Skip to main content
ACS Omega logoLink to ACS Omega
. 2023 Jan 23;8(5):4436–4452. doi: 10.1021/acsomega.2c05953

Carbon Nanostructure Embedded Novel Sensor Implementation for Detection of Aromatic Volatile Organic Compounds: An Organized Review

Nibedita Nath , Anupam Kumar , Subhendu Chakroborty §,*, Siba Soren ∥,*, Arundhati Barik , Kaushik Pal #,*, Fernando Gomes de Souza Jr
PMCID: PMC9909795  PMID: 36777592

Abstract

graphic file with name ao2c05953_0019.jpg

For field-like environmental gas monitoring and noninvasive illness diagnostics, effective sensing materials with exceptional sensing capabilities of sensitive, quick detection of volatile organic compounds (VOCs) are required. Carbon-based nanomaterials (CNMs), like CNTs, graphene, carbon dots (Cdots), and others, have recently drawn a lot of interest for their future application as an elevated-performance sensor for the detection of VOCs. CNMs have a greater potential for developing selective sensors that target VOCs due to their tunable chemical and surface properties. Additionally, the mechanical versatility of CNMs enables the development of novel gas sensors and places them ahead of other sensing materials for wearable applications. An overview of the latest advancements in the study of CNM-based sensors is given in this comprehensive organized review.

1. Introduction

Volatile organic compounds (VOCs) are substances that are evaporated and enter the environment under ordinary circumstances. They are present in a variety of products. Due to higher volatility, mobility, and resistance to degradation, VOCs can travel over large distances in the atmosphere.1 The most common VOCs are halogenated hydrocarbons, e.g., vinyl chloride and chloroethene, as well as aromatic hydrocarbons like: benzene, methylbenzene, xylene, and ethylbenzene.

VOCs come from both natural and man-made sources. Plant emissions, naturally occurring forest fires, and anaerobic moor processes are examples of natural sources. Household and commercial processes that produce VOCs include food processing, the use of fertilizers and pesticides, septic systems, chlorination, transportation, burning hydrocarbon fuels, storing and distributing petroleum, cleaning textiles, printing, the pharmaceutical industry, and more2 (Figure 1).

Figure 1.

Figure 1

Possible origins of VOC emission. Reprinted with permission from ref (2). Copyright 2021 MDPI.

Even though VOCs are utilized in many aspects of daily life, their exposure is dangerous. Ocular and sore throats, vomiting, nausea, headaches, loss of coordination, skin allergies, and other symptoms are brought on by small-scale exposure to VOCs.35 VOCs have been linked to long-term effects in humans, including harm to the central nervous system, reproductive system, lungs, liver, and kidneys.6 It is generally recognized that some VOCs might cause cancer.7,8 All living things frequently die when exposed to VOCs over an extended period. The primary causes of global warming are several VOCs, including methane. Additionally, certain VOCs are involved in the destruction of the stratospheric ozone.9

Significant approaches have been developed to discover VOC analytes in difficult matrices. Gas chromatography–mass spectrometry (GC-MS) is one of the most prominent analytical techniques for identifying VOCs.10,11 This is because of its consistent accuracy and specificity. Real-time monitoring and analysis of VOC gas have proven difficult due to its high cost, large equipment, and time-consuming sample preconcentration.1214 According to their obvious benefits of being portable, affordable, and incredibly sensitive, gas sensors have opened new opportunities for VOC detection in relation to the aforementioned difficulties. This enables the reliability of online analysis and real-time monitoring of VOC analytes. Organic semiconductors, metal oxides, noble metals, sulfides, and carbon-based materials can be used to make functional VOC gas sensors. Carbon materials, particularly those with intrinsic nanoscale characteristics, have been identified as one of the most promising options for detecting VOC gas.1518

Nowadays, carbon-based nanomaterials are encouraging the scientific community to make powerful sensor devices due to their characteristic structure, especially graphene and its derivatives such as graphene oxide (GO), CNTs, reduced graphene oxide (rGO), Cdots, and MXene. Several ways of interacting with VOCs, such as π–π stacking, electrostatic forces, and noncovalent bonding, make them suitable candidates for adsorbents.19 Carbon-based nanomaterials (CNMs) are appealing prospects for the future of automated sensor technology because of their excellent sensing detectability and intriguing transduction capabilities. A majority of atoms in 0–2-dimensional CNMs are exposed to the outside environment due to their huge specific surface area and remarkable sensitivity.2022 Low-dimensional CNMs like CNTs, graphene, and MXene have a large specific surface area to reach a high detection sensitivity for VOCs; however, the majority of their atoms are discharged into the air.23,24 We focused on the most recent developments in CNMs for VOC sensing in this comprehensive review.

2. General Overview of the Shape and Size of rGO, CNTs, and CDs

The carbon nanomaterials can be synthesized with controlled shape and size from 0D to 2D, whereas 3D carbon materials are considered as graphite, diamond, etc. Except for graphite and diamond, we have focused on graphene, CNTs, CDs, and other carbon-based nanomaterials. When we talk about nanomaterials, their shapes and sizes are more critical because of surface catalytic reactions. The reduced graphene (rGO) obtained by heating or using a reducing agent on GO sheets is not spherical. Its nominal effective diameter can be calculated instead of the actual size using dynamic light-scattering (DLS) techniques. Through this technique, different scientists measured the average sizes of rGO sheet combinations of honeycomb structures of a carbon atom, which were found to be 2.93 μm, 0.5–5 μm, etc.25,26 The CNT, which is a circular tube-like structure due to its rolled up graphene sheets, can be a single-walled or multi-walled CNT. Hence, it is considered a 1D carbon material; its length can be extended to some micrometers long, whereas its diameter can be found to be in the nm range.27,28 Carbon dots (CDs) are gaining much more attention because of their biocompatibility, excellent physicochemistry, and tunable photoluminescence and are being explored as biosensors.29,30 They are a new type of carbon nanostructure of >10 nm size and the most water-soluble carbon material.31 The effective visualization of their shape and size can be studied by a TEM image, which has been shown in Figure 2.

Figure 2.

Figure 2

Graphical representations of TEM images of (a,b) graphene, (c,d) CNTs, and (e,f) Cdots. Reprinted with permission from refs (32), (33), and (34). Copyright 2014 Springer Nature.

The other derivatives or oxidized forms of graphene by oxygen functional groups such as carboxyl (−C=O), epoxy (−C–O–C−), and hydroxyl (−OH) make graphene oxide suspended in water and other polar media.35 This degree of oxidation is responsible for the electrical conductivity of modified graphene and the degree of oxidation depending on various synthetic routes. The carbon atoms of modified oxidized graphene are partially sp3 hybridized and can move above and below the plane.36 Another derivative of graphene-like sheets can be obtained by reducing graphene oxide through chemical reduction or simple heating methods. Park and Ruoff examined the elemental analysis of carbon and oxygen atoms (atomic C/O ratio, ∼10) for rGO made by combustion techniques and discovered that rGO is not the same as pure graphene since it contains a considerable quantity of oxygen. Furthermore, when cooled to lower temperatures, its conductivity reduces by 3 orders of magnitude, causing it to display nonmetallic behavior.37 In this way, a very clear difference can be seen between pristine graphene and graphene derivatives such as GO and rGO. Both pristine graphene and its derivatives are two-dimensional structures. Still, pristine graphene is relatively inefficient for the adsorption of VOC molecules. In contrast, its derivatives have more adsorption sites due to the presence of different oxygen groups and hence offer high sensitivity toward VOC molecules of the film.38 On the other hand, the current–voltage (IV) characteristics for both GO and rGO exhibit ohmic contact, which is suitable for gas-sensing applications.39

3. sp2-Hybridized Electronic Behavior for Sensing

It is very much essential to note the origin of electrical conductivity properties of graphene, CNTs, and Cdots for sensing properties. It will help the understanding of surface interaction as a chemisorption reaction which changes the electrical properties of carbon-based materials. The 2D graphene sheet shows high carrier mobility due to planar sp2-hybridized atomic orbitals, and as a result, reduced graphene oxide (rGO) became a conductive active material for molecular sensors. The rolling form of the graphene sheet (i.e., a carbon nanotube, CNT) results in π confinement due to σ–π rehybridization at its circular curvature structure and creates an asymmetric distribution of σ-bonds and π-orbitals on both sides (inside and outside) of the nanotube, forming a high π-electron conjugation outside the nanotube.40 Thus, the surface of CNTs can donate or withdraw electrons from any electron-donating or electron-withdrawing entity during molecular interaction, thereby modulating overall electrical resistance.41,42 Several groups have approached the molecular orbital (MO) theory for the electronic structure of CQDs, which demonstrated that transitions have occurred in n → π* and π → π* as available transition energies.4345 According to Larson and Hu, the aromatic sp2-hybridized carbons are responsible for their π-states. The presence of π-electron conjugation results in high charge transfer in carbon-based materials.46 Generally, the detection of VOCs is governed by their interaction with carbon-based materials by diffusion and sensed by the active center of carbon-based materials,4750 microgravimetric sensors,5155 and resistive sensors.5659 When a molecule species interacts with the sidewalls of carbon-based or any semiconductor materials during gas detection, changes in conductance60,61 (due to charge transfer or mobility change) or capacitance62 (from inherent or induced dipole moments) occur.6365 The interaction between the VOCs and carbon-based compounds (rGO, CNTs, and CDs) depends strongly on the architecture and design of the composites or materials which may have a larger contact area, thus allowing greater sensitivity. The CNT has two variables, SWCNT and MWCNT. Depending upon chirality, SWCNTs can be semiconducting or metallic,66,67 and MWCNTs have only a metallic character and are therefore unsuitable for fabricating gas transistors.

4. Sensing of VOCs

Recently, point-of-care applications, various disease diagnoses, atmospheric gas tracking, and other fields have significantly increased the demand for accurate VOC gas analysis.6870 Meanwhile, research into the construction of sophisticated sensors for VOC detection is expanding rapidly (Stewart et al. 2020; Das et al. 2020).71,72 A thorough explanation of the VOC-detecting device’s sensing process is initially required to construct a powerful VOC sensor.

Research on developing gas sensors has primarily focused on nanoparticles because of their superior physicochemical characteristics and high surface-to-volume ratio. The recent discovery of CNMs has prompted7375 research toward sensing technologies. These materials have been shown to be more effective in producing higher VOC sensors.7678 Three steps commonly take place in gas sensors manufactured using CNMs for VOC detection: (i) trapping of VOC gas, (ii) their interaction with the sensing active site, and (iii) dispersing of the VOC gas. Focusing on the complex interactions between the VOC gas and the active center that detects them, CNMs have been used to create a range of VOC gas sensors, including optical76,7981 and microgravimetric ones.7786 Regarding optical sensors, spectrometric or colorimetric variations in the sensing materials are brought on by the intermolecular interactions among VOC samples and the active center76,7981 and resistance-based sensors.87,78,8890

5. CNM-Based Sensor for VOC Detection

Due to the considerable electrochemical, mechanical, optical, and thermal properties of CNMs, many research studies have been done in this area during the last three decades. Fullerenes, carbon nanotubes, graphene, carbon dots, carbon nanohorns, and carbon black are examples of zero-, one-, two-, and three-dimensional CNMs that have shown such inherent properties that can be easily modified in the development of cutting-edge innovation for sensing applications. Nanomaterials have opened up new pathways and options for sensing analysts or target molecules.91

Carbon-based nanostructures have several benefits over other commonly used sensor materials, including simple manufacturing procedures, unexpected physiochemical characteristics, greater sensing properties, environmentally friendly materials, improved detection accuracy, high flexibility, sensitivity, and reliability. They also have a large surface-to-volume ratio, high stability, high electrical conductivity, biocompatibility, reduced size, and an improved surface area. CNMs are being studied as a result of further utilization of efficient sensor technology.92,93

Carbon nanostructures like CNTs and graphene can be used to detect extremely small levels of danger and greenhouse gases. Therefore, using the CNM-based sensor to develop susceptible, small energy gas sensors is both of significant academic interest and of enormous economic significance.94 The many CNM-based sensors utilized for VOC detection are listed in Table 1 in detail.

Table 1. Various CNM-Based Sensors for VOC Detection.

