Next Article in Journal
High-Performance Zwitterionic Organohydrogel Fiber in Bioelectronics for Monitoring Bioinformation
Previous Article in Journal
From Piezoelectric Nanogenerator to Non-Invasive Medical Sensor: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Sensing Materials Targeting Clinical Volatile Organic Compound (VOC) Biomarkers: A Review

by
Akhilesh Kumar Pathak
1,*,
Kankan Swargiary
1,
Nuntaporn Kongsawang
2,
Pannathorn Jitpratak
2,
Noppasin Ajchareeyasoontorn
1,
Jade Udomkittivorakul
1 and
Charusluk Viphavakit
1,*
1
International School of Engineering (ISE), Intelligent Control Automation of Process Systems Research Unit, Chulalongkorn University, Bangkok 10330, Thailand
2
Biomedical Engineering Program, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand
*
Authors to whom correspondence should be addressed.
Biosensors 2023, 13(1), 114; https://doi.org/10.3390/bios13010114
Submission received: 29 November 2022 / Revised: 22 December 2022 / Accepted: 5 January 2023 / Published: 9 January 2023
(This article belongs to the Section Biosensor and Bioelectronic Devices)

Abstract

:
In general, volatile organic compounds (VOCs) have a high vapor pressure at room temperature (RT). It has been reported that all humans generate unique VOC profiles in their exhaled breath which can be utilized as biomarkers to diagnose disease conditions. The VOCs available in exhaled human breath are the products of metabolic activity in the body and, therefore, any changes in its control level can be utilized to diagnose specific diseases. More than 1000 VOCs have been identified in exhaled human breath along with the respiratory droplets which provide rich information on overall health conditions. This provides great potential as a biomarker for a disease that can be sampled non-invasively from exhaled breath with breath biopsy. However, it is still a great challenge to develop a quick responsive, highly selective, and sensitive VOC-sensing system. The VOC sensors are usually coated with various sensing materials to achieve target-specific detection and real-time monitoring of the VOC molecules in the exhaled breath. These VOC-sensing materials have been the subject of huge interest and extensive research has been done in developing various sensing tools based on electrochemical, chemoresistive, and optical methods. The target-sensitive material with excellent sensing performance and capturing of the VOC molecules can be achieved by optimizing the materials, methods, and its thickness. This review paper extensively provides a detailed literature survey on various non-biological VOC-sensing materials including metal oxides, polymers, composites, and other novel materials. Furthermore, this review provides the associated limitations of each material and a summary table comparing the performance of various sensing materials to give a better insight to the readers.

