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 + 1360 word(s) 1360 2021-03-24 07:43:09 |
2 format correct Meta information modification 1360 2021-04-07 12:29:58 |

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.
Tang, X. Graphene Sensors. Encyclopedia. Available online: https://encyclopedia.pub/entry/8500 (accessed on 29 July 2024).
Tang X. Graphene Sensors. Encyclopedia. Available at: https://encyclopedia.pub/entry/8500. Accessed July 29, 2024.
Tang, Xiaohui. "Graphene Sensors" Encyclopedia, https://encyclopedia.pub/entry/8500 (accessed July 29, 2024).
Tang, X. (2021, April 07). Graphene Sensors. In Encyclopedia. https://encyclopedia.pub/entry/8500
Tang, Xiaohui. "Graphene Sensors." Encyclopedia. Web. 07 April, 2021.
Graphene Sensors
Edit

Graphene is one of the most promising materials for gas-sensor applications.

graphene sensor functionalized graphene ammonia detection

1. Energy Band Structure of Graphene

Graphene is a two-dimensional (2D) material with a monolayer of sp2-bonded carbon atoms. Electrons are delocalized and free to move in this 2D sheet. The valence and conduction bands of graphene are conical valleys, which are attached at Dirac points. The energy dispersion curve is linear around the Dirac points, leading to an extremely large mobility of the charge carriers. Pristine graphene is a zero bandgap semimetal, and its Fermi level is located at the Dirac point, corresponding to zero conductivity at T = 0 K under vacuum [1]. In practical situations, graphene possesses defects that cause a p-type doping [2]. The zero or quasi-zero bandgap of pristine graphene can be a hurdle to use it as a sensing layer [3]. However, the electron transfer with target gas molecules leads to the Fermi level shift. More precisely, when graphene reacts with electron-donor or electron-acceptor gas molecules, the zero conductivity disappears. Moreover, doping or defects can also shift the Fermi level, which makes graphene become a p-type or n-type semiconductor. Due to its energy band structure, graphene is very sensitive to Fermi level changes. Its conductivity can be easily modified by doping or a back gate voltage when the graphene sheet is fabricated as a field effect transistor sensor [4].

2. Physical Properties of Graphene

Single-layer graphene has the largest surface-to-volume ratio of 2630 m2/g [5] and exposes all carbon atoms to environment. This can provide the largest number of binding sites per unit volume to yield high sensitivity for sensors. The high carrier mobility (200,000 cm2/s.V) [6] and excellent electrical conductivity (1738 Siemens/m) [7] of graphene at room temperature inherently ensure low electrical noise and energy consumption in graphene sensors. Carrier concentration in graphene strongly affects the graphene resistance, and it can be modulated by the adsorption of target molecules injecting electrons or holes into graphene. The significant change, coupled with a low electrical noise, of the graphene resistance makes it have the ability for detecting a single molecule. Particularly, graphene is easy to be functionalized and compatible with organic molecules through a π–π stacking interaction and/or electrostatic interaction [8]. For graphene oxide (GO), the oxygen-containing functional groups (such as hydroxyl and epoxy) are attached to honeycomb-like carbon of graphene. Therefore, GO can be readily modified with target molecules through the functional groups. These facts make graphene and graphene oxide ideal materials for designing gas sensors.

3. Preparation Methods of Graphene

Historical graphene was first prepared by mechanical exfoliation of graphite [4] [9]. For example, highly oriented pyrolytic graphite (HOPG) is attached to the photoresist layer over a glass substrate. Using scotch tape, thick flakes of graphite are repeatedly peeled off. Thin flakes left in the photoresist are released in acetone. The target substrate (such as SiO2/Si wafer) is dipped in the acetone and then washed with plenty of water and propanol, which removes mostly thick flakes. In this case, some thin flakes are captured on the target substrate’s surface due to van der Waals and/or capillary forces. Although mechanical exfoliation can provide high-quality single-layer graphene, it is not a scalable technique, and other techniques were proposed.