CNMs VOCs temperature (°C) response/recovery time (s) ref
CNT/SnO2 CH3OH/C2H5OH 250–300 20/30 (95)
ZnO/MWCNT nanorod C2H5OH 300 5/8 (96)
SnO2/MWCNT C3H6O RT 7/8 (97)
PMMA/POSS/CNT HCHO RT <5 s (98)
Graphene/ZnO C2H5OH 340 5/20 (99)
Graphene/Ni-SnO2 C3H6O 350 5.4 (100)
Graphene/ZnO acetone RT (101)
Graphene/ZnO HCHO RT 36 (102)
rGO/SnO2 aerogel C6H6O RT 2.43/1.06 (103)
rGO/α-Fe2O3 nanofiber C3H6O 375 3/9 (104)
rGO/ZnSnO3 HCHO 103 31/– (105)
rGO/DF-PDI TEA RT 64/128 (106)
ZnO quantum dot/graphene HCHO RT 30/40 (107)
Au@NGQDs/TiO2 HCHO 150 18/20 (108)
3D Mxene framework acetone RT 1.5 min/1.7 min (109)
Ti3C2Tx C2H5OH RT (110)

6. CNT-Based Sensor for VOC Detection

A one-dimensional carbon allotrope known as a carbon nanotube (CNT) is composed of sp2 carbon atoms organized in cylindrical tubes with diameters ranging from 1 to 100 nm. Single-wall carbon nanotubes (SWCNTs) and multiwalled carbon nanotubes (MWCNTs) are the two types of CNTs used most commonly. MWCNTs, on the other hand, are composed of multiple sheets of concentric single-walled graphene cylinders with an interlayer spacing of 3.4 and are held together by van der Waals forces.111 SWCNTs are the rolled form of the graphene sheet. Hexagonal rings make up CNTs and control their size, curvature, and electronic properties. Chirality refers to the configuration of carbon hexagonal rings in nanotubes.112

Since 20 years, CNTs have been made on a massive scale for numerous uses. Arc discharge,113 CVD,114 and laser ablation115 are common techniques for making CNTs. The CVD process and the discharge method are more advantageous for the mass production of high-quality CNTs.114

One of the oldest clinical practices is breathing analysis used for medical purposes. The binding of volatile compounds to a sensor array, the creation of sensor modifications, which produce distinctive patterns of signals, and the integration of signal patterns allowing categorization are the main phases in the breathing analysis working principle, which also is based on the human olfactory system.116 A quick, painless, inexpensive, noninvasive method for early disease detection and ongoing physiological monitoring is breathing analysis.116118 Analyzing the diaphragmatic movement when breathing can help detect human disorders like atypical sleep problems, asthma, heart attack, and lung cancer early on.119,120 Due to their appealing qualities, such as good biocompatibility, wearing comfort, low cost, as well as sensitivity to breathing activities in the aspect of lower frequencies and slight amplitude body motions, triboelectric nanogenerators (TENG) have recently been widely used for self-powered respiration monitoring. TENG-based respiration sensors can accurately and continuously track physiological respiratory behaviors and exhaled chemical regents for individualized health care.121 Further, the 'Liu' research group designed a respiration-driven triboelectric sensor (RTS) for concurrent biomechanical and biochemical breath measurement that can directly transform breath flow into electricity.122 The Wang research group also created an integrated triboelectric self-powered respiration sensor (TSRS) to track both human breathing patterns and the amount of NH3 in exhaled gases.123

'Su' and his research team created a triboelectric self-powered respiration sensor (TSRS) in 2019 to monitor human breathing patterns and the amount of NH3 in exhaled gases. Figure 3 shows a TSRS that is mounted to a person’s chest and tracks their breathing.124 Su et al. created an alveolus-inspired membrane sensor (AIMS) for self-powered wearable nitrogen dioxide detection and personal physiological evaluation in 2020. This sensor may be used to detect human breath activities for breath analysis.125

Figure 3.

Figure 3

Wearable TSRSs make it possible to monitor breathing in real time. (a) The stability of TSRS driven by regular breathing in humans. (b) The TSRS’s output voltage when four distinct breathing patterns are used. (c) Active monitoring of human breathing after 50 and 100 squats, correspondingly. (d) Evaluation of output signals both with and without NH3 injection from exhaled breath. Reprinted with permission from ref (124). Copyright 2019 Elsevier.

Using a self-validating 64-channel sensor array based on semiconducting single-walled carbon nanotubes, the Panes–Ruiz research group showed precise detection of H2S below breathing concentration levels in humid airflow (sc-SWCNTs). The repeatable sensor construction method is based on controlled multiplexing dielectrophoretic sc-SWCNT deposition. The sensing region is developed and produces gold nanoparticles that solve detection at room temperature by leveraging the affinity of gold and sulfur atoms in the gas. The estimated limit of detection (LOD) for sensing devices functionalized with optimum nanoparticle dispersion is 3 ppb, and their sensitivity is 0.122%/ppb. Our sensors surpass certain electrochemical sensors that seem to be available on the market in terms of improved stability and response levels despite self-validation.126

According to the kind and concentration of various surfactants, Chatterjee et al. investigated how several lung cancer VOCs found in human breath impacted the selectivity and sensitivity using CNT-based sensors (nonionic, cationic, and anionic). The performance of the sensors is found to be affected by both the interaction of the surfactant with the analyts and the supramolecular assembly with CNTs. Surfactant–CNT sensors were produced using the spray LbL approach. Water and other polar VOCs, particularly methanol and other alcohols, have demonstrated good sensitivity to CNT-DOC sensors. TX405-CNT sensors were sensitive to three chemicals: benzene, chloroform, and n-pentane. Except for isopropanol, the SDBS-CNT sensors could detect water, acetone, chloroform, and ethanol but not many biomarkers. N-Pentane, acetone, isoprene, and ethanol were all reactive substances for the BnzlkCl-CNT sensors. Although it was shown that pristine CNTs responded best to the majority of the aromatic VOCs in the collection, CTAB-CNT sensors were just moderately sensitive to most VOCs and lacked significant selectivity.127

Moreover, 'Sinha' research group used zinc oxide (ZnO) and CNTs to create a composite-based chemiresistive sensor. A mechanism has been linked to the VOC adsorption process’s temperature- and material-dependent switching. Due to their simplicity in low-temperature synthesis and the large variety of VOC sensing, stability, or other advantageous qualities, ZnO/CNT composites have been selected in this work as a crucial material above other metal oxides. Additionally, it is anticipated that a p–n heterojunction will develop at every point in which CNT and ZnO are in contact. This will also impact the composite sensor’s ability to sense. A potential is created as an outcome of the interaction of n-type (ZnO) and p-type (CNT) materials. As a result, if VOCs were adsorbed onto the surface of a composite sensor, their potential should decrease.128 The total sensor system also benefits from this drop in junction potential. Figure 4 depicts an electron transport route schematically in connection to the phenomenon.129

Figure 4.

Figure 4

CNT/ZnO composite sensor’s VOC detection technique in the low- and high-temperature regions. Reprinted with permission from ref (129). Copyright 2021 Elsevier.

For the purpose of detecting volatile VOCs, the Turner research group developed vapor quantum resistive sensors (vQRSs) utilizing random networks made of carbon nanotubes (CNTs) that have been functionalized with caffeine. Energy-dispersive X-ray spectroscopy (EDS) and a localized transmission electron microscopy (TEM) image both demonstrate that caffeine has moved to the CNT–CNT junction (CNTj). Furthermore, at the CNT intersection, we conducted localized EDS (Figure 5a,b). According to the elemental analysis, about 5% of atomic nitrogen is present. Such findings prove that caffeine molecules can cluster at the CNTj of random networks and thus are noncovalently bound to CNT surfaces.

Figure 5.

Figure 5

Caf-CNTj characterization: (a) high-Resolution TEM; (b) matching EDS spectra; and (inset) Caf-CNTs element table. Reprinted with permission from ref (130). Copyright 2020 Elsevier.

Specific VOCs like acetone, toluene, and ethanol have increased sensitivity and distinct selectivity and a quick response time with caffeine-enhanced CNT junction (Caf-CNTj) vQRSs. According to molecular dynamics simulations, the sensor’s sensitivity is diffusion-limited, with ethanol actively contributing (MD). The created Caf-CNTj sensors could be a strong contender for various activities performed by an electronic nose, including breath analysis and food quality monitoring.130

Bohli et al. created a room-temperature toluene and benzene sensor in 2019 using MWCNTs functionalized with a long-chain thiol self-assembled monolayer, 1-hexadecanethiol (HDT), and decorated with Au nanoparticles. The thiol monolayer adhering to the MWCNTs was examined using FT-IR and high-resolution TEM, which were used to describe the gold nanoparticle decorating. The ability of Au-MWCNT and HDT/Au-MWCNT sensors to detect VOCs at levels as low as ppm is evidence that the self-assembled layer enhances the sensing selectivity, sensitivity by a factor of 17, and response dynamics. The response of the Au-MWCNT sensor to various vapor injection amounts for toluene and benzene at room temperature is shown in Figure 6a,b.131

Figure 6.

Figure 6

(a, b) Au-MWCNT sensor response to injections of toluene and benzene at various concentrations. Reprinted with permission from ref (131). Copyright 2019 Belestein.

NiWO4 microflowers (MFs) were used to embellish MWCNTs in composites for the detection of NH3. The material’s porous nature, the large specific surface area, and the p–n heterojunction formed by the MWNTs and NiWO4 contribute to the better sensing capabilities of this composite. The sensor based on 10% (MWN10) has the best gas sensitivity, with a sensitivity of 13.07 to 50 ppm of NH3 at room temperature and a detecting lower limit of 20 ppm, according to the gas sensitivity of the sensor based on daisy-like NiWO4/MWCNTs.132

7. Graphene-Embedded Sensor for VOC Detection

Graphene has garnered a great deal of scientific attention since Novoselov and associates used mechanical exfoliation to synthesize it for the first time in 2004,133 winning them the Nobel Prize for their contribution to Physics in 2010.134 Pure graphene and graphene oxide (GO) are the family members of “graphene”. Graphene is a single layer of graphite formed of predictable hexagonal arrangements of 2D carbon atoms. This arrangement is identical to that of graphite. Graphene is relatively light, weighing just around 0.77 mg per square meter.135

Smart electronic material 'Graphene' has many unique properties and distinguishing characteristics. The structure’s extremely high electron mobility (200 000 cm2 V1 s1), much more significant than carbon nanotubes (CNTs), is made possible by its zero band gap.136 It possesses exceptional mechanical strength, an extremely high surface-to-volume ratio,137,138 high capacitance,139,140 outstanding thermal conductivity,141 exceptional electrical conductivity, and the possibility for atomically clean graphene sheets on the graphene lattice.142 In addition to such amazing properties, graphene also enables sensitive analyte detection because of its exceptionally low electrical noise.143 Functionalized reduced graphene oxide (RGO) VOC sensors for detecting the ppm level of VOCs was developed by Tombel et al. (Figure 7).144 Using a Ti/Pt integrated electrode and functionalized nanomaterial on RGO thin film, a single semiconductor sensor was assembled on a SiO2/Si substrate (IDE). The three VOCs that were the focus of the detection studies were acetone, toluene, and isoprene. These experiments were conducted at ambient temperatures of 30 °C and 40% relative humidity.145

Figure 7.

Figure 7

VOC sensor from the top. Reprinted with permission from ref (145). Copyright 2021 AIP Publishing.

Reduced graphene oxide (rGO)/metalloporphyrin-based VOC sensors are simple and affordable to make and can precisely detect VOCs linked to a number of diseases often present in human breath. They have been described by the Lee research group. Using a drop casting technique, we created an rGO-metalloporphyrin-sensing array and examined it with EDS and scanning electron microscopy (SEM). Next, three disease-related breath VOCs (acetone, ammonia, and isopropanol), as well as carbon monoxide, were introduced to the detecting array.146 ZnO nanowire–rGO nanocomposites are created by combining ZnO nanowires and graphene oxide (GO) in various ratios. This nanocomposite is utilized as a sensor to detect NH3, and it performs substantially superior to pure reduced graphene-oxide-based gas sensors, exhibiting an outstanding response (19.2%) to NH3 at ambient temperature.147

Gupta et al. investigated the use of thin coating rGO as a sensing material on a QCM sensor to detect VOC vapors at ambient temperature. To make the rGO suspension, the aqueous suspension of GO paste is first reduced with ascorbic acid. The rGO thin films were cast using the drop-cast technique and dried at room temperature. For their contaminant-free wrinkled form, multilayered graphene structure, high mobility of 2 × 104 cm2/(V s), and low resistivity of 4 × 10–1 cm, acetone molecules are more readily absorbed on the surface of rGO thin films. As a result of the rGO sensing material used in the QCM gas sensor, rapid reactions are possible. The QCM gas sensors’ quick reaction and recovery periods of 20–30 s have been proven using commercially available rGO thin sensing films.148

It is claimed that to produce nanostructured detectors GO nanodomains of p-type were carefully inserted into an n-type 3D ZnO nanoarchitecture. These ultraporous nanoheterojunction networks’ features were investigated using physical and chemical approaches, demonstrating how GO affects the networks’ ability to sense chemicals and light. By employing UV light activation of the sensing reactions, these nanocomposite materials were also employed to detect many common VOCs, including ethyl alcohol, propanone, and ethylbenzene, down to ambient temperature. Here, Figure 8a–d shows how exposure to UV light, temperature, and pure ZnO or 32:1 ZnO/GO films’ chemical sensing affects reactions to acetone.149

Figure 8.