1. Introduction

Nowadays, exhaled breath analysis is considered an important area of research in order to diagnose several diseases at an early stage. In ancient times, several physicians observed certain odors in the human exhaled breath which are associated with a certain specific disease and its stage. For example, a diabetic patient was observed to exhale a ‘fruity’ odor while renal failure has shown an association with some ‘fishy’ smell [1]. Later, in modern science, these exhaled odors were recognized as volatile organic compounds (VOCs) [2,3]. These VOCs are produced in the human body either from exogenous VOCs or endogenous VOCs. The exogenous VOCs contain the compounds inhaled from the outside environment, which are produced following the oral ingestion of food and compounds obtained from smoking cigarettes. On the other hand, endogenous VOCs are compounds made from symbiotic bacteria and/or blood-borne compounds which can be released into the environment via the lungs. Therefore, considering the potential of VOCs being used as biomarkers, their real-time monitoring and analysis are essential and recognized as a new frontier in health inspections and disease diagnostics [4,5].
The concentrations of available VOCs in the exhaled breath are determined in the range of nanomolar to picomolar levels. Hence, it is always a challenging task to distinguish normal individuals from the patients suffering from some chronic disease via endogenously formed VOCs. However, collecting the exhaled breath samples is painless, extremely simple, and noninvasive which provides a great benefit to the patients. Therefore, it is worth investigating human breath VOC detection. Gas chromatography–mass spectrometry (GC–MS) has been considered as the most accurate, selective, and sensitive method which has been utilized for several years to detect these VOCs from exhaled breath. The GC–MS technique utilizes the ion-pair extraction of the breath sample and quantification by mass spectrometry to monitor VOC concentrations as low as part per trillion (ppt). Although the instrument is promising, it faces the most significant limitations due to its cost, time consumption, off-site analyses, and the requirement for skilled technicians. To overcome this limitation, several researchers have proposed various user-friendly interfaces for the sensing system. To detect these VOC biomarkers in the exhaled human breath such as exploiting catalyst [6], electrochemical [7], resistance [8], and optical sensors [9], coating the novel-sensing materials on these devices is the common practice to enhance and improve the sensing performance.
Nanoscale sensing materials of sizes varying in the range of 2 nm to 100 nm (within the Debye length of the surface) are utilized as the highly sensitive sensing elements in vapor/gas sensors due to their large surface-to-volume-ratio and their unique optical, chemical, and electrical properties compared to the non-nanoscale sensing materials [10]. These vapor/gas-sensitive materials usually include metal oxide, carbon-based materials, and polymers with various morphological structures, integrated over conventional transducers (e.g., optical, capacitive, resistive, and gravimetric). The sensing properties of these materials depend on the chemisorption of negatively charged oxygen adsorbents in the air [11]. Gas chemisorption is an energy-activated process and the oxygen species require different temperatures to take place on the surface of grains. In air, the oxygen molecules capture electrons from the materials and then exist as O2 (<100 °C), O (100 °C to 300 °C), and O2− (>300 °C) on the surface of sensing materials [11,12]. Generally, the relatively high temperature may endow adsorbed VOC gaseous molecules with high activation and could speed up the rate of gas diffusion. It is conducive to the occurrence of chemical reactions between the oxygen species and the target VOC molecules [13,14]. It also depends on the chemical reaction between gas/vapor and sensing material, or the diffusion of VOC into the bulk of the sensing material. Therefore, an intensive study to employ these sensing materials for vapor/gas sensing systems is required. In this review, the non-biological materials are mainly considered and discussed for the VOC sensors. With these materials, a sensor can achieve rapid response, simple synthesis, and good thermal and chemical stabilities. In addition, the associated disadvantages, and challenges of these materials in terms of the sensitivity, response time, detection limit, selectivity, and their corresponding sensing mechanisms are also presented and discussed. For instance, a summary of VOC’s sensing materials is shown in Figure 1.
This review is organized in the following manner: Section 2 contains information on the main sources of VOCs in the human body, while Section 3 provides a brief discussion on major VOCs present in the exhaled breath and the associated diseases/disorders. In Section 4, we discussed various classical approaches, that have been utilized for several years to detect VOCs in human breath. In Section 5, Section 6, Section 7 and Section 8 we comprehensively provide a literature review on various sensing materials including metal oxide (zinc oxide, molybdenum trioxide, titanium dioxide), carbon-based materials (graphene, multi-wall carbon nanotubes, activated carbon), polymers (polydimethylsiloxane, polypyrrole, polythiophene), and other materials (MIL-n series, Irmofs series, Zeolites), respectively. Section 9 and Section 10 include the future prospects and clinical challenges along with a comparative analysis between the gold standard and the other material-based devices, respectively. Finally, an overall conclusion is drawn in Section 11.

2. Origins of VOCs in the Human Body

The VOC composition in the exhaled breath is complex and determined by several external and internal factors, e.g., metabolic, inflammatory, and oxidative stress processes and environmental conditions [15]. These VOCs originate from several sources, including the respiratory tract, tissues, organs, and microbiomes. The source of VOCs can be categorized into two types, (i) exogenous and (ii) endogenous. The exogenous VOCs are defined as the VOCs resulting from external exposure. Exogenous VOCs comprise a wide range of sources and have largely been overlooked as environmental contaminants in biological samples. Nowadays, exogenous VOCs gain huge interest as most of them interact with biological systems and provide valuable information relevant to health and diseases. Whereas, endogenous VOCs are produced throughout the body and are selected and distributed in the bloodstream. From the blood, these VOCs are exchanged into air in the lungs and are then exhaled in respiratory droplets and atmospheric gases. Exhaled VOCs provide a source of useful biomarkers with clear associations with the body’s metabolism. Breath Biopsy enables non-invasive collection and analysis of the VOC biomarkers from breath. Usually, it takes one minute for blood to flow around the entire circulatory system. By sampling the breath for a minute or longer, even at low levels of the VOC biomarkers from all parts of the body, it can be pre-concentrated, detected, and identified. It is easy to increase the sensitivity to aid in detecting subtle changes during the very early stages of the disease by simply extending the sample collection time [16]. In the next section, we provide detailed information on various VOCs and the associated disorders.

3. Diseases and Disorders Indicated by Important Biomarkers

The most important VOC biomarkers of the diseases in the human body are isoprene (C5H8), acetone (C3H6O), ammonia (NH3), methane (CH4), aldehyde (R−CH=O), and hydrogen sulfide (H2S) [17]. This section will briefly summarize the metabolic processes that extract these VOC biomarkers and outline the disorders which are indicated by the amount that is present in the exhaled breath. The detailed information on these VOCs is listed below [18]:
(i). Isoprene: Isoprene generally appears copiously in the exhaled breath of patients suffering from some chronic diseases. In 2001, McGrath et al. investigated the level of isoprene in exhaled breath of patients suffering from heart failure [19]. Breath isoprene production in subjects with chronic heart failure was significantly reduced in the patients from 83 pmol/min·kg compared to the controls which are 168 pmol/ min·kg. Later in 2009, Amal et al. diagnosed the stage of lung cancer by investigating the concentration of isoprene in the exhaled breath [20]. According to the report, the person affected with lung cancer exhaled a lower concentration of isoprene, acetone, and methanol compared to the healthy control. In 2019, Phillip et al. performed a historical study of 53 pediatric Type-I diabetes mellitus (T1DM) patients in parallel to the children suffering from chronic kidney disease and healthy control [21]. The outcome suggested that the T1DM patients exhaled significantly higher concentrations of ethanol, isoprene, and pentanal compared to the control one. Additionally, the alveolar concentrations of isopropanol and pentanal were significantly different in patients with adequate and poor long-term metabolic control. From the report, it was observed that isoprene is linked to the biosynthesis of cholesterol and is most probably stored in fat tissue.
(ii). Acetone: Acetone is a metabolite of glucose and human fat. It appears as the most common VOC in exhaled human breath. It was, firstly, recognized as the diabetes biomarker by Petters et al. in 1857 [22]. Glucose has been popularly known to be a primary source of energy in the human body. The insulin, which is a hormone, generated by the pancreas enables cells to absorb glucose molecules. When the pancreas cannot generate insulin, it is considered T1DM. On the other hand, if the pancreas can generate insulin but fails to be used by the body, it is known as Type-II diabetes mellitus (T2DM). Usually, the body can extract energy from glucose by breaking body fat to produce energy [18]. This process is called ketogenesis. Ketogenesis is the source of all ketone bodies including acetone in humans. All ketone substances, including acetone, are produced during ketogenesis in the human body. The concentration of acetone in the exhaled human breath is directly associated with the concentration level of blood sugar. The acetone level in exhaled breath and blood sugar has a strong linear correlation, which is utilized to diagnose the stage of diabetes [23]. The higher concentration of the exhaled breath acetone indicates the severity of diabetes and hence can be an alarming situation. For a non-diabetic person, the exhaled breath acetone level is ≤0.9 parts per million (ppm). For the moderately diabetic person, the acetone level is 0.9 ppm to 1.8 ppm, and for the severe-stage diabetic people this value can be several tens of ppm.
(iii). Ammonia: Ammonia plays a vital role in nutritional benefits including the synthesis of amino sugar, the synthesis of pyrimidines and purines, producing non-essential amino acids, and maintaining the acid-base balance in the human body. However, when the ammonia concentration exceeds its control level, it becomes toxic to the human body. Therefore, this excess amount of ammonia has to be released from the body by ornithine or the urea cycle which converts this excess ammonia into urea and excretes it in the form of urine via the kidney [24]. This conversion cycle takes place in the kidney and liver. Therefore, any problem associated with kidney and liver functioning will lead to an excess amount of ammonia concentration in the exhaled breath. Hence, the high concentration of ammonia in the exhaled breath indicates several diseases such as liver dysfunction, kidney failure, hepatic encephalopathy, type-II Alzheimer, swelling of the brain, etc. [25,26].
(iv). Methane: Methane is considered to be an important gas appearing in exhaled human breath. It is produced by the methane-producing bacteria present in the intestine under anaerobic circumstances [27]. Methane is not present in the breath under normal conditions, but when it is when excessively created. The variation in methane levels can also be observed under other conditions including inflammatory bowel disease, obesity, irritable bowel syndrome, etc. [28]. Methane is produced and can be considered a biomarker for oxidative stress, heart disease, breast cancer, hepatic disease. Therefore, hydrocarbons, including methane, pentane, and ethane, can serve as potential biomarkers for several diseases that involve oxidative stress such as neck and head cancer [29].
(v). Aldehydes: Aldehydes, such as hydroxy alkenals, alkenals, and dialdehydes, are the products of lipid peroxidation and alcohol oxidation. Their elevated levels were observed in cancer patients [30]. During lipid peroxidation, polyunsaturated fatty acids interact with free radicals and produce aldehydes. Patients suffering from oxidative stress, liver cancer/disease, and Wilson’s disease have been observed to show a higher concentration of aldehydes in their exhaled breath and/or blood [31]. In several metabolic or genetic disorders such as aging, diabetes, Parkinson’s and Alzheimer’s disease, elevated levels of aldehydes (especially formaldehyde, glyoxal) in blood, urine, and breath have been reported [32].
(vi). Hydrogen sulfide: Hydrogen sulfide is a popular VOC with a malodor. It is a significant gasotransmitter in animals and humans signaling several physical processes including neuromodulation, inflammation, cytoprotection, apoptosis, and vascular tone regulation [33,34]. Hydrogen sulfide in the exhaled breath is considered to be the potential biomarker for airway inflammation, asthma, and also dental and oral health [35,36,37].
In Table 1 we summarize all these VOCs along with their molecular formula and associated disorders.