Chemical vapor deposition (CVD) holds great promises for large-scale production of high-quality graphene [10],[11]. Conceptually, graphene CVD is a simple technique: it involves the decomposition of hydrocarbon precursors (such as CH4) on catalytic metal substrates at high temperature (about 1000 °C) in a controlled atmosphere or atmospheric pressure. Cu, Ni, Ti, Ru, Pd, Pt, Ir, Co, Au, and Rh can be used as the catalytic metal substrates [12]. Among them, Cu is a very attractive catalyst and extensively chosen, since very low carbon solubility of Cu allows self-limited graphene growth, leading to highly homogeneous graphene sheets.

Graphene has also been grown on the surface of single-crystal silicon carbide (SiC) by thermal decomposition [13]. Namely, Si sublimation from SiC surface yields single-layer or multi-layer graphene structure at the graphene–SiC interface. However, it is quite difficult to exfoliate or transfer from the SiC substrate.

For large-scale production of low-cost graphene, one of the most popular approaches is the use of strong oxidizing agents to obtain GO and then convert GO to reduced graphene oxide (rGO) by thermal treatment or chemical reduction. In 1958, Hummers [14] reported a method for the GO synthesis by using KMnO4 and NaNO3 in concentrated H2SO4. This method has broadly been used in gas-sensor fabrication because of its high efficiency. However, it still has a few drawbacks. For instance, the oxidation procedure releases toxic gasses and the control of the graphene oxide sheets is difficult. Recently, several modified methods were proposed in the literature to improve the control of the GO and rGO synthesis [15],[16]. It is worth noting that the reduction in GO is not complete, and there are defects and OH groups in the final rGO [17],[18].

4. Functionalized Graphene NH3 Sensors

Pristine graphene without bandgap behaves like a semimetal [19]. It is not a good choice for the NH3 detection, because it does not possess any functional groups or defects. Monocrystalline graphene of mechanical exfoliation is well known as pristine graphene. However, it includes few isolated point defects. When NH3 molecules are attached on the isolated point defects, the resistance of the pristine graphene sensor does not significantly change. Since the donor electrons of NH3 can transfer through low resistance pathways around the isolated point defects [20]. On the contrary, functionalized graphene is one of the most promising materials for the NH3 sensor. Thanks to the abundant defects (grain boundaries, line defects, or a few point defects) and functional groups on graphene, they strongly adsorb NH3 molecules. Around such defects, no low-resistance pathways exist. So, the resistance change caused by NH3 adsorption is significant, leading to a higher sensitivity.

Non-functionalized graphene is nonselective to gases [21]. A variety of gas molecules can be adsorbed on it to give similar resistance changes. For p-type graphene, the adsorption of a reducing gas (either NH3 or H2S) donates electrons to the graphene and depletes the concentration of holes, increasing the graphene resistance. The adsorption of an oxidizing gas (either NO2 or SO2) accepts electrons from the graphene and enhances the concentration of holes, decreasing the graphene resistance. As a consequence, different gases may generate the same sensing signals [22]. Indeed, the first non-functionalized graphene sensor can adsorb not only NH3, but also CO, H2O, and NO2 in the same conditions. Ultraviolet or infrared radiation is able to make graphene sensitive to different types of gases, allowing for certain selectivity. However, the light adsorption efficiency of graphene is weak [23], [24]. The selective capabilities of graphene can be easily enhanced by grafting functional groups on its surface. This strategy is followed by most of the researchers.

The functionalization (generating specific links and increasing the adsorbing sites) is a viable method to improve chemical inert and poor selectivity of graphene. The graphene functionalization can be classified to covalent and non-covalent strategies [25]. The former transforms graphene’s π-orbitals from sp2 into sp3, disrupting graphene electronic and mechanical properties. Whereas, the latter provides functional groups to the graphene surface by electrostatic or π–π interactions, enhancing the graphene bonding ability and simultaneously preserving the graphene original properties (the high carrier mobility and favorable noise characteristics). It is, thus, important to optimize and improve non-covalent functionalization techniques.

Over the past few years, a number of functionalization procedures have been developed. Various materials, from metallic nanoparticles and metal oxides to organic molecules and conducting polymers, are used to accomplish the graphene functionalization for promoting a high sensitivity and great selectivity to NH3 sensing. According to the interaction nature between the target molecule and the receptor attached to graphene (or rGO), sensing reactions are distinguished to physisorption and chemisorption. The physisorption depends on the van der Waals forces of attraction with low binding energy, leading to full and fast recovery, but low sensitivity and poor selectivity. The chemisorption depends on the formation of chemical bonds between NH3 molecules and the sensing layers, resulting in good selectivity, but slow and incomplete recovery. For certain sensing layers, physisorption and chemisorption can happen at the same time. Both adsorptions can cause charge transfer, yielding p-doping or n-doping of the graphene film.

References

  1. Meunier, V.; Souza Filho, A.G.; Barros, E.B.; Dresselhaus, M.S.; Physical properties of low-dimensional s p 2 -based carbon nanostructures. Rev. Mod. Phys. 2016, 88, 025005, https://doi.org/10.1103/RevModPhys.88.025005.
  2. Suk, J.W.; Lee, W.H.; Lee, J.; Chou, H.; Piner, R.D.; Hao, Y.; Akinwande, D.; Ruoff, R.S.; Enhancement of the Electrical Properties of Graphene Grown by Chemical Vapor Deposition via Controlling the Effects of Polymer Residue. Nano Lett. 2013, 13, 1462–1467, 10.1021/nl304420b.
  3. Donarelli, M.; Ottaviano, L.; 2D Materials for Gas Sensing Applications: A Review on Graphene Oxide, MoS2, WS2 and Phosphorene. Sensors 2018, 18, 3638, https://doi.org/10.3390/s18113638.
  4. Novoselov, K.S.; Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666-669, 284146851.
  5. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S.; Graphene and Graphene Oxide: Synthesis, Properties, and Applications. Adv. Mater. 2010, 22, 3906–3924, 10.1002/adma.201001068.
  6. Bolotin, K.I.; Sikes, K.J.; Jiang, Z.; Klima, M.; Fudenberg, G.; Hone, J.; Kim, P.; Stormer, H.L.; Ultrahigh electron mobility in suspended graphene. Solid State Commun. 2008, 146 , 351–355, 10.1016/j.ssc.2008.02.024.
  7. Weiss, N.O.; Zhou, H.; Liao, L.; Liu, Y.; Jiang, S.; Huang, Y.; Duan, X.; Graphene: An Emerging Electronic Material. Adv. Mater. 2012, 24, 5782–5825, 10.1002/adma.201201482.
  8. Yang, Y.; Asiri, A.M.; Tang, Z.; Du, D.; Lin, Y.; Graphene based materials for biomedical applications. Mater. Today 2013, 16, 365–373, http://dx.doi.org/10.1016/j.mattod.2013.09.004.
  9. Li, X.; Cai, W.; An, J.; Kim, S.; Nah, J.; Yang, D.; Piner, R.; Velamakanni, A.; Jung, I.; Tutuc, E.; et al.et al. Large-Area Synthesis of High-Quality and Uniform Graphene Films on Copper Foils. Science 2009, 324, 1312–1314, 10.1126/science.1171245 .
  10. Reckinger, N.; Tang, X.; Joucken, F.; Lajaunie, L.; Arenal, R.; Dubois, E.; Hackens, B.; Henrard, L.; Colomer, J.-F.; Oxidation-assisted graphene heteroepitaxy on copper foil. Nanoscale 2016, 8, 18751–18759, 10.1039/c6nr02936a.
  11. Reckinger, N.; Casa, M.; Scheerder, J.E.; Keijers, W.; Paillet, M.; Huntzinger, J.-R.; Haye, E.; Felten, A.; Van de Vondel, J.; Sarno, M.; et al.et al. Restoring self-limited growth of single-layer graphene on copper foil via backside coating. Nanoscale 2019, 11, 5094–5101, 10.1039/c8nr09841g.
  12. Yusop, M.Z.M.; Elias, M.S.; Dzarfan, M.H.; Aziz, M.; Ismail, A.F.; Effects of copper, nickel, and its alloy as catalysts for graphene growth via chemical vapor deposition method: A review. Malays. J. Fundam. Appl. Sci. 2019, 15, 508–515, ISSN:2289-599X.
  13. Berger, C.; Electronic Confinement and Coherence in Patterned Epitaxial Graphene. Science 2006, 312, 1191–1196, 10.1126/science.1125925 .
  14. Hummers,W.S.; Offeman, R.E.; Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339, 10.1021/ja01539a017.
  15. Lotya, M.; Hernandez, Y.; King, P.J.; Smith, R.J.; Nicolosi, V.; Karlsson, L.S.; Blighe, F.M.; De, S.; Wang, Z.; McGovern, I.T.; et al.et al. Liquid Phase Production of Graphene by Exfoliation of Graphite in Surfactant/Water Solutions. J. Am. Chem. Soc. 2009, 131, 3611–3620, 10.1021/ja807449u.
  16. Alam, S.N.; Sharma, N.; Kumar, L.; Synthesis of Graphene Oxide (GO) by Modified Hummers Method and Its Thermal Reduction to Obtain Reduced Graphene Oxide (rGO)*. Graphene 2017, 06, 1–18, 10.4236/graphene.2017.61001.
  17. Hidayah, N.M.S.; Liu, W.-W.; Lai, C.-W.; Noriman, N.Z.; Khe, C.-S.; Hashim, U.; Lee, H.C.; Comparison on Graphite, Graphene Oxide and Reduced Graphene Oxide: Synthesis and Characterization. Penang, Malaysia 2017, AIP Conference Proceedings 1892, 150002, 10.1063/1.5005764.
  18. Gadgil, B.; Damlin, P.; Kvarnström, C.; Graphene vs. reduced graphene oxide: A comparative study of graphene-based nanoplatforms on electrochromic switching kinetics. Carbon 2016, 96, 377–381, 10.1016/j.carbon.2015.09.065.
  19. Tang, X.; Reckinger, N.; Poncelet, O.; Louette, P.; Ureña, F.; Idrissi, H.; Turner, S.; Cabosart, D.; Colomer, J.-F.; Raskin, J.-P.; et al.et al. Damage Evaluation in Graphene Underlying Atomic Layer Deposition Dielectrics. Sci. Rep. 2015, 5, 13523, : 10.1038/srep13523.
  20. Salehi-Khojin, A.; Estrada, D.; Lin, K.Y.; Bae, M.-H.; Xiong, F.; Pop, E.; Masel, R.I.; Polycrystalline Graphene Ribbons as Chemiresistors. Adv. Mater. 2012, 24, 53–57, 10.1002/adma.201102663.
  21. Zhang, Y.-H.; Chen, Y.-B.; Zhou, K.-G.; Liu, C.-H.; Zeng, J.; Zhang, H.-L.; Peng, Y.; Improving gas sensing properties of graphene by introducing dopants and defects: A first-principles study. Nanotechnology 2009, 20, 185504, 10.1088/0957-4484/20/18/185504.
  22. Yuan, W.; Shi, G.; Graphene-based gas sensors. J. Mater. Chem. A 2013, 1, 10078, 10.1039/C3TA11774J.
  23. Nair, R.R.; Blake, P.; Grigorenko, A.N.; Novoselov, K.S.; Booth, T.J.; Stauber, T.; Peres, N.M.R.; Geim, A.K.; Fine Structure Constant Defines Visual Transparency of Graphene. Science 2008, 320, 1308, 10.1126/science.1156965 .
  24. Liu, B.; Tang, C.; Chen, J.; Xie, N.; Tang, H.; Zhu, X.; Park,; Multiband and Broadband Absorption Enhancement of Monolayer Graphene at Optical Frequencies from Multiple Magnetic Dipole Resonances in Metamaterials. Nanoscale Res. Lett. 2018, 13, 153, 10.1186/s11671-018-2569-3.
  25. 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, 10.1021/cr3000412.
More
Information
Subjects: Polymer Science
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 548
Revisions: 2 times (View History)
Update Date: 07 Apr 2021
1000/1000
Video Production Service