Figure 8

(a–d) The responses of a pure ZnO sensor and a hybrid 32:1 ZnO/GO sensor to acetone concentrations ranging from 1 ppm to 20 ppb in simulated air (20% O2–80% N2). Reprinted with permission from ref (149). Copyright 2019 RSC.

The two main forms of sensitizing agents, N-doped graphene quantum dots (N-GQDs-NSs) coupled to nanosheets and N-doped carbon nanoparticles (N-CNPs) (Figure 9), have been produced from natural carbon powder utilizing the straightforward methods of oxidation and centrifuge separation. Analysis using the techniques of XRD, UV–vis, FT-IR, FE-SEM, XPS, Raman spectroscopy, AFM, HR-TEM, and FL revealed that nitrogen had been successfully incorporated into carbon nanoparticles, providing them an average plane dimension of 50 nm and a reasonably smooth surface. The capacity of the produced samples to identify volatile organic substances using a simple optical-fiber-based sensor setup was used to create and establish the samples’ adaptability for sensitizing agents. Comparative laboratory studies have explored that the recommended sensor’s efficiency can respond quickly within just a few tens of seconds when exposed to methanol vapor. When exposed to various alcohol vapors in an atmosphere with ambient air, their lower limits of detection were, respectively, 4.3, 4.9, and 10.5 ppm.150

Figure 9.

Figure 9

Experimental setup and potential utilization to evaluate the performance of polymer fiber optic sensor N-CNPs and N-GQD-NSs is schematically depicted. Reprinted with permission from ref (150). Copyright 2019 Springer Nature.

For rapid identification of VOCs, field effect transistors (FETs) with a p-TiO2 nanoparticle (NPs)/GO heterojunction are used. P-type anatase TiO2 NPs made from sol–gel were used as the channel materials and were implanted in a few stacked GO channels. To determine the ideal ratio of GO and TiO2 NPs within the hybrid channel, extensive microscopic, spectroscopic, and electrical characterizations were carried out. On a SiO2/Si substrate, a back-gated FET sensor was created and put through testing while being exposed to various VOCs. At 100 °C, tests with p-TiO2/GO FET sensors at zero gate voltage (VGS = 0) demonstrated ethanol selectivity. The lower detection limit of ethanol was enhanced to 500 ppb by utilizing an appropriate gate voltage (VGS > 0, near the Dirac point), significantly enhancing its sensitivity and p-TiO2–GO hybrid for field-assisted sensitivity amplification. Transient behavior in the presence of reducing vapor ethanol, on the other hand, served as evidence for the p-type conductance of the composites of p-TiO2 NPs and p-GO. The pure GO sensor (S9) could not respond at 100 ppm, but the S1–S8 sensors produced a response within the dynamic range of 25–300 ppm. Every sensor demonstrated a stable baseline with predictable sensing behavior (Figure 10).151

Figure 10.

Figure 10

Transient behavior of S1–S9 sensors in the 25–300 ppm range of ethanol concentration at 100 °C. VGS = 0 and VDS = 0.5. Reprinted with permission from ref (151). Copyright 2021 Elsevier.

A susceptible HCHO sensor is suggested and demonstrated using a heterostructure of the Sn3O4/rGO composite. It features a low working temperature and a wide detection range. The hydrothermal Sn3O4/rGO composite exhibits a large specific surface area and a microstructure like a flower. The sensing features of sensors based on pure Sn3O4 and Sn3O4/rGO composites are thoroughly analyzed to learn how the heterostructure influences the sensing performance of HCHO. According to the observed data, the Sn3O4/rGO composite sensor has a significantly higher sensing response at a lower working temperature than the pure Sn3O4 sensor. Figure 11a,b presents data analysis and the schematic configuration of the gas sensor.152

Figure 11.

Figure 11

Schematic structure of a whole gas sensor in (a) and the test and analysis system in (b). Reprinted with permission from ref (152). Copyright 2020 Elsevier.

As one of the most frequently harmful irritating gases, NO2 can cause asthma and other illnesses in people.153 Researchers are concentrating on creating light, affordable wearable electronics with excellent performance, and adaptability due to the recent rapid advancement of gas sensor research.154 Chen and a colleague created a smartphone-enabled, completely integrated wireless system for real-time NO2 monitoring using a flexible ZnS nanoparticle/nitrogen-doped reduced graphene oxide (ZnS NP/N-rGO) sensor. The device exhibited a response of 2.2–10 ppm of NO2, a rapid recovery performance of 724 s, a power consumption of 0.52 W, and an extremely low theoretical limit of detection (69 ppb) for graphene-based sensors.155

8. CD-Based Sensor for VOC Detection

For their excellent environmentally friendly nature, brightness, tunable fluorescence, inertness, low cost, straightforward synthetic route, and availability for a wide range of starting materials, carbon dots (Cdots), an emerging nanomaterial of the carbon-based materials family, have attracted significant research interest.156 For use in acetone gas-sensing applications, the Mishra research group reports producing ZnS quantum dots (QDs) via a hot-injection technique (Figure 12). The generated ZnS QDs were characterized using XRD and TEM analysis. The ZnS QD sensor exhibits rapid response and recovery times. At a 100 ppm acetone concentration at 175 °C, it also demonstrates outstanding stability, high sensitivity, and strong selectivity. Table 2 also compares the ZnS nanomaterial-based acetone sensor with earlier works or research tools based on metal-oxide nanoparticles. The ZnS QD sensor can be a viable sensor to detect acetone vapor from exhaled air for the invasive type 1 diabetes inspection.157

Figure 12.

Figure 12

Schematic picture of the hot-injection process used to create ZnS quantum dots (QDs) and an illustration of the ZnS crystal structure. Reprinted with permission from ref (157). Copyright 2021 MDPI.

Table 2. Comparative Study of ZnS-Based Nanosensors with Metal Oxide Nanosensors for Acetone Sensing.

ZnS-based sensor T/°C acetone (ppm) response time (s) recovery time (s) detection limit stability/selectivity ref
ZnS QDs 175 100 5.5 6.7s 12 ppm 89.1%/91.1% (157)
ZnS nanowire 320 100 ∼30 ∼45 –/21.1 (158)
Au/ZnS 260 100 –/– (159)
ZnO@ZnS core/shell 300 500 ∼10 ∼16 cycling for 55 s/- (160)
Cr2O3/ZnS 300 200 –/– (161)
20Nb/WO3 325 1 ∼7 ∼625 ∼14.3/22.3 (162)
TiO2/α-Fe2O3 300 300 ∼22 ∼86 ∼23/23 (163)
NiO/Zn2SnO4 300 100 ∼1 ∼60 <100 ppb ∼5 cycles for 1000 s/49.6 (164)
SnO2/ZnSnO3 290 300 ∼5 ∼115 ∼31/32 (165)

An acetone sensor was created using a quartz crystal microbalance (QCM), which was modified with graphene quantum dots (GQDs) to sense gases. GQDs were produced via citrate pyrolysis, and their characteristics were determined by high-resolution TEM (HR-TEM). They looked at the sensor’s gas sensitivity to acetone at low concentrations. With a sensitivity of 16.78 Hz/ppm and a minimal detection limit of 2.5 ppm, it demonstrated good linearity at less than 240 ppm acetone concentrations. The sensor showed high acetone selectivity in a combination of acetone, butanol, and isopropanol (Figure 13). The same sensor’s response time was constant for varying acetone concentrations, and the response and recovery durations were 32 and 48 s, respectively.166

Figure 13.

Figure 13

Schematic of VOC detection device. Reprinted with permission from ref (166). Copyright 2021.

9. Other Carbon Nanomaterial-Based Sensors for the Detection of VOCs

The gas-sensing properties of MXene may be predicted using the quantum mechanical approach and basic principles based on density functional theory (DFT). Using the Schrödinger equation, DFT-based first-principles examine the electronic structure of materials.167 The absorption energy of material to certain gas molecules, as well as the associated charge transfer characteristics, are important factors in predicting sensing qualities based on gas sensing. Yu et al. performed the first computational analysis to demonstrate MXenes’ potential as a gas sensor or capturer, employing monolayer Ti2CO2 for NH3 detection.168 Because metallic Ti2C(OH)2, Ti2CF2, and Ti2C do not have semiconducting properties, only Ti2CO2 was studied using first-principles modeling. Using the most stable structure, the O-functionalized monolayer of Ti2CO2, gas molecules’ possible absorption positions were investigated to compute adsorption energy and charge transfer. Aside from demonstrating gas absorption and charge transfer in MXenes, plane-wave-based DFT computation allows for the tailoring of MXene selectivity via controlled oxygen functionalization. Junkaew et al. investigated gas molecule adsorption behavior on four MXenes (M2C, M = Ti, V, Nb, Mo) and their O-terminated surfaces with electronic charge characteristics.168 When a gas molecule is absorbed, it retains its molecule form or is dissociated on the surface of the MXenes. Theoretical simulations predicted varied absorption energies of multiple gaseous species on various types of MXene surface conditions and their associated charge transfer for gas-sensing properties. However, just a few laboratory experiments on the sensing properties of different MXenes have been published. We report a novel experimental investigation into the gas-sensing properties of MXenes. Layered transition metal carbides, nitrides, or MXene are composed of more than 60 different types, including Ti2C, V2C, Nb2C, Ti4N3, TiNbC, Ti3CN, and Mo2TiC2.169171 By chemically treating precursors of the Mn+1AXn phase, also known as the MAX phase, where M is an early transition metal, A is typically an element IIIA or IVA; X is C and/or N; and n = 1, 2, and 3 MXenes are produced.172 MXenes are a focus of research because of a number of intriguing characteristics that promise to lead the way for future sensor development. A fascinating class of two-dimensional materials called MXenes has lately attracted interest for their possible use in gas sensing.173

Here, we show the efficiency of a new Mo2CTx MXene sensor in detecting VOCs. On a Si/SiO2 substrate, the suggested sensor is a chemiresistive device made by photolithography. The effectiveness of the sensor is examined in relation to different MXene manufacturing conditions. Different VOC concentrations, including ethanol, toluene, benzene, acetone, and methanol, are examined at room temperature. The effectiveness of several MXene-based gas sensors is compared in Table 3.174

Table 3. Comparison of MXene-Based Gas Sensors at Room Temperature (RT).

MXene sensor operating temperature VOCs with highest sensing response LOD cross-sensitivity ratio (at 100 ppm) ref
Mo2CTx RT toluene 220 ppb 2.65 (175)
Ti3C2Tx RT NH3 1.47 (176)
Ti3C2Tx RT C2H5OH 50 ppb 1.75 (177)
V2 C2Tx RT C2H5OH 3.61 (178)
CuO/Ti3C2Tx 250 °C toluene 320 ppb (179)
S-doped Ti3C2Tx RT toluene (180)

In this case, the model materials for hybridization and their application to the detection of different volatile organic chemicals are Ti3C2Tx and WSe2 (Figure 14). The Ti3C2Tx/WSe2 hybrid sensor has excellent adaptability for a wide range of volatile organic compounds, low noise levels, and exceptionally fast response and recovery times. The hybrid sensor’s sensitivities to ethanol have risen about 12-fold when compared to pure Ti3C2Tx. Furthermore, by restricting the contact of water molecules from Ti3C2Tx’s edges, the hybridization process offers a successful defense against MXene oxidation. Detecting volatile organic molecules containing oxygen is highly sensitive and selective, and an enhanced method for Ti3C2Tx/WSe2 heterostructured materials is presented.173

Figure 14.

Figure 14

(a) Ti3C2Tx/WSe2 hybridization and sensor construction. Ti3C2Tx/WSe2 nanohybrid preparation procedures are shown schematically. (b) Schematic representation of inkjet-printed gas sensors used in a wireless monitoring system to detect volatile organic chemicals: WSe2, CTA + −WSe2, and Ti3C2Tx dispersions. (c) Zeta potential distributions. Reprinted with permission from ref (173). Copyright 2020 Springer Nature.

It is highlighted how well the Zhao research team’s novel MXene, V4C3Tx, functions as an acetone sensor. It was made from V4AlC3 by selectively etching the Al layer with aqueous HF RT. A V4C3Tx-based acetone sensor operates well due to its low operating temperature of 25 °C, low detection limit of 1 ppm below the 1.8 ppm diabetes diagnostic threshold, and good selectivity toward acetone in a mixed mixture of water vapor and acetone. It might lead to a much faster and earlier diagnosis of diabetes. V4C3Tx MXene is used for the first time in this work in the context of acetone detection.180

We provide a simple solvothermal method for fabricating W18O49/Ti3C2Tx composites that build 1D W18O49 nanorods (NRs) upon the surfaces of 2D Ti3C2Tx Mxene sheets. The 1D W18O49 NRs were uniformly distributed throughout the surfaces of a 2D Ti3C2Tx sheet to produce the 1D/2D W18O49/Ti3C2Tx hybrids. Comparing the W18O49/Ti3C2Tx composites to the Ti3C2Tx sheets and W18O49 NRs, it was discovered that they markedly increased the ability to sense acetone. The hybrid sensor demonstrated a robust response to low acetone concentrations, rapid response and recovery times, excellent selectivity, long-term stability, and an extremely low acetone detection limit. The superb sensing performance of the hybrid sensor can be attributed to the uniform dispersion of W18O49 NRs on the surface, the elimination of fluorine-containing groups during the solvothermal process, and the beneficial interactions at the interface between the W18O49/Ti3C2Tx sheets and W18O49 NRs.181

Schottky-barrier-equipped Ti3C2Tx–ZnO nanosheet hybrids have been produced for NO2 recovery and detection under UV light. By using HF solution to etch the Ti3AlC2 Al layer, Ti3C2Tx nanosheets were produced (MAX). ZnO nanosheets’ porous structure is crucial for the adsorption of NO2 gas. With the aid of UV irradiation during the recovery process, the Ti3C2Tx–ZnO nanosheet-based gas sensor displayed better NO2-detecting skills, including a high sensitivity of 367.63% to 20 ppm of NO2 and a quicker response/recovery time of 22 s/10 s.182

The first oxygenated amorphous carbon (a-COx)/graphite (G) nanofilament-based buckypaper sensor was developed by Homaeigohar (Figure 15). By forming hydrogen bonds, oxygen-containing groups influence the group’s capacity to extract electrons, the density of hole carriers, and, therefore, the resistivity, which in turn affect the group’s selectivity toward VOCs like ethanol and acetone. However, the creation of charge-transfer complexes caused by the toluene aromatic ring’s electrostatic interactions with the electrons of the graphitic crystals may be the main cause of the sensor’s great responsiveness to nonpolar toluene.183

Figure 15.

Figure 15

Electrically conductive oxygenated amorphous carbon (a-COx)/graphite (G) (a-COx/G) nanofibers’ ability for sensing VOCs. Reprinted with permission from ref (183). Copyright 2019 MDPI.

10. Conclusions, Outlook, and Perspective

The development of VOC sensing utilizing CNMs is outlined in this comprehensive review. Due to the wide range of uses for VOC environmental monitoring, the difficulty of selecting VOC detection is growing every day. Toxic VOC detection is crucial for protecting the environment and human health. Due to their unique morphology and features, developing CNMs has been the foundation for sensor technology over the preceding two decades. CNMs like graphene, CNTs, and their derivatives, which have a large number of adsorption sites, low density, tunable electrical properties, high carrier mobility, low operating temperatures, longer lives, and ease of recovery, are appropriate sensing materials for a variety of toxic pollutant gases and volatile organic compounds. Numerous researchers have observed an increase in sensitivity that allows for detecting harmful VOCs at ppb levels. The selectivity of a CNM-based sensor toward a specific VOC at room temperature has been shown in numerous papers. According to the literature, metal oxide and nanocarbon hybrids have outstanding sensing capabilities, with nanocarbon serving as the primary component. Composites made of nanocarbon have shown a strong potential for use in VOC sensing.

However, some measure of research should be focused on the synthesis or design of porous or hierarchical structures of carbon-based materials for the suitable VOC molecule adsorption with more reaction sites which can help with a better understanding of structural-adsorption/absorption property relationships that will garnish sensing mechanisms. Also, by analysis of theoretical calculation, the change of electronic properties of host–guest items can be explored for selective gas sensors.

Acknowledgments

The author Dr S. Chakroborty is grateful to the experimental Research lab at IES University Bhopal and Dr. Kaushik Pal should sincerely acknowledge University Centre for Research and Development (UCRD) at Chandigarh University, Mohali, Punjab. Most distinguished Prof. Fernando Gomes de Souza junior highly acknowledges his research funding FAPERJ-CNE E-26/201.154/2021 to support this work.

The authors declare no competing financial interest.

References

  1. Faroon O.; Taylor J.; Roney N.; Fransen M. E.; Bogaczyk S.; Diamond G.. Toxicological Profile for Carbon Tetrachloride; Department of Health and Human Services, Public Health Service Agency for Toxic Substances and Disease Registry: Atlanta, GA, USA, 2005. [Google Scholar]
  2. Pandey P.; Yadav R. A Review on Volatile Organic Compounds (VOCs) as Environmental Pollutants: Fate and Distribution. Int. J. Plant Environ. 2018, 4, 14–26. 10.18811/ijpen.v4i02.2. [DOI] [Google Scholar]
  3. Choi S.; Kim N.; Cha H.; Ha R. Micro sensor node for air pollutant monitoring: hardware and software issues. Sensors 2009, 9, 7970–7987. 10.3390/s91007970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Sax N.; Lewis R.. Hawleys Condensed Chemical Directory, 11th ed.; VNR: New York, 1987; pp 177–178, 704–969. [Google Scholar]
  5. Lim C.; Yu J.; Kin D.; Byun H.; Lee D.. IEEE Sensors J. 2006. [Google Scholar]
  6. Mizukoshi A.; Kumagai K.; Yamamoto N.; Noguchi M.; Yoshiuchi K.; Kumano H.; Yanagisawa Y. A novel methodology to evaluate health impacts caused by voc exposures using real-time VOC and holter monitors. Int. J. Environ. Res. Public Health 2010, 7, 4127–4138. 10.3390/ijerph7124127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Xu Z.; Mulchandani A.; Chen W. Detection of Benzene, toluene, ethyl benzene, and xylenes (BTEX) using toluene dioxygenase-peroxidase coupling reactions. Biotechnol. Prog. 2003, 19, 1812–1815. 10.1021/bp0341794. [DOI] [PubMed] [Google Scholar]
  8. Ghanem M.; Guo Y.; Hassard J.; Osmond M.; Richards M.. Sensor Grids for Air Pollution Monitoring; 2004. https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.127.1135&rep=rep1&type=pdf.
  9. Henriksen T.; Dahlback A.; Larsen S.; Moan J. Ultraviolet radiation and skin cancer: Effect of an ozone layer depletion. Photochem. photobiol. 1990, 51, 579–582. 10.1111/j.1751-1097.1990.tb01968.x. [DOI] [PubMed] [Google Scholar]
  10. Peng G.; Tisch U.; Adams O.; Hakim M.; Shehada N.; Broza Y. Y.; Billan S.; Abdah-Bortnyak R.; Kuten A.; Haick H. Diagnosing lung cancer in exhaled breath using gold nanoparticles. Nat. Nanotechnol. 2009, 4, 669–673. 10.1038/nnano.2009.235. [DOI] [PubMed] [Google Scholar]
  11. Kumar V.; Kim K.-H.; Kumar P.; Jeon B.-H.; Chun J.; Kim J.-C. Functional hybrid nanostructure materials: advanced strategies for sensing applications toward volatile organic compounds. Coord. Chem. Rev. 2017, 342, 80–105. 10.1016/j.ccr.2017.04.006. [DOI] [Google Scholar]
  12. Zhou X.; Xue Z.; Chen X.; Huang C.; Bai W.; Lu Z.; Wang T. Nanomaterial based gas sensors used for breath diagnosis. J. Mater. Chem. B 2020, 8, 3231–3248. 10.1039/C9TB02518A. [DOI] [PubMed] [Google Scholar]
  13. Bruderer T.; Gaisl T.; Gaugg M. T.; Nowak N.; Streckenbach B.; Müller S.; Moeller A.; Kohler M.; Zenobi R. On-line analysis of exhaled breath: focus review. Chem. Rev. 2019, 119, 10803–10828. 10.1021/acs.chemrev.9b00005. [DOI] [PubMed] [Google Scholar]
  14. Lee J.; Jung M.; Barthwal S.; Lee S.; Lim S.-H. MEMS gas preconcentrator filled with CNT foam for exhaled VOC gas detection. Biochip J. 2015, 9, 44–49. 10.1007/s13206-014-9106-y. [DOI] [Google Scholar]
  15. Tomic M.; Setka M.; Vojkuvka L.; Vallejos S. VOCs sensing by metal oxides, conductive polymers, and carbon-based materials. Nanomaterials 2021, 11, 552–585. 10.3390/nano11020552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Tripathi K.M.; Kim T.; Losic D.; Tung T. T. Recent advances in engineered graphene and composites for detection of volatile organic compounds (VOCs) and noninvasive diseases diagnosis. Carbon 2016, 110, 97–129. 10.1016/j.carbon.2016.08.040. [DOI] [Google Scholar]
  17. Chen W. Y.; Lai S.-N.; Yen C.-C.; Jiang X.; Peroulis D.; Stanciu L. A. Surface functionalization of Ti3C2Tx MXene with highly reliable super hydrophobic protection for volatile organic compounds sensing. ACS Nano 2020, 14, 11490–11501. 10.1021/acsnano.0c03896. [DOI] [PubMed] [Google Scholar]
  18. Chen Q.; Liu D.; Lin L.; Wu J. Bridging interdigitated electrodes by electrochemical-assisted deposition of graphene oxide for constructing flexible gas sensor. Sens. Actuator. B Chem. 2019, 286, 591–599. 10.1016/j.snb.2019.02.024. [DOI] [Google Scholar]
  19. Scida K.; Stege P. W.; Haby G.; Messina G. A.; García C. D. Recent applications of carbon-based nanomaterials in analytical chemistry: critical review. Anal. Chim. Acta 2011, 691, 6–17. 10.1016/j.aca.2011.02.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Llobet E. Gas sensors using carbon nanomaterials: A review. Sens. Actuators B Chem. 2013, 179, 32–45. 10.1016/j.snb.2012.11.014. [DOI] [Google Scholar]
  21. Homaeigohar S.; Elbahri M. Graphene membranes for water desalination. NPG Asia Mater. 2017, 9, e427 10.1038/am.2017.135. [DOI] [Google Scholar]
  22. Homaeigohar S.; Davoudpour Y.; Habibi Y.; Elbahri M. The Electrospun Ceramic Hollow Nanofibers. Nanomaterials 2017, 7, 383–415. 10.3390/nano7110383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Dey A. Semiconductor metal oxide gas sensors: a review. Mater. Sci. Eng., B 2018, 229, 206–217. 10.1016/j.mseb.2017.12.036. [DOI] [Google Scholar]
  24. Zappa D.; Galstyan V.; Kaur N.; Arachchige H. M. M.; Sisman O.; Comini E. Metal oxide-based heterostructures for gas sensors-A review. Anal. Chim. Acta 2018, 1039, 1–23. 10.1016/j.aca.2018.09.020. [DOI] [PubMed] [Google Scholar]
  25. Liu S.; Zeng T. H.; Hofmann M.; Burcombe E.; Wei J.; Jiang R.; Kong J.; Chen Y. Antibacterial Activity of Graphite, Graphite Oxide, Graphene Oxide, and Reduced Graphene Oxide: Membrane and Oxidative Stress. ACS Nano 2011, 5, 6971–6980. 10.1021/nn202451x. [DOI] [PubMed] [Google Scholar]
  26. Su Y.; Kravets V. G.; Wong S. L.; Waters J.; Geim A. K.; Nair R. R. Impermeable barrier films and protective coatings based on reduced graphene oxide Nat. Commun. 2014, 5, 4843. 10.1038/ncomms5843. [DOI] [PubMed] [Google Scholar]
  27. Meyyappan M.; Delzeit L.; Cassell A.; Hash D. Carbon nanotube growth by PECVD: a review. Plasma Sources Sci. Technol. 2003, 12, 205–216. 10.1088/0963-0252/12/2/312. [DOI] [Google Scholar]
  28. Dresselhaus M. S.; Dresselhaus G.; Jorio A. Unusual properties and structure of carbon nanotubes. Annu. Rev. Mater. Res. 2004, 34, 247–278. 10.1146/annurev.matsci.34.040203.114607. [DOI] [Google Scholar]
  29. Baker S. N.; Baker G. A. Angew. Chem., Int. Ed. 2010, 49, 6726–6744. 10.1002/anie.200906623. [DOI] [PubMed] [Google Scholar]
  30. Lin L.; Rong M.; Luo F.; Chen D.; Wang Y.; Chen X. TrAC, Trends Anal. Chem. 2014, 54, 83–102. 10.1016/j.trac.2013.11.001. [DOI] [Google Scholar]
  31. Li H.; Kang Z.; Liu Y.; Lee S. T. Carbon nanodots: synthesis, properties and applications. J. Mater. Chem. 2012, 22, 24230–24255. 10.1039/c2jm34690g. [DOI] [Google Scholar]
  32. Bo Z.; Shuai X.; Mao S.; Yang H.; Qian J.; Chen J.; Yan J.; Cen K. Green preparation of reduced graphene oxide for sensing and energy storage Applications. Sci. Report 2015, 4, 4684–4691. 10.1038/srep04684. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Ma J.; Wang J. N. Purification of Single-Walled Carbon Nanotubes by a Highly Efficient and Nondestructive Approach. Chem. Mater. 2008, 20, 2895–2902. 10.1021/cm8001699. [DOI] [Google Scholar]
  34. Perikala M.; Bhardwaj A. Highly Stable White-Light-Emitting Carbon Dot Synthesis Using a Non-coordinating Solvent. ACS Omega 2019, 4, 21223–21229. 10.1021/acsomega.9b02686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Geim A. K.; Novoselov K. S. The rise of graphene. Nanosci. Technol. A Collect. Rev. Nat. J. 2007, 6, 11–19. 10.1142/9789814287005_0002. [DOI] [PubMed] [Google Scholar]
  36. Yu W.; Sisi L.; Haiyan Y.; Jie L. Progress in the functional modification of graphene/graphene oxide: A review. RSC Adv. 2020, 10, 15328–15345. 10.1039/D0RA01068E. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Park S.; Ruoff R. S. Chemical methods for the production of graphenes. Nat. Nanotechnol. 2009, 4, 217–224. 10.1038/nnano.2009.58. [DOI] [PubMed] [Google Scholar]
  38. Hazra A.; Samane N.; Basu S. A.. A Review on Metal Oxide-Graphene Derivative Nano-Composite Thin Film Gas Sensors. Multilayer Thin Films - Versatile Applications for Materials Engineering; InTechOpen, 2020. 10.5772/intechopen.90622. [DOI] [Google Scholar]
  39. Sharma N.; Sharma V.; Jain Y.; Kumari M.; Gupta R.; Sharma S. K.; Sachdev K. Synthesis and Characterization of Graphene Oxide (G.O.) and Reduced Graphene Oxide (rGO) for Gas Sensing Application. Macromol. Symp. 2017, 376 (1), 1700006–1700010. 10.1002/masy.201700006. [DOI] [Google Scholar]
  40. Lucas A.; Lambin P.; Smalley R. On the energetics of tubular fullerenes. J. Phys. Chem. Solids 1993, 54, 587–593. 10.1016/0022-3697(93)90237-L. [DOI] [Google Scholar]
  41. Qi P.; Vermesh O.; Grecu M.; Javey A.; Wang Q.; Dai H.; Peng S.; Cho K. J. Toward large arrays of multiplex functionalized carbon nanotube sensors for highly sensitive and selective molecular detection. Nano Lett.. 2003, 3, 347–351. 10.1021/nl034010k. [DOI] [PubMed] [Google Scholar]
  42. Pradhan B. Mater. Res. Soc. Symp. Proc. 2001, 633. 10.5772/intechopen.85045. [DOI] [Google Scholar]
  43. Larson D. R.; Zipfel W. R.; Williams R. M.; Clark S. W.; Bruchez M. P.; Wise F. W.; Webb W. W. Water-soluble quantum dots for multiphoton fluorescence imaging in vivo. Science 2003, 300, 1434–1436. 10.1126/science.1083780. [DOI] [PubMed] [Google Scholar]
  44. Hu S.; Tian R.; Wu L.; Zhao Q.; Yang J.; Liu J.; Cao S. Chemical regulation of carbon quantum dots from synthesis to photocatalytic activity. Chemistry-An Asian Journal 2013, 8, 1035–1041. 10.1002/asia.201300076. [DOI] [PubMed] [Google Scholar]
  45. Park Y.; Yoo J.; Lim B.; Kwon W.; Rhee S.-W. Improving the functionality of carbon nanodots: doping and surface functionalization. J. Mater. Chem. 2016, 4, 11582–11603. 10.1039/C6TA04813G. [DOI] [Google Scholar]
  46. McQuade D. T.; Hegedus A. H.; Timothy M.; Swager T. M. Signal Amplification of a ″Turn-On″ Sensor: Harvesting the Light Captured by a Conjugated Polymer. J. Am. Chem. Soc. 2000, 122 (49), 12389–12390. 10.1021/ja003255l. [DOI] [Google Scholar]
  47. Wang X.; Sun X.; Hu P. A.; Zhang J.; Wang L.; Feng W.; Lei S.; Yang B.; Cao W. Colorimetric sensor based on self-assembled polydiacetylene/grapheme stacked composite film for vapor-phase volatile organic compounds. Adv. Funct. Mater. 2013, 23, 6044–6050. 10.1002/adfm.201301044. [DOI] [Google Scholar]
  48. Shi J.; Zhu Y.; Zhang X.; Baeyens W. R. G.; Garcıa-Campana A. M. Recent developments in nanomaterial optical sensors. Trends Anal. Chem. 2004, 23, 351–360. 10.1016/S0165-9936(04)00519-9. [DOI] [Google Scholar]
  49. Song Y.; Wei W.; Qu X. Colorimetric biosensing using smart materials. Adv. Mater. 2011, 23, 4215–4236. 10.1002/adma.201101853. [DOI] [PubMed] [Google Scholar]
  50. Mehta B.; Benkstein K. D.; Semancik S.; Zaghloul M. E. Gas sensing with bare and graphene-covered optical nano-antenna structures. Sci. Rep 2016, 6, 21287. 10.1038/srep21287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Wu Z.; Chen X.; Zhu S.; Zhou Z.; Yao Y.; Quan W.; Liu B. Enhanced sensitivity of ammonia sensor using graphene/polyanilinenanocomposite. Sensor. Actuator. B Chem. 2013, 178, 485–493. 10.1016/j.snb.2013.01.014. [DOI] [Google Scholar]
  52. Zhang D.; Wang D.; Li P.; Zhou X.; Zong X.; Dong G. Facile fabrication of high performance QCM humidity sensor based on layer-by-layer self-assembled polyaniline/graphene oxide nanocomposite film. Sensor Actuator. B Chem. 2018, 255, 1869–1877. 10.1016/j.snb.2017.08.212. [DOI] [Google Scholar]
  53. Chen W. Y.; Jiang X.; Lai S.-N.; Peroulis D.; Stanciu L. Nanohybrids of aMXene and transition metal dichalcogenide for selective detection of volatile organic compounds. Nat. Commun. 2020, 11, 1302. 10.1038/s41467-020-15092-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Guo H.; Guo A.; Gao Y.; Liu T. Optimization of a VOC sensor with a bilayered diaphragm using FBAR as strain sensing elements. Sensors 2017, 17, 1764. 10.3390/s17081764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Zhang Y.; Luo J.; Flewitt A. J.; Cai Z.; Zhao X. Film bulk acoustic resonators (FBARs) as biosensors: a review. Biosens. Bioelectron 2018, 116, 1–15. 10.1016/j.bios.2018.05.028. [DOI] [PubMed] [Google Scholar]
  56. Kaur A.; Singh I.; Kumar J.; Bhatnagar C.; Dixit S. K.; Bhatnagar P. K.; Mathur P. C.; Covas J. A.; da Conceicao Paiva M. Enhancement in the performance of multiwalled carbon nanotube: poly (methylmethacrylate) composite thin film ethanol sensors through appropriate nanotube functionalization. Mater. Sci. Semicond. Process. 2015, 31, 166–174. 10.1016/j.mssp.2014.11.030. [DOI] [Google Scholar]
  57. Tung T. T.; Castro M.; Kim T. Y.; Suh K. S.; Feller J.-F. High stability silver nanoparticles-graphene/poly (ionic liquid)-based chemoresistive sensors for volatile organic compounds’ detection. Anal. Bioanal. Chem. 2014, 406, 3995–4004. 10.1007/s00216-013-7557-y. [DOI] [PubMed] [Google Scholar]
  58. Elakia M.; Gobinath M.; Sivalingam Y.; Palani E.; Ghosh S.; Nutalapati V.; Surya V. J. Investigation on visible light assisted gas sensing ability of multiwalled carbon nanotubes coated with pyrene based organic molecules. Phys. E Low-dimens. Syst. Nanostruct. 2020, 124, 114232. 10.1016/j.physe.2020.114232. [DOI] [Google Scholar]
  59. Hussain T.; Sajjad M.; Singh D.; Bae H.; Lee H.; Larsson J. A.; Ahuja R.; Karton A. Sensing of volatile organic compounds on two-dimensional nitrogenated holey graphene, graphdiyne, and their heterostructure. Carbon 2020, 163, 213–223. 10.1016/j.carbon.2020.02.078. [DOI] [Google Scholar]
  60. Collins P. G.; Bradley K.; Ishigami M.; Zettl A. Extreme Oxygen Sensitivity of Electronic Properties of Carbon Nanotubes. Science 2000, 287, 1801–1804. 10.1126/science.287.5459.1801. [DOI] [PubMed] [Google Scholar]
  61. Kong J.; Franklin N. R.; Zhou C.; Chapline M. G.; Peng S.; Cho K.; Dai H. Nanotube Molecular Wires as Chemical Sensors. Science 2000, 287, 622–625. 10.1126/science.287.5453.622. [DOI] [PubMed] [Google Scholar]
  62. Snow E. S.; Perkins F. K.; Houser E. J.; Badescu S. C.; Reinecke T. L. Chemical detection with a single-walled carbon nanotube capacitor. Science 2005, 307, 1942–1945. 10.1126/science.1109128. [DOI] [PubMed] [Google Scholar]
  63. Zheng W.; Xu Y.; Zheng L.; Yang C.; Pinna N.; Liu X.; Zhang J. MoS2 Van der Waals p-n Junctions Enabling Highly Selective Room-Temperature NO2 Sensor. Adv. Funct. Mater. 2020, 30, 2000435. 10.1002/adfm.202000435. [DOI] [Google Scholar]
  64. Zhang N.; Yu K.; Li Q.; Zhu Q. Z.; Wan Q. Room-temperature high-sensitivity H2S gas sensor based on dendritic ZnO nanostructures with macroscale in appearance. Appl. Phys. 2008, 103, 104305. 10.1063/1.2924430. [DOI] [Google Scholar]
  65. Mishra R. K.; Choi G.-J.; Choi H.-J.; Gwag J. S. ZnS Quantum Dot Based Acetone Sensor for Monitoring Health-Hazardous Gases in Indoor/Outdoor Environment. Micromachines (Basel) 2021, 12, 598–609. 10.3390/mi12060598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Saito R.; Dresselhaus G.; Dressealhaus M. S.. Physical properties of carbon nanotubes; Imperial College Press: London, 1998; p 259. [Google Scholar]
  67. Bondavalli P.; Legagneux P.; Pribat D. Carbon nanotubes based transistors as gas sensors: State of the art and critical review. Sensors and Actuators B-Chemical 2009, 140, 304–18. 10.1016/j.snb.2009.04.025. [DOI] [Google Scholar]
  68. Panda A.; Arumugasamy S. K.; Lee J.; Son Y.; Yun K.; Venkateswarlu S.; Yoon M. Chemical-free sustainable carbon nano-onion as a dual-mode sensor platform for noxious volatile organic compounds. Appl. Surf. Sci. 2021, 537, 147872. 10.1016/j.apsusc.2020.147872. [DOI] [Google Scholar]
  69. Ruzsanyi V.; Wiesenhofer H.; Ager C.; Herbig J.; Aumayr G.; Fischer M.; Renzler M.; Ussmueller T.; Lindner K.; Mayhew C. A portable sensor system for the detection of human volatile compounds against transnational crime. Sens. Actuator. B Chem. 2021, 328, 129036–129044. 10.1016/j.snb.2020.129036. [DOI] [Google Scholar]
  70. Andre L.; Desbois N.; Gros C. P.; Brandes S. Porous materials applied to biomarker sensing in exhaled breath for monitoring and detecting noninvasive pathologies. Dalton Trans. 2020, 49, 15161–15170. 10.1039/D0DT02511A. [DOI] [PubMed] [Google Scholar]
  71. Stewart K.; Limbu S.; Nightingale J.; Pagano K.; Park B.; Hong S.; Lee K.; Kwon S.; Kim J.-S. Molecular understanding of a p-conjugated polymer/solid-state ionic liquid complex as a highly sensitive and selective gas sensor. J. Mater. Chem. C 2020, 8, 15268–15276. 10.1039/D0TC03093G. [DOI] [Google Scholar]
  72. Das T.; Mohar M. Development of a smartphone-based real time costeffective VOC sensor. Heliyon 2020, 6, e05167 10.1016/j.heliyon.2020.e05167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Tripathi K.M.; Kim T.; Losic D.; Tung T. T. Recent advances in engineered graphene and composites for detection of volatile organic compounds (VOCs) and noninvasive diseases diagnosis. Carbon 2016, 110, 97–129. 10.1016/j.carbon.2016.08.040. [DOI] [Google Scholar]
  74. Vishinkin R.; Haick H. Nanoscale sensor technologies for disease detection via volatolomics. Small 2015, 11, 6142–6164. 10.1002/smll.201501904. [DOI] [PubMed] [Google Scholar]
  75. Ahmad R.; Majhi S. M.; Zhang X.; Swager T.M.; Salama K. N. Recent progress and perspectives of gas sensors based on vertically oriented ZnOnanomaterials. Adv. Colloid Interface Sci. 2019, 270, 1–27. 10.1016/j.cis.2019.05.006. [DOI] [PubMed] [Google Scholar]
  76. Wang X.; Sun X.; Hu P. A.; Zhang J.; Wang L.; Feng W.; Lei S.; Yang B.; Cao W. Colorimetric sensor based on self-assembled polydiacetylene/grapheme stacked composite film for vapor-phase volatile organic compounds. Adv. Funct. Mater. 2013, 23, 6044–6050. 10.1002/adfm.201301044. [DOI] [Google Scholar]
  77. Zhihua Y.; Liang Z.; Kaixin S.; Weiwei H. Characterization of quartz crystal microbalance sensors coated with graphene films. Procedia Eng. 2012, 29, 2448–2452. 10.1016/j.proeng.2012.01.330. [DOI] [Google Scholar]
  78. Inyawilert K.; Wisitsoraat A.; Sriprachaubwong C.; Tuantranont A.; Phanichphant S.; Liewhiran C. Rapid ethanol sensor based on electrolytically-exfoliated graphene-loaded flame-made In-doped SnO2 composite film. Sensor. Actuator. B Chem. 2015, 209, 40–55. 10.1016/j.snb.2014.11.086. [DOI] [PubMed] [Google Scholar]
  79. Shi J.; Zhu Y.; Zhang X.; Baeyens W. R. G.; Garcıa-Campana A. M. Recent developments in nanomaterial optical sensors. Trends Anal. Chem. 2004, 23, 351–360. 10.1016/S0165-9936(04)00519-9. [DOI] [Google Scholar]
  80. Song Y.; Wei W.; Qu X. Colorimetric biosensing using smart materials. Adv. Mater. 2011, 23, 4215–4236. 10.1002/adma.201101853. [DOI] [PubMed] [Google Scholar]
  81. Mehta B.; Benkstein K. D.; Semancik S.; Zaghloul M. E. Gas sensing with bare and graphene-covered optical nano-antenna structures. Sci. Rep 2016, 6, 21287. 10.1038/srep21287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Wu Z.; Chen X.; Zhu S.; Zhou Z.; Yao Y.; Quan W.; Liu B. Enhanced sensitivity of ammonia sensor using graphene/polyanilinenanocomposite. Sensor. Actuator. B Chem. 2013, 178, 485–493. 10.1016/j.snb.2013.01.014. [DOI] [Google Scholar]
  83. Zhang D.; Wang D.; Li P.; Zhou X.; Zong X.; Dong G. Facile fabrication of high performance QCM humidity sensor based on layer-by-layer self-assembled polyaniline/graphene oxide nanocomposite film. Sensor Actuator. B Chem. 2018, 255, 1869–1877. 10.1016/j.snb.2017.08.212. [DOI] [Google Scholar]
  84. Chen W.; Wang Z.; Gu S.; Wang M. J.; Wang Y.; Wei Z. Hydrophobic aminofunctionalizedgraphene oxide nanocomposite for aldehydes detection in fish fillets. Sensor Actuator B Chem. 2020, 306, 127579. 10.1016/j.snb.2019.127579. [DOI] [Google Scholar]
  85. Guo H.; Guo A.; Gao Y.; Liu T. Optimization of a VOC sensor with a bilayered diaphragm using FBAR as strain sensing elements. Sensors 2017, 17, 1764. 10.3390/s17081764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Zhang Y.; Luo J.; Flewitt A. J.; Cai Z.; Zhao X. Film bulk acoustic resonators (FBARs) as biosensors: a review. Biosens, Bioelectron. 2018, 116, 1–15. 10.1016/j.bios.2018.05.028. [DOI] [PubMed] [Google Scholar]
  87. Hussain T.; Sajjad M.; Singh D.; Bae H.; Lee H.; Larsson J. A.; Ahuja R.; Karton A. Sensing of volatile organic compounds on two-dimensional nitrogenated holey graphene, graphdiyne, and their heterostructure. Carbon 2020, 163, 213–223. 10.1016/j.carbon.2020.02.078. [DOI] [Google Scholar]
  88. Kaur A.; Singh I.; Kumar J.; Bhatnagar C.; Dixit S. K.; Bhatnagar P. K.; Mathur P. C.; Covas J. A.; da Conceicao Paiva M. Enhancement in the performance of multiwalled carbon nanotube: poly (methylmethacrylate) composite thin film ethanol sensors through appropriate nanotube functionalization. Mater. Sci. Semicond. Process. 2015, 31, 166–174. 10.1016/j.mssp.2014.11.030. [DOI] [Google Scholar]
  89. Tung T. T.; Castro M.; Kim T. Y.; Suh K. S.; Feller J.-F. High stability silver nanoparticles-graphene/poly (ionic liquid)-based chemoresistive sensors for volatile organic compounds’ detection. Anal. Bioanal. Chem. 2014, 406, 3995–4004. 10.1007/s00216-013-7557-y. [DOI] [PubMed] [Google Scholar]
  90. Elakia M.; Gobinath M.; Sivalingam Y.; Palani E.; Ghosh S.; Nutalapati V.; Surya V. J. Investigation on visible light assisted gas sensing ability of multiwalled carbon nanotubes coated with pyrene based organic molecules. Phys. E Low-dimens. Syst. Nanostruct. 2020, 124, 114232. 10.1016/j.physe.2020.114232. [DOI] [Google Scholar]
  91. Kour R.; Arya S.; Young S.-J.; Gupta V.; Bandhoria P.; Khosla A. Recent Advances in Carbon Nanomaterials as Electrochemical Biosensors. J. Electrochem. Soc. 2020, 167, 037555. 10.1149/1945-7111/ab6bc4. [DOI] [Google Scholar]
  92. Alwin David S.; Rajkumar R.; Karpagavinayagam P.; Fernando J.; Vedhi C.. Chapter 25 - Sustainable carbon nanomaterial-based sensors: Future vision for the next 20 years. Carbon Nanomaterials-Based Sensors; Jamballi G. M., Chaudhery Mustansar H.; Elsevier, 2022; pp 429–443, ISBN 9780323911740. [Google Scholar]
  93. Gakhar T.; Basu S.; Hazra A. Carbon-metal oxide nanocomposites for selective detection of toxic and hazardous volatile organic compounds (VOC) -A review. Green Anal Chem. 2022, 1, 100005. 10.1016/j.greeac.2022.100005. [DOI] [Google Scholar]
  94. Dariyal P.; Sharma S.; Chauhan G. S.; Singh B. P.; Dhakate S. R. Recent trends in gas sensing via carbon nanomaterials: outlook and challenges. Nanoscale Adv. 2021, 3, 6514–6544. 10.1039/D1NA00707F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Wongchoosuk C.; Wisitsoraat A.; Tuantranont A.; Kerdcharoen T. Portable electronic nose based on carbon nanotube-SnO2 gas sensors and its application for detec- tion of methanol contamination in whiskeys. Sensors Actuators, B Chem. 2010, 147, 392–399. 10.1016/j.snb.2010.03.072. [DOI] [Google Scholar]
  96. Liang Y. X.; Chen Y. J.; Wang J. H. Low-resistance gas sensors fabricated from mul- tiwalled carbon nanotubes coated with a thin tin oxide layer. Appl. Phys. Lett. 2004, 85, 666–668. 10.1063/1.1775879. [DOI] [Google Scholar]
  97. Mubeen S.; Lai M.; Zhang T.; Lim J. H.; Mulchandani A.; Deshusses M. A.; Myung N. V. Hybrid tin oxide-SWNT nanostructures based gas sensor. Electrochim. Acta 2013, 92, 484–490. 10.1016/j.electacta.2013.01.029. [DOI] [Google Scholar]
  98. Sachan A.; Castro M.; Choudhary V.; Feller J.-F. vQRS based on hybrids of CNT with PMMA-POSS and PS-POSS Copolymers to reach the sub-PPM detection of ammonia and formaldehyde at room temperature despite moisture. Chemosensors 2017, 5, 22. 10.3390/chemosensors5030022. [DOI] [Google Scholar]
  99. Song N.; Fan H.; Tian H. PVP assisted in situ synthesis of functionalized graphene/ZnO (FGZnO) nanohybrids with enhanced gas-sensing property. J. Mater. Sci. 2015, 50, 2229–2238. 10.1007/s10853-014-8785-z. [DOI] [Google Scholar]
  100. Singkammo S.; Wisitsoraat A.; Sriprachuabwong C.; Tuantranont A.; Phanich- phant S.; Liewhiran C. Electrolytically exfoliated graphene-loaded flame-made Ni-doped SnO2 composite film for acetone sensing. ACS Appl. Mater. Interfaces 2015, 7, 3077–3092. 10.1021/acsami.5b00161. [DOI] [PubMed] [Google Scholar]
  101. Liu L.; Zhang D.; Zhang Q.; Chen X.; Xu G.; Lu Y.; Liu Q. Smartphone-based sensing system using ZnO and graphene modified electrodes for VOCs detection. BiosensBioelectron. 2017, 93, 94–101. 10.1016/j.bios.2016.09.084. [DOI] [PubMed] [Google Scholar]
  102. Mu H.; Zhang Z.; Zhao X.; Liu F.; Wang K.; Xie H. High sensitive formaldehyde graphene gas sensor modified by atomic layer deposition zinc oxide films. Appl. Phys. Lett. 2014, 105, 033107. 10.1063/1.4890583. [DOI] [Google Scholar]
  103. Guo D.; Cai P.; Sun J.; He W.; Wu X.; Zhang T.; Wang X.; Zhang X. Reduced-graphene-oxide/metal-oxide p-n heterojunction aerogels as efficient 3D sensing frameworks for phenol detection. Carbon 2016, 99, 571–578. 10.1016/j.carbon.2015.12.074. [DOI] [Google Scholar]
  104. Guo L.; Kou X.; Ding M.; Wang C.; Dong L.; Zhang H.; Feng C.; Sun Y.; Gao Y.; Sun P.; Lu G. Reduced graphene oxide/A-Fe2O3 composite nanofibers for application in gas sensors. Sensors Actuators. B Chem. 2017, 244, 233–242. 10.1016/j.snb.2016.12.137. [DOI] [Google Scholar]
  105. Sun J.; Bai S.; Tian Y.; Zhao Y.; Han N.; Luo R.; Li D.; Chen A. Hybridization of ZnSnO3 and rGO for improvement of formaldehyde sensing properties. Sens. Actuator. B Chem. 2018, 257, 29–36. 10.1016/j.snb.2017.10.015. [DOI] [Google Scholar]
  106. Nag S.; Castro M.; Choudhary V.; Feller J.-F. Sulfonatedpoly(ether ether ketone) [SPEEK] nanocomposites based on hybrid nanocarbons for the detection and discrimination of some lung cancer VOC biomarkers. J. Mater. Chem. B 2017, 5, 348–359. 10.1039/C6TB02583H. [DOI] [PubMed] [Google Scholar]
  107. Huang Q.; Zeng D.; Li H.; Xie C. Room temperature formaldehyde sensors with enhanced performance, fast response and recovery based on zinc oxide quantum dots/graphenenanocomposites. Nanoscale 2012, 4, 5651–5658. 10.1039/c2nr31131c. [DOI] [PubMed] [Google Scholar]
  108. Shao S.; Kim H. W.; Kim S. S.; Chen Y.; Lai M. NGQDs modified nanoporous TiO2/graphene foam nanocomposite for excellent sensing response to formaldehyde at high relative humidity. Appl. Surf. Sci. 2020, 516, 145932. 10.1016/j.apsusc.2020.145932. [DOI] [Google Scholar]
  109. Yuan W.; Yang K.; Peng H.; Li F.; Yin F. A flexible VOCs sensor based on a 3D Mxene framework with a high sensing performance. J. Mater. Chem. 2018, 6, 18116–18124. 10.1039/C8TA06928J. [DOI] [Google Scholar]
  110. Kim S. J.; Koh H.-J.; Ren C. E.; Kwon O.; Maleski K.; Cho S.-Y.; Anasori B.; Kim C.-K.; Choi Y.-K.; Kim J.; Gogotsi Y.; Jung H.-T. Metallic Ti3C2Tx MXene Gas Sensors with Ultrahigh Signal-to- Noise Ratio. ACS Nano 2018, 12, 986–993. 10.1021/acsnano.7b07460. [DOI] [PubMed] [Google Scholar]
  111. Di Crescenzo A.; Ettorre V.; Fontana A. Non-covalent and reversible functionalization of carbon nanotubes. Beilstein J. Nanotechnol 2014, 5, 1675–1690. 10.3762/bjnano.5.178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Britz D. A.; Khlobystov A. N. Noncovalent interactions of molecules with single walled carbon nanotubes. ChemSoc. Rev. 2006, 35, 637–659. 10.1039/b507451g. [DOI] [PubMed] [Google Scholar]
  113. Iijima S.; Ichihashi T. Single-shell carbon nanotubes of 1-nm diameter. Nature 1993, 363, 603–605. 10.1038/363603a0. [DOI] [Google Scholar]
  114. Dai H. Nanotube growth and characterization. Carbon Nanotubes 2001, 80, 29–53. 10.1007/3-540-39947-X_3. [DOI] [Google Scholar]
  115. Dariyal P.; Sharma S.; Chauhan G. S.; Singh B. P.; Dhakate S. R. Recent trends in gas sensing via carbon nanomaterials: outlook and challenges. Nanoscale Adv. 2021, 3, 6514–6544. 10.1039/D1NA00707F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Cannac O.; Martinez-Almoyna L.; Hraiech S. Critical illness-associated cerebral microbleeds in COVID-19 acute respiratory distress syndrome. Neurology 2020, 95, 498. 10.1212/WNL.0000000000010537. [DOI] [PubMed] [Google Scholar]
  117. Zhang S.; Bick M.; Xiao X.; Chen G.; Nashalian A.; Chen J. Leveraging triboelectric nanogenerators for bioengineering. Matter 2021, 4, 845–887. 10.1016/j.matt.2021.01.006. [DOI] [Google Scholar]
  118. Tat T.; Libanori A.; Au C.; Yau A.; Chen J. Advances in triboelectric nanogenerators for biomedical sensing. Biosens. Bioelectron. 2021, 171, 112714. 10.1016/j.bios.2020.112714. [DOI] [PubMed] [Google Scholar]
  119. Khan Y.; Ostfeld A. E.; Lochner C. M.; Pierre A.; Arias A. C. Monitoring of Vital Signs with Flexible and Wearable Medical Devices. Adv. Mater. 2016, 28 (22), 4373–4395. 10.1002/adma.201504366. [DOI] [PubMed] [Google Scholar]
  120. Yu Y.; Ye L.; Song Y.; Guan Y.; Zang J. Wrinkled nitrile rubber films for stretchable and ultra-sensitive respiration sensors. Extreme Mech. Lett. 2017, 11, 128–136. 10.1016/j.eml.2016.12.003. [DOI] [Google Scholar]
  121. Su Y.; Chen G.; Chen C.; Gong Q.; Xie G.; Yao M.; Tai H.; Jiang Y.; Chen J. Self-Powered Respiration Monitoring Enabled By a Triboelectric Nanogenerator. Adv. Mater. 2021, 33, 2101262. 10.1002/adma.202101262. [DOI] [PubMed] [Google Scholar]
  122. Liu B.; Libanori A.; Zhou Y.; Xiao X.; Xie G.; Zhao X.; Su Y.; Wang S.; Yuan Z.; Duan Z.; Liang J.; Jiang Y.; Tai H.; Chen J. Simultaneous Biomechanical and Biochemical Monitoring for Self-Powered Breath Analysis. ACS Appl. Mater. Interfaces 2022, 14, 7301–7310. 10.1021/acsami.1c22457. [DOI] [PubMed] [Google Scholar]
  123. Wang S.; Jiang Y.; Tai H.; Liu B.; Duan Z.; Yuan Z.; Pan H.; Xie G.; Du X.; Su Y. An integrated flexible self-powered wearable respiration sensor. Nano Energy 2019, 63, 103829. 10.1016/j.nanoen.2019.06.025. [DOI] [Google Scholar]
  124. Su Y.; Li W.; Yuan L.; Chen C.; Pan H.; Xie G.; Conta G.; Ferrier S.; Zhao X.; Chen G.; Tai H.; Jiang Y.; Chen J. Piezoelectric fiber composites with polydopamine interfacial layer for self-powered wearable biomonitoring. Nano Energy 2021, 89, 106321. 10.1016/j.nanoen.2021.106321. [DOI] [Google Scholar]
  125. Su Y.; Wang J.; Wang B.; Yang T.; Yang B.; Xie G.; Zhou Y.; Zhang S.; Tai H.; Cai Z.; Chen G.; Jiang Y.; Chen L.-Q.; Chen J. Alveolus-Inspired Active Membrane Sensors for Self-Powered Wearable Chemical Sensing and Breath Analysis. ACS Nano 2020, 14 (5), 6067. 10.1021/acsnano.0c01804. [DOI] [PubMed] [Google Scholar]
  126. Panes-Ruiz L. A.; Riemenschneider L.; Al Chawa M. M.; Loffler M.; Rellinghaus B.; Tetzlaff R.; Bezugly V.; Ibarlucea B.; Cuniberti G. Selective and self-validating breath-level detection of hydrogen sulfide in humid air by gold nanoparticle-functionalized nanotube arrays. Nano Res. 2022, 15, 2512–2521. 10.1007/s12274-021-3771-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Chatterjee S.; Castro M.; Feller J. F. Tailoring selectivity of sprayed carbon nanotube sensors (CNT) towards volatile organic compounds (VOC) with surfactants. Sens Actuators B: Chem. 2015, 220, 840–849. 10.1016/j.snb.2015.06.005. [DOI] [Google Scholar]
  128. Septiani N. L. W.; Yuliarto B.; Nugraha; Dipojono H. K. Multiwalled carbon nanotubes-zinc oxide nanocomposites as low temperature toluene gas sensor. Appl. Phys. A: Mater. Sci. Process. 2017, 123, 166. 10.1007/s00339-017-0803-y. [DOI] [Google Scholar]
  129. Sinha M.; Neogi S.; Mahapatra R.; Krishnamurthy S.; Ghosh R. Material dependent and temperature driven adsorption switching (p- to n- type) using CNT/ZnO composite-based chemiresistive methanol gas sensor. Sens Actuators B: Chem. 2021, 336, 129729. 10.1016/j.snb.2021.129729. [DOI] [Google Scholar]
  130. Turner A.; McCoy T.; Cao W.; Karoui A.; Maswadeh W. M.; Vlahovic B.; Elsayed-Ali H. E.; Daniel B.; Castro M.; Sadasivuni K. K.; Elahi M.; Adedeji A.; Kumar B. Enhanced detection of volatile organic compounds (VOCs) by caffeine modified carbon nanotube junctions. Nano-Struct Nano-Objects 2020, 24, 100578. 10.1016/j.nanoso.2020.100578. [DOI] [Google Scholar]
  131. Bohli N.; Belkilani M.; Casanova-Chafer J.; Llobet E.; Abdelghani A. Multiwalled carbon nanotube based aromatic volatile organic compound sensor: sensitivity enhancement through1-hexadecanethiol functionalisation. Beilstein J. Nanotechnol. 2019, 10, 2364–2373. 10.3762/bjnano.10.227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Yang M.; Au C.; Deng G.; Mathur S.; Huang Q.; Luo X.; Xie G.; Tai H.; Jiang Y.; Chen C.; Cui Z.; Liu X.; He C.; Su Y.; Chen J. NiWO4 Microflowers on Multi-Walled Carbon Nanotubes for High- Performance NH3 Detection. ACS Appl. Mater. Interfaces 2021, 13, 52850–52860. 10.1021/acsami.1c10805. [DOI] [PubMed] [Google Scholar]
  133. Novoselov K. S.; Geim A. K.; Morozov S. V.; Jiang D.; Zhang Y.; Dubonos S. V.; Grigorieva I. V.; Firsov A. A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. 10.1126/science.1102896. [DOI] [PubMed] [Google Scholar]
  134. Jacoby M. Nobel prize in physics. Chem. Eng. News Archive 2010, 88, 8. 10.1021/cen-v088n041.p008. [DOI] [Google Scholar]
  135. Geim A. K. Graphene: status and prospects. Science. 2009, 324, 1530–1534. 10.1126/science.1158877. [DOI] [PubMed] [Google Scholar]
  136. Hong S.; Myung S. Nanotube electronics—a flexible approach to mobility. Nat. Nanotechnol 2007, 2, 207–208. 10.1038/nnano.2007.89. [DOI] [PubMed] [Google Scholar]
  137. Chen J.; Bi H.; Sun S. R.; Tang Y. F.; Zhao W.; Lin T. Q.; Wan D. Y.; Huang F. Q.; Zhou X. D.; Xie X. M.; Jiang M. H. Highly conductive and flexible paper of 1D silver-nanowire-doped graphene. ACS Appl. Mater. Interfaces 2013, 5, 1408–1413. 10.1021/am302825w. [DOI] [PubMed] [Google Scholar]
  138. Layek R. K.; Kuila A.; Chatterjee D. P.; Nandi A. K. Amphiphilicpoly(N-vinyl pyrrolidone) grafted graphene by reversible addition and fragmentation polymerization and the reinforcement of poly(-vinyl acetate) films. J. Mater. ChemA 2013, 1, 10863–10874. 10.1039/c3ta11853c. [DOI] [Google Scholar]
  139. 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. [DOI] [Google Scholar]
  140. Hwang J.; Kim D. S.; Ahn D.; Hwang S. W. Transport properties of a DNA-conjugated single-wall carbon nanotube field-effect transistor. Jpn. J. Appl. Phys. 2009, 48, 06FD08. 10.1143/JJAP.48.06FD08. [DOI] [Google Scholar]
  141. Sridhar V.; Kim H. J.; Jung J. H.; Lee C.; Park S.; Oh I. K. Defect-engineered three-dimensional graphene-nanotubepalladium nanostructures with ultrahigh capacitance. ACS Nano 2012, 6, 10562–10570. 10.1021/nn3046133. [DOI] [PubMed] [Google Scholar]
  142. Chen S.; Wu Q. Z.; Mishra C.; Kang J.; Zhang H.; Cho K.; Cai W.; Balandin A. A.; Ruoff R. S. Thermal conductivity of isotopically modified graphene. Nat. Mater. 2012, 11, 203–207. 10.1038/nmat3207. [DOI] [PubMed] [Google Scholar]
  143. Ishigami M.; Chen J. H.; Cullen W. G.; Fuhrer M. S.; Williams E. D. Atomic structure of graphene on SiO2. Nano Lett. 2007, 7, 1643–1648. 10.1021/nl070613a. [DOI] [PubMed] [Google Scholar]
  144. Tung T. T.; Nine M.d. J.; Krebsz M.; Pasinszki T.; Coghlan C. J.; Tran D. N. H.; Losic D. Recent Advances in Sensing Applications of Graphene Assemblies and Their Composites. AdvFunct Mater. 2017, 27, 1702891. 10.1002/adfm.201702891. [DOI] [Google Scholar]
  145. Tombel N. S. M.; Badaruddin S. A. M.; Yakin F. S. M.; Zaki H. F. M.; Syono M. I. Detection of low PPM of volatile organic compounds using nanomaterial functionalized reduced graphene oxide sensor. AIP Conf. Proc. 2021, 2368, 020004. [Google Scholar]
  146. Lee B. M.; Eetemadi A.; Tagkopoulos I. Reduced Graphene Oxide-Metalloporphyrin Sensors for Human Breath Screening. ApplSci. 2021, 11, 11290. 10.3390/app112311290. [DOI] [Google Scholar]
  147. Wang T.; Sun Z.; Huang D.; Yang Z.; Ji Q.; Hu N.; Yin G.; He D.; Wei H.; Zhang Y. Studies on NH3 gas sensing by zinc oxide nanowire-reduced graphene oxide nanocomposites. Sens. Actuators B Chem. 2017, 252, 284–294. 10.1016/j.snb.2017.05.162. [DOI] [Google Scholar]
  148. Gupta M.; Athirah N.; Fahmi Hawari H. Graphene derivative coated QCM-based gas sensor for volatile organic compound (VOC) detection at room temperature. IJEECS 2020, 18, 1279–1286. 10.11591/ijeecs.v18.i3.pp1279-1286. [DOI] [Google Scholar]
  149. Pargoletti E.; Hossain U. H.; Di Bernardo I.; Chen H.; Tran-Phu T.; Lipton-Duffin J.; Cappelletti G.; Tricoli A. Room-temperature photodetectors and VOC sensors based on graphene oxide-ZnO nanoheterojunctions. Nanoscale 2019, 11, 22932–22945. 10.1039/C9NR08901B. [DOI] [PubMed] [Google Scholar]
  150. Parvizi R.; Azad S.; Dashtian K.; Ghaedi M.; Heidari H. Natural Source-Based Graphene as Sensitising Agents for Air Quality Monitoring. Sci. Rep. 2019, 9, 3798. 10.1038/s41598-019-40433-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Gakhar T.; Hazra A. p-TiO2/G.O. heterojunction based VOC sensors: A new approach to amplify sensitivity in FET structure at optimized gate voltage. Measurement 2021, 182, 109721. 10.1016/j.measurement.2021.109721. [DOI] [Google Scholar]
  152. Yin F.; Li Y.; Yue W.; Gao S.; Zhang C.; Chen Z. Sn3O4/rGOheterostructure as a material for formaldehyde gas sensor with a wide detecting range and low operating temperature. Sens. Actuators B Chem. 2020, 312, 127954. 10.1016/j.snb.2020.127954. [DOI] [Google Scholar]
  153. Agrawal A. V.; Kumar N.; Kumar M. Strategy and future prospects to develop room-temperature-recoverable NO2 gas sensor based on two-dimensional molybdenum disulfide. Nano-Micro Lett. 2021, 13, 1–58. 10.1007/s40820-020-00558-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Zhu Z. Y.; Liu C. C.; Jiang F. X.; Liu J.; Liu G. Q.; Ma X. M.; et al. Flexible fiber-shaped hydrogen gas sensor via coupling palladium with conductive polymer gel fiber. J. Hazard. Mater. 2021, 411, 125008. 10.1016/j.jhazmat.2020.125008. [DOI] [PubMed] [Google Scholar]
  155. Chen X.; Wang T.; Han Y.; Lv W.; Li B.; Su C.; Zeng M.; Yang J.; Hu N.; Su Y.; Yang Z. Wearable NO2 sensing and wireless application based on ZnS nanoparticles/nitrogen-doped reduced graphene oxide. Sens. Actuators B Chem. 2021, 345, 130423. 10.1016/j.snb.2021.130423. [DOI] [Google Scholar]
  156. Dolai S.; Bhunia S. K.; Rajendran S.; Vipinachandran V. U.; Ray S. C.; Kluson P. Tunable fluorescent carbon dots: synthesis progress, fluorescence origin, selective and sensitive volatile organic compounds detection. Crit Rev. Solid State Mater. Sci. 2021, 46, 349–370. 10.1080/10408436.2020.1830750. [DOI] [Google Scholar]
  157. Mishra R. K.; Choi G.-J.; Choi H.-J.; Gwag J. S. ZnS Quantum Dot Based Acetone Sensor for Monitoring Health-Hazardous Gases in Indoor/Outdoor Environment. Micromachines (Basel). 2021, 12, 598–609. 10.3390/mi12060598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Wang X.; Xie Z.; Huang H.; Liu Z.; Chen D.; Shen G. Gas sensors, thermistor and photodetector based on ZnS nanowires. J. Mater. Chem. 2012, 22, 6845–6850. 10.1039/c2jm16523f. [DOI] [Google Scholar]
  159. Zhang L.; Dong R.; Zhu Z.; Wang S. Au nanoparticles decorated ZnS hollow spheres for highly improved gas sensor performances. Sens. Actuators B Chem. 2017, 245, 112–121. 10.1016/j.snb.2017.01.179. [DOI] [Google Scholar]
  160. Zhang W.; Wang S.; Wang Y.; Zhu Z.; Gao X.; Yang J.; Zhang H. X. ZnO@ZnS core/shell microrods with enhanced gas sensing properties. RSC Adv. 2015, 5, 2620–2629. 10.1039/C4RA12803F. [DOI] [Google Scholar]
  161. Sun G.-J.; Kheel H.; Ko T.-G.; Lee C.; Kim H. M. Prominent Ethanol Sensing with Cr2O3 Nanoparticle-Decorated ZnSNanorods Sensors. J. Korean Phys. Society 2016, 69, 390–397. 10.3938/jkps.69.390. [DOI] [Google Scholar]
  162. Choi H. J.; Chung J.-H.; Yoon J.-W.; Lee J.-H. Highly selective, sensitive, and rapidly responding acetone sensor using ferroelectric ε-WO3 spheres doped with Nb for monitoring ketogenic diet efficiency. Sens. Actuators B Chem. 2021, 338, 129823. 10.1016/j.snb.2021.129823. [DOI] [Google Scholar]
  163. Wang C.; Wang Y.; Cheng P.; Xu L.; Dang F.; Wang T.; Lei Z. In-situ generated TiO2/α-Fe2O3 heterojunction arrays for batch manufacturing of conductometric acetone gas sensors. Sens. Actuators B Chem. 2021, 340, 129926. 10.1016/j.snb.2021.129926. [DOI] [Google Scholar]
  164. Zhou C.; Meng F.; Chen K.; Yang X.; Wang T.; Sun P.; Liu F.; Yan X.; Shimanoe K.; Lu G. High sensitivity and low detection limit of acetone sensor based on NiO/ Zn2SnO4 p-n heterojunction octahedrons. Sens Actuators B Chem. 2021, 339, 129912. 10.1016/j.snb.2021.129912. [DOI] [Google Scholar]
  165. Cheng P.; Lv L.; Wang Y.; Zhang B.; Zhang Y.; Zhang Y.; Lei Z.; Xu L. SnO2/ZnSnO3 double-shelled hollow microspheres based high-performance acetone gas sensor. Sens. Actuators B Chem. 2021, 332, 129212. 10.1016/j.snb.2020.129212. [DOI] [Google Scholar]
  166. Wang J.; Lin S.; Lin Y.; Wang X. Preparation of Graphene Quantum Dots and Their Sensing Properties in Quartz Crystal Microbalance Acetone Sensor. Sens.Mater. 2021, 33, 499–511. 10.18494/SAM.2021.3075. [DOI] [Google Scholar]
  167. Junkaew A.; Arroyave R. Enhancement of the selectivity of MXenes (M2C, M = Ti, V, Nb, Mo) via oxygen-functionalization: promising materials for gas-sensing and -separation. Phys. Chem. Chem. Phys. 2018, 20, 6073–6080. 10.1039/C7CP08622A. [DOI] [PubMed] [Google Scholar]
  168. Yu X.-F.; Li Y. C.; Cheng J.-B.; Liu Z.-B.; Li Q.-Z.; Li W.-Z.; Yang X.; Xiao B. Monolayer Ti2CO2: A Promising Candidate for NH3 Sensor or Capturer with High Sensitivity and Selectivity. ACS Appl. Mater. 2015, 7, 13707–13713. 10.1021/acsami.5b03737. [DOI] [PubMed] [Google Scholar]
  169. Ghidiu M.; Lukatskaya M. R.; Zhao M.-Q.; Gogotsi Y.; Barsoum M. W. Conductive two-dimensional titanium carbide ’clay’ with high volumetric capacitance. Nature 2014, 516, 78–81. 10.1038/nature13970. [DOI] [PubMed] [Google Scholar]
  170. Lukatskaya M. R.; Mashtalir O.; Ren C. E.; Dall’Agnese Y.; Rozier P.; Taberna P. L.; Naguib M.; Simon P.; Barsoum M. W.; Gogotsi Y. Cation Intercalation and High Volumetric Capacitance of Two-Dimensional Titanium Carbide. Science 2013, 341, 1502–1505. 10.1126/science.1241488. [DOI] [PubMed] [Google Scholar]
  171. Naguib M.; Mashtalir O.; Carle J.; Presser V.; Lu J.; Hultman L.; Gogotsi Y.; Barsoum M. W. Two-Dimensional Transition Metal Carbides. ACS Nano 2012, 6, 1322–1331. 10.1021/nn204153h. [DOI] [PubMed] [Google Scholar]
  172. Anasori B.; Xie Y.; Beidaghi M.; Lu J.; Hosler B. C.; Hultman L.; Kent P. R. C.; Gogotsi Y.; Barsoum M. W. Two-Dimensional, Ordered, Double Transition Metals Carbides (MXenes). ACS Nano 2015, 9, 9507–9516. 10.1021/acsnano.5b03591. [DOI] [PubMed] [Google Scholar]
  173. Chen W. Y.; Jiang X.; Lai S.-N.; Peroulis D.; Stanciu L. Nanohybrids of a MXene and transition metal dichalcogenide for selective detection of volatile organic compounds. Nat. Commun. 2020, 11, 1302. 10.1038/s41467-020-15092-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Guo W.; Surya S. G.; Babar V.; Ming F.; Sharma S.; Alshareef H. N.; Schwingenschlögl U.; Salama K. N. Selective Toluene Detection with Mo2CTx MXene at Room Temperature. ACS Appl. Mater. Interfaces 2020, 12, 57218–57227. 10.1021/acsami.0c16302. [DOI] [PubMed] [Google Scholar]
  175. Lee E.; VahidMohammadi A.; Prorok B. C.; Yoon Y. S.; Beidaghi M.; Kim D.-J. Room Temperature Gas Sensing of Two- Dimensional Titanium Carbide (MXene). ACS Appl. Mater. Interfaces 2017, 9, 37184–37190. 10.1021/acsami.7b11055. [DOI] [PubMed] [Google Scholar]
  176. Kim S. J.; Koh H.-J.; Ren C. E.; Kwon O.; Maleski K.; Cho S.-Y.; Anasori B.; Kim C.-K.; Choi Y.-K.; Kim J.; Gogotsi Y.; Jung H.-T. Metallic Ti3C2Tx MXene gas sensors with ultrahigh signal-to-noise ratio. ACS Nano 2018, 12, 986–993. 10.1021/acsnano.7b07460. [DOI] [PubMed] [Google Scholar]
  177. Lee E.; VahidMohammadi A.; Yoon Y. S.; Beidaghi M.; Kim D.-J. Two-Dimensional Vanadium Carbide MXene for Gas Sensors with Ultrahigh Sensitivity Toward Nonpolar Gases. ACS Sens. 2019, 4, 1603–1611. 10.1021/acssensors.9b00303. [DOI] [PubMed] [Google Scholar]
  178. Hermawan A.; Zhang B.; Taufik A.; Asakura Y.; Hasegawa T.; Zhu J.; Shi P.; Yin S. CuO Nanoparticles/Ti3C2Tx MXene Hybrid Nanocomposites for Detection of Toluene Gas. ACS Appl. Nano Mater. 2020, 3, 4755–4766. 10.1021/acsanm.0c00749. [DOI] [Google Scholar]
  179. Shuvo S. N.; Ulloa Gomez A. M.; Mishra A.; Chen W. Y.; Dongare A. M.; Stanciu L. A. Sulfur-Doped Titanium Carbide MXenes for Room Temperature Gas Sensing. ACS Sens. 2020, 5, 2915. 10.1021/acssensors.0c01287. [DOI] [PubMed] [Google Scholar]
  180. Zhao W.-N.; Yun N.; Dai Z.-H.; Li Y.-F. A high-performance trace level acetone sensor using an indispensable V4C3TxMXene. RSC Adv. 2020, 10, 1261–1270. 10.1039/C9RA09069J. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Sun S.; Wang M.; Chang X.; Jiang Y.; Zhang D.; Wang D.; Zhang Y.; Lei Y. W18O49/Ti3C2Tx Mxenenanocomposites for highly sensitive acetone gas sensor with low detection limit. Sens. Actuators B Chem. 2020, 304, 127274. 10.1016/j.snb.2019.127274. [DOI] [Google Scholar]
  182. Fan C.; Shi J.; Zhang Y.; Quan W.; Chen X.; Jianhua Yang J.; Zeng M.; Zhou Z.; Su Y.; Wei H.; Yang Z. Fast and recoverable NO2 detection achieved by assembling ZnO on Ti3C2Tx MXene nanosheets under U.V. illumination at room temperature. Nanoscale 2022, 14, 3441. 10.1039/D1NR06838E. [DOI] [PubMed] [Google Scholar]
  183. Homaeigohar S. Amphiphilic Oxygenated Amorphous Carbon-Graphite Buckypapers with Gas Sensitivity to Polar and Non-Polar VOCs. Nanomaterials 2019, 9, 1343. 10.3390/nano9091343. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES