpurua, J.; Alvarez-Puebla, R.A.; Auguié, B.; Baumberg, J.J.; Bazan, G.C.; Bell, S.E.; Boisen, A.; Brolo, A.G. Present and future of surface-enhanced Raman scattering. ACS Nano 2019, 14, 28–117.
  81. Ding, S.-Y.; You, E.-M.; Tian, Z.-Q.; Moskovits, M. Electromagnetic theories of surface-enhanced Raman spectroscopy. Chem. Soc. Rev. 2017, 46, 4042–4076.
  82. Shen, M.; Rodriguez-Lopez, J.; Huang, J.; Liu, Q.; Zhu, X.-H.; Bard, A.J. Electrochemistry and Electrogenerated Chemiluminescence of Dithienylbenzothiadiazole Derivative. Differential Reactivity of Donor and Acceptor Groups and Simulations of Radical Cation− Anion and Dication− Radical Anion Annihilations. J. Am. Chem. Soc. 2010, 132, 13453–13461.
  83. Hu, L.; Xu, G. Applications and trends in electrochemiluminescence. Chem. Soc. Rev. 2010, 39, 3275–3304.
  84. Ding, C.; Zhang, W.; Wang, W.; Chen, Y.; Li, X. Amplification strategies using electrochemiluminescence biosensors for the detection of DNA, bioactive molecules and cancer biomarkers. TrAC Trends Anal. Chem. 2015, 65, 137–150.
  85. Richter, M.M. Electrochemiluminescence (ecl). Chem. Rev. 2004, 104, 3003–3036.
  86. Zhou, Y.; Yin, H.; Zhao, W.-W.; Ai, S. Electrochemical, electrochemiluminescent and photoelectrochemical bioanalysis of epigenetic modifiers: A comprehensive review. Coord. Chem. Rev. 2020, 424, 213519.
  87. Valenti, G.; Rampazzo, E.; Kesarkar, S.; Genovese, D.; Fiorani, A.; Zanut, A.; Palomba, F.; Marcaccio, M.; Paolucci, F.; Prodi, L. Electrogenerated chemiluminescence from metal complexes-based nanoparticles for highly sensitive sensors applications. Coord. Chem. Rev. 2018, 367, 65–81.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , ,
View Times: 678
Revisions: 3 times (View History)
Update Date: 21 Jul 2022
1000/1000