4. Techniques of Detecting VOC Biomarkers in the Exhaled Breath

The aforementioned discussions show that the human breath contains a variety of VOCs that serve as biomarkers for various diseases and metabolic problems. Therefore, the real-time monitoring of such VOCs in the exhaled human breath is highly essential to enable non-invasive illness detection. In the following subsections, a detailed discussion has been carried out to understand various techniques developed to detect VOCs in very low concentrations of part per million volumes (ppmv), part per billion volumes (ppbv), and part per trillion volumes (pptv).

4.1. Gas Chromatography–Mass Spectrometry (GC–MS) Techniques

GC–MS is a technique in which a mixture of molecules of various compounds travel by a carrier gas (normally helium) via a column that separates molecules and detects them by a detector. In the past decades, this technique has been utilized widely along with various types of detectors to detect VOCs in exhaled breath for the purpose of health monitoring. A block schematic diagram of GC–MS is shown in Figure 2. In 2003, Sanchez et al. detected 25 VOCs in human exhaled breath using a series couple column which includes some of the important biomarkers such as ethanol, methanol, acetone, isoprene, pentane, etc. [54]. The limit of detection (LOD) was observed to be 1–5 ppb in 0.8 L of the exhaled breath. Lord et al. also developed a GC–MS-based analytical detection system which could selectively detect acetone and ethanol in the exhaled breath and reduce the moisture effect to a large extent [55]. Later, Giardina et al. also proposed a low-temperature glassy carbon-based solid-phase mass extraction microfiber which was capable of extracting at least five types of cancer related to VOC biomarkers from the simulated breath. The extracted sample was analyzed using GC–MS with good sensitivity [56]. Lamote et al. utilized e-Nose and GC–MS to distinguish between malignant pleural mesothelioma patients and asymptomatic asbestos-exposed people at a risk of the mentioned disease. Schnabel et al. utilized GC–Time of flight–mass spectrometry (GC–TOF–MS) to demonstrate the non-invasive monitoring of ventilator associated pneumonia in ICU patients by exhaled breath analysis [57]. They identified nearly 12 VOC biomarkers for this purpose. GC–MS was also utilized by Acevedo et al. for gastric cancer [58]. The study proposed the approach to distinguish between healthy people and patients suffering from gastric cancer by the exhaled breath analysis. From all this literature, we can say that GC–MS is considered a potential technique for quantitative analysis for non-invasive detection.

4.2. Selected-Ion Flow-Tube Mass Spectrometry (SIFT-MS)

SIFT-MS is a tool of analytical chemistry which is similar to gas chromatography for the quantitative monitoring of VOCs [59]. In this technique, the VOC samples are ionized by the reagent ions, such as NO+, H3O+, O2+, etc., which can be later analyzed by a quadrupole mass spectrometer. Figure 3 shows the block diagram of SIFT-MS. SIFT-MS was first reported to (i) determine the VOC trace presented in the exhaled human breath for the prognosis of the disease and (ii) understand pathophysiological and physiological conditions. In 1996, Smith et al. used SIFT-MS to detect ammonia from the exhaled breath of a known Helicobacter pyroli-infected person which was observed to be increased by ∼4 ppm after an oral dose of 2 g nonradioactive urea [60]. In 1999, Spanel et al. utilized SIFT-MS and O2+ as reagent ions in order to quantitatively detect isoprene in the exhaled breath [61]. Later in 2002, Diskin et al. investigated the variation in concentration of common breath VOC biomarkers, such as ammonia, isoprene, ethanol, acetaldehyde, and acetone over a period of 30 days using SIFT-MS with healthy individuals [62]. In the same year, Abbott et al. utilized SIFT-MS to detect acetonitrile in the breath and urinary samples of several smokers and nonsmokers [63]. The result exhibited that the acetonitrile concentration in the exhaled breath was achieved within the range of 17 ppb to 124 ppb, while the urinary acetonitrile concentrations were obtained within the range of 0 μg/L to 150 μg/L that were close to the concentrations previously determined in the blood. Other researchers also utilized SIFT-MS to determine the concentration of various VOCs and diagnosed the stage of related disorders [64,65,66].

4.3. Proton-Transfer-Reaction Mass Spectrometry (PTR-MS)

Classical PTR-MS is a tool of analytical chemistry which utilizes gas-phase hydronium ion as n ion with purity >99.5% of source reagent. The PTR-MS is utilized to monitor the absolute concentration of the selected VOCs as low as pptv without any calibration [67]. Figure 4 illustrates the block schematic diagram of PTR-MS. PTR-MS exhibits excellent sensitivity and selectivity which play a crucial role in the exhaled breath analysis to monitor the pathophysiological and physiological state of the human subjects. In 2004, Amann et al. utilized PTR-MS to study the variation in the concentration of different VOCs exhaled in patients during sleep with carbohydrate malabsorption, and inter- and intra-subject variability of a certain mass [68]. Kar et al. utilized PTR-MS to detect cholesterologenesis by investigating the level of isoprene in the exhaled breath [69]. Later, Schmutzhard et al. reported the potential of PTR-MS to diagnose neck and head squamous cell carcinoma by monitoring isoprene in the exhaled breath [70]. Overall, this technique has been exploited widely to detect various VOCs and hence diagnosed the stages of corresponding diseases.

4.4. Advantages and Limitations of Classical VOC Detection Techniques

These techniques utilize the ion-pair extraction of the analytes and quantification by mass spectrometry to detect the VOC concentrations. However, a limitation of these techniques is that they can be time-consuming and expensive. They also require a skilled technician to operate which can be done only for off-site analysis. A detailed discussion of the advantages and limitations of these techniques is provided in Table 2.

5. Metal Oxides (MOs)

Oxide is a classification of chemical compound that has one or more oxygen atoms including other elements in its composition such as CO2, H2O, etc. MOs are chemical compounds that are formed by metal and oxygen. In general, MOs are known for their active selective material for sensing applications that encounters optical, chemical, and electronic inducement for analyte molecules detection [71,72,73].

5.1. Zinc Oxide (ZnO)

MO nanostructures, such as ZnO, are n-type semiconductors that have a wide band gap energy (3.37 ev at 300 K), large excitation binding energy (60 meV) at RT [74], and a high surface-to-volume ratio, making it a promising candidate for VOC and gas-sensing detection by significantly improving its response [75]. ZnO has been extensively used as an active sensing material as it provides simple synthesis preparation, good mechanical stability, biocompatibility, and offers an admissible response to a different variety of VOCs [76]. Many researchers have reported and studied various forms of ZnO nanostructures, i.e., nanoparticles, nanorods, nanowires, nanosheets, nanotubes, thin films, and 3D hierarchical structures including their application in VOCs vapor detection [74,77,78]. These various structures can be prepared by well-known techniques such as radio frequency sputtering [75,79], hydrothermal [78,80,81], sol-gel coating [82], and solvothermal combined with calcination [74]. The sensing mechanism of the ZnO-based gas sensor is mainly based on the adsorption/desorption of the VOC and gas species onto the oxide active sites and the electronic effects that are produced by the contact resistance of the modified ZnO gas sensors with the noble metals [83].
Among the different VOCs, acetone and ethanol are most studied and reported to be detected by different ZnO nanostructures while inadequate numbers of studies are done for other VOCs such as acetylene, n-butanol, cyclohexane, and benzene. Hardan et al. investigated the use of ZnO thin films on thermally oxidized Si substrates to detect isopropanol, acetone, and ethanol and reported the highest sensitivity of acetone detection at around 500 ppm at 400 °C [75]. The mechanism was based on an ionosorption model where the change of the resistance of the metal oxide semiconductor (MOS) in the presence of the test gas was outlined. Bora et al. studied and investigated the ZnO nanorods coated on multimode optical fiber for different chemicals such as ethanol, methanol, toluene, and benzene for its vapor sensing and reported the ethanol average sensitivity to be 22.2% at 50 ppm [84]. The study offered a simple and cost-effective optical sensor utilizing a light scattering from ZnO nanorod structures with a specific length, diameter, and surface coverage of the nanorods. In 2016, Muthukrishnan et al. reported an alternate sol-gel dip coating approach to obtain ZnO thin film to reduce the cost and complications [82]. They investigated the sensing response of acetone, ethanol, and acetaldehyde gases by a chemoresistive technique and reported the response/recovery time in order of seconds. The highest selectivity for acetone was found to be 71.32% among other test gases along with the lowest limit of detection (LOD) of 2 ppm acetone at RT with a response of 1.08, which is calculated as the ratio between the change in the electrical resistance of the sensor in the air to the test gases. In 2017, a porous coral ZnO was studied and reported by Zhu et al. for the detection of formaldehyde, methylbenzene, methanol, and acetone with a concentration of 600 ppm at 360 °C [74]. With this special porous structure of ZnO, the fabricated gas sensor revealed a substantial gas response with a fast response and recovery speed for all four different gases. Recently, in 2022, Swargiary et al. also developed ZnO nanorods coated over Single-mode–Mutlimode–Single-mode (SMS) fiber to detect isopropanol (IPA) vapor at various concentrations as a biomarker for diabetes patients [78]. The fabricated sensors exhibited a high sensitivity of 0.053 nm/% for IPA vapor detection at RT. Figure 5 shows the morphology of ZnO nanorods coated on the SMS for VOC sensing.
Although the different ZnO nanostructures coated sensors have been reported to show an excellent response to the different VOCs, a high selectivity and sensitivity sensor to effectively detect multiple vapors in a mixed volatile environment is still challenging. For this reason, it can be considered as a drawback for using ZnO nanostructure alone as a sensing material for VOCs [85]. Some researchers reported that tuning ZnO nanostructures by doping, composite structures and/or using surface modification with different noble metals gives rise to the improvement of VOCs selectivity. In 2017, Ma et al. reported the Ag-doped ZnO/SnO2 sensor with hollow nanofibers in structures with a rapid response time of 5 s, with an excellent selective detection to 1 ppm ethanol at 200 °C. According to the report, a noble metal on the surface of ZnO is considered to be a site for adsorbate, catalysts, and surface reaction promoters and can improve the thermal stability of the nanostructure. The main advantages of this structure are its porosity and surface roughness morphology which makes it a promising candidate for efficient ethanol gas detection in the environment [86]. In the same year, Wang et al. exploited ZnO doped with Mn (MZO) to selectively detect 20 ppm acetone at 340 °C and CdO-activated MZO to selectively detect 20 ppm ethanol at 240 °C [85]. These gas sensors were also used as a sensor array to differentiate the acetone and ethanol mixtures at different ratios displaying the potential of a high-performance gas sensor. Furthermore, high selectivity and sensitivity with very stable response/recovery characteristics towards methanol among other VOCs were also reported by Mandal et al. with a rose-like ZnO microcube/MoO3 micrograss-based composite structure [87]. The most recent article reported in 2022 by Dong et al. prepared a novel composite of SnO2/ZnO hollow cubes that were synthesized by using a self-template hydrothermal method followed by a calcination process [88]. They explored the formation mechanism of the composite structure and its morphology, which is shown in Figure 6.
ZnO nanostructure formation was found to be on the thin wall of hollow cubes. SnO2 nanorods were found to adhere to those thin walls. With this structure, the VOC vapor sensing was carried out for formaldehyde, acetone, isopropanol, ethanol, n-butanol, methanol, benzene, and another three reductive gases (ammonia, hydrogen, and carbon monoxide) at 220 °C. Among all the VOC vapors, the hollow SnO2/ZnO cube sensor displayed excellent sensitivity, selectivity, and repeatability towards 100 ppm formaldehyde with stability for a long period of time and the highest response of 148 is achieved. This response was calculated from the ratio between the baseline resistance and the balanced resistance of the composite structure in the air and the test gases.
The selectivity of ZnO-based VOC sensors can be further improved by working on the features of the sensitive material as well as the operating parameters of the sensor. In addition, various limitations, and challenges, such as ultralow concentration detection, humidity interference from the individuals, and variability in VOC profiles of the individuals from the time of food intakes, have to be critically considered. Therefore, by changing the property of ZnO nanostructure morphologies with the doping and/or composite materials, a high selectivity sensor in a mixed gaseous environment can be accomplished with an effort to overcome the associated limitations. This can serve as the future development of multifunctional, compatible, and robust ZnO-based VOC sensors.

5.2. Nickel Oxide (NiO)

Nickel oxide (NiO) is an intrinsic semiconductor and is widely used for VOC sensing due to its unique characteristics such as wide bad gap, stability at high temperatures, transparency, and excellent chemical stability. The accomplishment of NiO depends on its morphology, dimension, and phase state such as porosity and dimensions.
In 2020, Nakate et al. studied the performance of the gas response of NiO nanosheets sensor via a hydrothermal chemical route which was prepared by mixing nickel nitrate solution and sodium hydroxide solution in de-ionized water [89]. The morphology of the NiO nanosheet is shown in Figure 7a. Ultra-thin nanosheet morphology of NiO can give a high surface area and active sites that play an important role in the high adsorption of gas to improve the efficiency of the sensor. The highest gas response reported in this work at 250 °C was 191% for 150 ppm hydrogen concentration with a response time of 150 s. In 2022, Li et al. developed ultrathin porous NiO nanosheets for acetone gas sensing by using the facile solvothermal method. They developed various NiO nanosheets on interdigitated electrodes by controlling the adjustment of ethanol to distilled water ratios, i.e., 0:8, 1:7, 1:3, 1:1, 3:1, and 8:0, which was referred to as NiO-0, NiO-12.5, NiO-25, NiO-50, NiO-75, and NiO-100, respectively. Among them, NiO-75 (75% ethanol solution-based NiO gas sensor) exhibited a high selectivity towards acetone at 225 °C, with a low LOD of 0.8 ppb along with a fast response and good stability due to its porous structure, large surface area and a crystal facet compared to the other NiO nanosheets as shown in Figure 7b [90]. The advantages of this method are low cost and easy to control. The obtained NiO nanosheet gas sensor (NiO-75) can be considered a suitable candidate for commercial acetone gas detection.
Meilin and Zhen developed mesoporous NiO for an ammonia sensor via the sol-gel method in 2019 [91]. The synthesized mesoporous NiO has a uniform rod structure with a pore size of 4 nm as shown in Figure 7c. The study involved stability which showed a long-term constant response from 0.2 ppm to 0.4 ppm ammonia concentration in 21 days at RT. In 2022, