Graphene/Polymer Nanocomposites for Human Health Monitoring: History
Please note this is an old version of this entry, which may differ significantly from the current revision.

Graphene/polymer nanocomposites (GPNs) are largely explored in the development of sensing devices to monitor human health parameters, due to the excellent electrical and mechanical properties of the graphene filler combined with the chemical versatility of the surrounding polymer matrix.

  • graphene
  • polymers
  • sensors
  • nanocomposites
  • human health monitoring

1. Graphene/Polymer Nanocomposites: Fabrication and Properties

Graphene-based nanocomposites with polymer matrix are commonly fabricated following three different methods: solution blending, in situ polymerization and melt mixing [1][2]. The most largely used technique is solution blending, which involves the solubilization of the polymer in a suitable solvent and the mixing with graphene to form a homogeneous dispersion. Generally, polymers such as polystyrene, polycarbonate, polyacrylamide, polyimides and poly(methyl methacrylate) are mixed with graphene oxide [3][4][5], which can be previously functionalized with isocyanates, alkylamine or alkyl-chlorosilanes in order to improve its dispersibility in organic solvents.

The fabrication of GPNs by in situ polymerization is based on the polymerization of the matrix in the presence of the selected filler, starting from a mixture of monomer and reinforcement [1]. Typically, this approach allows to obtain a good grade of dispersion of graphene-based nanofillers avoiding their previous exfoliation. In the melt mixing technique, the filler is dispersed in the polymer matrix exploiting high temperatures and shear forces [6]. The polymer phase is brought to melting at high temperatures, thus facilitating the dispersion or intercalation of the graphene oxide nanoplatelets without the use of organic, often toxic, solvents.

The properties of the GPN nanocomposites are strictly related to the spatial distribution and alignment of the graphene nanofiller, and to its interfacial adhesion with the polymer phase. In particular, GPNs with low loadings of functionalized graphene sheets generally exhibit a shift in the glass transition temperature [4], if compared with the value of the uncharged polymer. This behavior can be ascribed to a reduced mobility of the polymer chains at the interface between the filler and the matrix [7][8]. Therefore, the effect of constraint applied on the chains can directly induce an increase in glass transition temperature [9].

In terms of thermal conductivity, the performance of GPNs can be evaluated referring to the 2D geometry of the graphene fillers. These are characterized by a lower interfacial thermal resistance that provides a higher thermal conductivity to the host polymer matrix [10][11]. Nevertheless, the 2D structure can be source of anisotropy in the nanocomposites arrangement, for which the in-plane thermal conductivity results as much as ten times higher than the cross-plane conductivity [12]. This is typically evaluated following the percolation theory, therefore considering phonons as the main mode for thermal conduction in polymers. Covalent bonding between the filler and the polymer matrix can reduce phonon scattering at the interface leading to an overall enhancement of the GPN thermal conductivity [13].

The electrical conductivity behavior of GPNs can be analyzed considering the influence of different factors and their overall effect. In particular, the characteristics of the specific graphene-based filler, such as its aspect ratio and morphology, as well as its inter-sheet junction, can affect the electrical performances of GPNs [14]. In the same way, processing and dispersion, and the related state of aggregation and alignment of the nanoparticles concur to determine the electrical behavior of the resulting nanocomposites [15]. Several theoretical models and experiments were aimed to assess the role of nanofiller shape, geometry and state of dispersion on the percolation threshold of graphene/polymer nanocomposites [16][17].

As mentioned above, the overall performance of polymer-based nanocomposite materials can be related to the quality and stability of the polymer/nanofiller interphase region. Typically, the physical and mechanical properties and the chemical composition of this region are different from those of the bulk polymer matrix [18][19]. In the case of an interphase stiffer than the surrounding polymer, this can result in higher overall stiffness and strength of the composite, but with lower resistance to fracture [20]. The interphase properties can affect the mechanical behavior of the nanocomposites also depending on the morphology and size of this region. In fact, several studies show that its thickness can be tailored with the aim to achieve both higher strength and improved toughness of the resulting nanocomposites [21][22][23]. Force-modulation atomic force microscopy (AFM) and nanoindentation are common techniques that are used to investigate the interphase and its properties [18][24]. In particular, AFM phase imaging is currently considered a useful tool to evaluate the thickness and the relative stiffness of the interphase, since it involves much lower interaction forces between the probe and the sample than force modulation or nanoindentation [20]. The arrangement of graphene-based fillers inside the polymer matrix is also investigated in order to assess the state of dispersion at the microstructural level and its impact on the nanocomposite properties. Results reveal that graphene-based fillers, such as graphene oxide or graphene nanoplatelets, can arrange differently in the host polymer, originating structural states that can be classified as stacked, intercalated or exfoliated [1].

The intercalated state can be considered a particular stacked structure, which is characterized by a greater interlayer spacing, but within a few nanometers [25]. Generally, in the exfoliated structure, graphene nanoplatelets have the largest interfacial contact with the polymer matrix, and this allows to improve the performances of the composites in different ways. Due to the interactions with the matrix, the exfoliated phase can exhibit a curved shape. In this case, the rumpled shape assumed by the filler can result in a mechanical interlocking acting as a possible mechanism of strengthening. The compatibility between the host polymer and the nanoplatelets is one of the major factors determining the filler morphology in the matrix: the nanoplatelets are characterized by a more extended conformation for high polymer/filler affinity or, conversely, a crumpled conformation when the affinity decreases [26]. Finally, the processing method used to fabricate the nanocomposite also affects its microstructure to a great extent: solution mixing or in situ polymerization generally induce an exfoliated and randomly oriented status of the nanoplatelets, whereas from the melt mixing technique a more oriented and intercalated or stacked structure of the nanoplatelets is generated [27].

 

2. Graphene/Polymer Nanocomposites: Applications in Human Health Monitoring

Many sensors based on nanocomposite materials find applications in the monitoring of human health parameters [28][29]. In this perspective, the sensing devices need to be comfortable to wear, biocompatible, and lightweight [30][31][32]. They need to interface with human body, showing at the same time high selectivity and sensitivity to detect and quantify specific signals or analytes.

Graphene, graphene oxide and chemically modified graphene are widely employed to fabricate nanocomposites suitable for detecting biological analytes, such as uric acid and ascorbic acid [33], hydroquinone and catechol [34], and nucleic bases [35][36]. The presence of functional groups on the nanocomposite surface is fundamental to create hydrogen bonds with the analytes, so the strength of these bonds and the distance between the interaction sites and the reaction center make possible the discrimination of the analytes.

DNA molecules can be immobilized on graphene surface by physical adsorption or chemical binding, thus creating sensitive platforms where each binding event with the analyte can be detected through the changes of the electric or electrochemical properties of these platforms [37][38]. Noncovalent interactions can be promoted through the physical adsorption, involving π-π stacking interactions between the DNA nucleobases and the aromatic surface of graphene. In particular, in the case of single-stranded DNA (ssDNA), stable aqueous dispersions of graphene/DNA can be obtained, without traces of sedimentation for months [39]. Double-stranded DNA (dsDNA) is also used as dispersing agent for graphene nanoplatelets. However, in this case less stable aqueous solutions are obtained due to the weaker hydrophobic interactions arising from the base pairing of the nucleobases. Nevertheless, the graphene/dsDNA affinity can be significantly enhanced by further functionalizing graphene oxide with polar groups, which are able to establish electrostatic interactions with the DNA bases. The immobilization of DNA on graphene through covalent bonds is generally carried out after functionalizing the DNA with amino groups, which are able to interact with the graphene oxide surface via carbodiimide chemistry [40]. In particular, amine-terminated ssDNA can be linked to the surface of graphene oxide directly or through the involvement of specific molecules that act as carriers.

Single-stranded DNA was covalently immobilized on a polyaniline/graphene (PAN/GN) nanocomposite, which was applied onto a glassy carbon electrode (GCE) and used for HIV-1 gene detection [41]. In particular, the negatively-charged phosphate backbone of the HIV-1 binds to the sensitive surface via π-π stacking interactions. The hybridization between the ssDNA probe and the target HIV-1 generates double-stranded DNA (dsDNA), which induces an increase of the electron transfer resistance that can be correlated with the concentration of the gene. The sensitivity and the selectivity of this nanocomposite were tested, and a low detection limit of 1.0 × 10−16 M for the target HIV-1 was measured. 

The development of new composite materials using biomolecules, such as enzymes, has allowed to extend even more the field of sensing applications in medical diagnosis and bio-industrial analysis. Several sensitive nanocomposites have been realized using natural polymers as matrix, such as gelatin, alginate and chitosan, due to their intrinsic biodegradability and biocompatibility that make it suitable for biomedical applications [42]. In particular, chitosan was combined with graphene to develop nanocomposite materials with sensing properties useful for monitoring human health [43][44][45].

Xie et al. developed an immunosensor based on graphene and chitosan-modified screen-printed carbon electrode (SPCE) [46]. The phospho-p53 capture antibody was adsorbed on the surface of the graphene-chitosan/SPCE. A sandwich immunocomplex was formed between the targeted phospho-p5315 antigen, the phospho-p53 capture antibody, the antigen, and biotinylated phospho-p5315 detection antibody, which was previously marked with horseradish peroxidase (HRP). The high surface area of graphene allowed to immobilize a large amount of capture antibody, increasing the sensitivity of this nanocomposite immunosensor.

Nanocomposite films based on glucose oxidase (GOD), platinum (Pt), functional graphene sheets (FGS) and chitosan were developed for glucose sensing [47]. The electrocatalytic action of FGS and Pt nanoparticles towards hydrogen peroxide (H2O2) was exploited to obtain a sensitive biosensor with a detection limit of 0.6 μM of glucose. The performance of this type of sensor can be ascribed to the large surface area and the fast electron transfer of graphene and Pt nanoparticles. This sensor showed good reproducibility and long-term stability, with a negligible response to other compounds such as ascorbic acid and uric acid. The GOD/Pt/FGS/chitosan sensitive nanocomposite can be useful for both clinical and home-care devices for a rapid monitoring of glucose.

Glucose sensing was also performed with composite films made of graphene, chitosan and uric acid, which were deposited onto glassy carbon electrodes [44]. A molecularly imprinted electrochemical sensor was obtained, and its sensitivity mechanism was analyzed by electrochemical impedance spectroscopy and chronocoulometric methods. A comparison between graphene-doped and undoped sensors was carried out, with results demonstrating an improvement in terms of sensitivity due to the high surface area and good electronic conduction of graphene.

Sensitive films based on EDTA-modified reduced graphene (EDTA-RG) and Nafion were fabricated and tested as dopamine detectors [48]. Graphene was chemically modified by silanization using N-(trimethoxysilylpropyl) ethylenediamine triacetic acid (EDTA-silane). The selectivity was investigated by using dopamine and ascorbic acid. Experimental tests demonstrated that the sulfuric groups of Nafion and the carboxylic groups of EDTA-RG interfere with the diffusion of ascorbic acid, thus enabling the selective detection of dopamine.

More recently, biosensing has seen advances towards more complex structures that are able to enhance the overall sensitivity of the detecting surface. A sensing composite material was realized using fractal nanoplatinum with a cauliflower-like morphology, which was developed on a reduced graphene oxide paper [49]. Platinum was electrodeposited on the graphene-nanocellulose sheets using pulsed sonoelectrodeposition. As a result, a conductive nanocomposite paper with a high electroactive surface was obtained and then functionalized using glucose oxidase (via chitosan encapsulation) or RNA aptamer (via covalent linking). In this way, the material sensitivity towards glucose or Escherichia coli bacteria can be activated. Depending on the type of the enzyme selected, good performances in terms of sensitivity and response times were obtained.

This entry is adapted from the peer-reviewed paper 10.3390/polym14051030

References

  1. Jeffrey R. Potts; Daniel R. Dreyer; Christopher W. Bielawski; Rodney S. Ruoff; Graphene-based polymer nanocomposites. Polymer 2011, 52, 5-25, 10.1016/j.polymer.2010.11.042.
  2. Dimitrios Papageorgiou; Ian A. Kinloch; Robert J. Young; Graphene/elastomer nanocomposites. Carbon 2015, 95, 460-484, 10.1016/j.carbon.2015.08.055.
  3. Dan Chen; Hong Zhu; Tianxi Liu; In Situ Thermal Preparation of Polyimide Nanocomposite Films Containing Functionalized Graphene Sheets. ACS Applied Materials & Interfaces 2010, 2, 3702-3708, 10.1021/am1008437.
  4. T. Ramanathan; Ahmed Abdala; S. Stankovich; D. A. Dikin; M. Herrera-Alonso; R. D. Piner; D. H. Adamson; Hannes Schniepp; X. Chen; R. S. Ruoff; et al. Functionalized graphene sheets for polymer nanocomposites. Nature Nanotechnology 2008, 3, 327-331, 10.1038/nnano.2008.96.
  5. Sasha Stankovich; Dmitriy A. Dikin; Geoffrey H. B. Dommett; Kevin M. Kohlhaas; Eric J. Zimney; Eric A. Stach; Richard D. Piner; Sonbinh Nguyen; Rodney S. Ruoff; Graphene-based composite materials. Nature 2006, 442, 282-286, 10.1038/nature04969.
  6. Josphat Phiri; Patrick Gane; Thad C. Maloney; General overview of graphene: Production, properties and application in polymer composites. Materials Science and Engineering: B 2017, 215, 9-28, 10.1016/j.mseb.2016.10.004.
  7. Amitabh Bansal; Hoichang Yang; Chunzhao Li; Kilwon Cho; Brian C. Benicewicz; Sanat K. Kumar; Linda S. Schadler; Quantitative equivalence between polymer nanocomposites and thin polymer films. Nature Materials 2005, 4, 693-698, 10.1038/nmat1447.
  8. Rodney D. Priestley; Christopher J. Ellison; Linda J. Broadbelt; John M. Torkelson; Structural Relaxation of Polymer Glasses at Surfaces, Interfaces, and In Between. Science 2005, 309, 456-459, 10.1126/science.1112217.
  9. Seyed Esmaeil Zakiyan; Hamed Azizi; Ismaeil Ghasemi; Influence of chain mobility on rheological, dielectric and electromagnetic interference shielding properties of poly methyl-methacrylate composites filled with graphene and carbon nanotube. Composites Science and Technology 2017, 142, 10-19, 10.1016/j.compscitech.2017.01.025.
  10. Aiping Yu; Palanisamy Ramesh; Mikhail E. Itkis; Elena Bekyarova; Robert C. Haddon; Graphite Nanoplatele-Epoxy Composite Thermal Interface Materials. The Journal of Physical Chemistry C 2007, 111, 7565-7569, 10.1021/jp071761s.
  11. Jinhong Du; Hui-Ming Cheng; The Fabrication, Properties, and Uses of Graphene/Polymer Composites. Macromolecular Chemistry and Physics 2012, 213, 1060-1077, 10.1002/macp.201200029.
  12. L. Monica Veca; Mohammed J. Meziani; Wei Wang; Xin Wang; Fushen Lu; Puyu Zhang; Yi Lin; Robert Fee; John W. Connell; Ya-Ping Sun; et al. Carbon Nanosheets for Polymeric Nanocomposites with High Thermal Conductivity. Advanced Materials 2009, 21, 2088-2092, 10.1002/adma.200802317.
  13. Sabyasachi Ganguli; Ajit K. Roy; David P. Anderson; Improved thermal conductivity for chemically functionalized exfoliated graphite/epoxy composites. Carbon 2008, 46, 806-817, 10.1016/j.carbon.2008.02.008.
  14. Pradip Kumar; Seunggun Yu; Faisal Shahzad; Soon Man Hong; Yoon-Hyun Kim; Chong Min Koo; Ultrahigh electrically and thermally conductive self-aligned graphene/polymer composites using large-area reduced graphene oxides. Carbon 2016, 101, 120-128, 10.1016/j.carbon.2016.01.088.
  15. Nariman Yousefi; Mohsen Moazzami Gudarzi; Qingbin Zheng; Seyed Hamed Aboutalebi; Farhad Sharif; Jang-Kyo Kim; Self-alignment and high electrical conductivity of ultralarge graphene oxide-polyurethane nanocomposites. Journal of Materials Chemistry 2012, 22, 12709-12717, 10.1039/c2jm30590a.
  16. Evgeniy Tkalya; Marcos Ghislandi; Ronald Otten; Mustafa Lotya; Alexander Alekseev; Paul van der Schoot; Jonathan Coleman; Gijsbertus de With; Cor Koning; Experimental and Theoretical Study of the Influence of the State of Dispersion of Graphene on the Percolation Threshold of Conductive Graphene/Polystyrene Nanocomposites. ACS Applied Materials & Interfaces 2014, 6, 15113-15121, 10.1021/am503238z.
  17. Yang Wang; Jerry W. Shan; George J. Weng; Percolation threshold and electrical conductivity of graphene-based nanocomposites with filler agglomeration and interfacial tunneling. Journal of Applied Physics 2015, 118, 065101, 10.1063/1.4928293.
  18. Ana M. Díez-Pascual; Marián A. Gómez-Fatou; Fernando Ania; Araceli Flores; Nanoindentation Assessment of the Interphase in Carbon Nanotube-Based Hierarchical Composites. The Journal of Physical Chemistry C 2012, 116, 24193-24200, 10.1021/jp309067e.
  19. Yizhuo Gu; Min Li; Ji Wang; Zuoguang Zhang; Characterization of the interphase in carbon fiber/polymer composites using a nanoscale dynamic mechanical imaging technique. Carbon 2010, 48, 3229-3235, 10.1016/j.carbon.2010.05.008.
  20. T. D. Downing; R. Kumar; W. M. Cross; L. Kjerengtroen; J. J. Kellar; Determining the interphase thickness and properties in polymer matrix composites using phase imaging atomic force microscopy and nanoindentation. Journal of Adhesion Science and Technology 2000, 14, 1801-1812, 10.1163/156856100743248.
  21. Muhammad Aqeel Ashraf; Wanxi Peng; Yasser Zare; Kyong Yop Rhee; Effects of Size and Aggregation/Agglomeration of Nanoparticles on the Interfacial/Interphase Properties and Tensile Strength of Polymer Nanocomposites. Nanoscale Research Letters 2018, 13, 1-7, 10.1186/s11671-018-2624-0.
  22. Jun Huang; Yu Wu; Lixin Huang; Evaluation of the mechanical properties of graphene-based nanocomposites incorporating a graded interphase based on isoparametric graded finite element model. Composite Interfaces 2020, 28, 543-575, 10.1080/09276440.2020.1794164.
  23. Chaoying Wan; Biqiong Chen; Reinforcement and interphase of polymer/graphene oxide nanocomposites. Journal of Materials Chemistry 2012, 22, 3637-3646, 10.1039/c2jm15062j.
  24. Liliane Bokobza; Bruno Bresson; Gilles Garnaud; Jingxue Zhang; Mechanical and AFM investigations of elastomers filled with multiwall carbon nanotubes. Composite Interfaces 2012, 19, 285-295, 10.1080/15685543.2012.712486.
  25. Weihua Kai; Yuuki Hirota; Lei Hua; Yoshio Inoue; Thermal and mechanical properties of a poly(e-caprolactone)/graphite oxide composite. Journal of Applied Polymer Science 2007, 107, 1395-1400, 10.1002/app.27210.
  26. Masukazu Hirata; Takuya Gotou; Shigeo Horiuchi; Masahiro Fujiwara; Michio Ohba; Thin-film particles of graphite oxide 1: High-yield synthesis and flexibility of the particles. Carbon 2004, 42, 2929-2937, 10.1016/j.carbon.2004.07.003.
  27. Hyunwoo Kim; Yutaka Miura; Christopher W. Macosko; Graphene/Polyurethane Nanocomposites for Improved Gas Barrier and Electrical Conductivity. Chemistry of Materials 2010, 22, 3441-3450, 10.1021/cm100477v.
  28. Bo Peng; Fengnian Zhao; Jianfeng Ping; Yibin Ying; Recent Advances in Nanomaterial-Enabled Wearable Sensors: Material Synthesis, Sensor Design, and Personal Health Monitoring. Small 2020, 16, 2002681, 10.1002/smll.202002681.
  29. Sichao Hou; Aiying Zhang; Ming Su; Nanomaterials for Biosensing Applications. Nanomaterials 2016, 6, 58, 10.3390/nano6040058.
  30. Han Jin; Yasmin Shibli Abu-Raya; Hossam Haick; Advanced Materials for Health Monitoring with Skin-Based Wearable Devices. Advanced Healthcare Materials 2017, 6, 1700024, 10.1002/adhm.201700024.
  31. Wanasinghe Arachchige Dumith Madushanka Jayathilaka; Kun Qi; Yanli Qin; Amutha Chinnappan; William Serrano-García; Chinnappan Baskar; Hongbo Wang; Jianxin He; Shizhong Cui; Sylvia W. Thomas; et al. Significance of Nanomaterials in Wearables: A Review on Wearable Actuators and Sensors. Advanced Materials 2018, 31, e1805921, 10.1002/adma.201805921.
  32. Zheng Lou; Lili Wang; Kai Jiang; Zhongming Wei; Guozhen Shen; Reviews of wearable healthcare systems: Materials, devices and system integration. Materials Science and Engineering: R: Reports 2019, 140, 100523, 10.1016/j.mser.2019.100523.
  33. Jen-Lin Chang; Kuo-Hsin Chang; Chi-Chang Hu; Wan-Ling Cheng; Jyh-Myng Zen; Improved voltammetric peak separation and sensitivity of uric acid and ascorbic acid at nanoplatelets of graphitic oxide. Electrochemistry Communications 2010, 12, 596-599, 10.1016/j.elecom.2010.02.008.
  34. Haijun Du; Jianshan Ye; Jiaqi Zhang; Xiaodan Huang; Chengzhong Yu; A voltammetric sensor based on graphene-modified electrode for simultaneous determination of catechol and hydroquinone. Journal of Electroanalytical Chemistry 2011, 650, 209-213, 10.1016/j.jelechem.2010.10.002.
  35. Cheng Xiang Lim; Hui Ying Hoh; Priscilla Kailian Ang; Kian Ping Loh; Direct Voltammetric Detection of DNA and pH Sensing on Epitaxial Graphene: An Insight into the Role of Oxygenated Defects. Analytical Chemistry 2010, 82, 7387-7393, 10.1021/ac101519v.
  36. Ming Zhou; Yueming Zhai; Shaojun Dong; Electrochemical Sensing and Biosensing Platform Based on Chemically Reduced Graphene Oxide. Analytical Chemistry 2009, 81, 5603-5613, 10.1021/ac900136z.
  37. M. Gabriella Santonicola; Label-Free Biosensing Platforms Based on Graphene/DNA Interfaces. Graphene Bioelectronics 2018, , , 10.1016/b978-0-12-813349-1.00008-1.
  38. Sabina Botti; Alessandro Rufoloni; Susanna Laurenzi; Stefano Gay; Tomas Rindzevičius; Michael Stenbæk Schmidt; M. Gabriella Santonicola; DNA self-assembly on graphene surface studied by SERS mapping. Carbon 2016, 109, 363-372, 10.1016/j.carbon.2016.07.069.
  39. Avinash J. Patil; Jemma L. Vickery; Thomas B. Scott; Stephen Mann; Aqueous Stabilization and Self-Assembly of Graphene Sheets into Layered Bio-Nanocomposites using DNA. Advanced Materials 2009, 21, 3159-3164, 10.1002/adma.200803633.
  40. Alessandra Bonanni; Adriano Ambrosi; Martin Pumera; Nucleic Acid Functionalized Graphene for Biosensing. Chemistry – A European Journal 2011, 18, 1668-1673, 10.1002/chem.201102850.
  41. Qiaojuan Gong; Haixia Han; Haiying Yang; Meiling Zhang; Xiaoling Sun; Yunxia Liang; Zhaorong Liu; Wenchan Zhang; Jinli Qiao; Sensitive electrochemical DNA sensor for the detection of HIV based on a polyaniline/graphene nanocomposite. Journal of Materiomics 2019, 5, 313-319, 10.1016/j.jmat.2019.03.004.
  42. Magda Silva; Natália M. Alves; Maria C. Paiva; Graphene-polymer nanocomposites for biomedical applications. Polymers for Advanced Technologies 2017, 29, 687-700, 10.1002/pat.4164.
  43. Xinhuang Kang; Jun Wang; Hong Wu; Ilhan A. Aksay; Jun Liu; Yuehe Lin; Glucose Oxidase-graphene-chitosan modified electrode for direct electrochemistry and glucose sensing. Biosensors and Bioelectronics 2009, 25, 901-905, 10.1016/j.bios.2009.09.004.
  44. Huiting Lian; Zhaohui Sun; Xiangying Sun; Bin Liu; Graphene Doped Molecularly Imprinted Electrochemical Sensor for Uric Acid. Analytical Letters 2012, 45, 2717-2727, 10.1080/00032719.2012.702173.
  45. Marco Orecchioni; Cecilia Menard-Moyon; Lucia Gemma Delogu; Alberto Bianco; Graphene and the immune system: Challenges and potentiality. Advanced Drug Delivery Reviews 2016, 105, 163-175, 10.1016/j.addr.2016.05.014.
  46. Yunying Xie; Aiqiong Chen; Dan Du; Yuehe Lin; Graphene-based immunosensor for electrochemical quantification of phosphorylated p53 (S15). Analytica Chimica Acta 2011, 699, 44-48, 10.1016/j.aca.2011.05.010.
  47. Hong Wu; Jun Wang; Xinhuang Kang; Chongmin Wang; Donghai Wang; Jun Liu; Ilhan A. Aksay; Yuehe Lin; Glucose biosensor based on immobilization of glucose oxidase in platinum nanoparticles/graphene/chitosan nanocomposite film. Talanta 2009, 80, 403-406, 10.1016/j.talanta.2009.06.054.
  48. Shifeng Hou; Marc L. Kasner; Shujun Su; Krutika Patel; Robert Cuellari; Highly Sensitive and Selective Dopamine Biosensor Fabricated with Silanized Graphene. The Journal of Physical Chemistry C 2010, 114, 14915-14921, 10.1021/jp1020593.
  49. S.L. Burrs; M. Bhargava; R. Sidhu; J. Kiernan-Lewis; C. Gomes; J.C. Claussen; E.S. McLamore; A paper based graphene-nanocauliflower hybrid composite for point of care biosensing. Biosensors and Bioelectronics 2016, 85, 479-487, 10.1016/j.bios.2016.05.037.
More
This entry is offline, you can click here to edit this entry!
Video Production Service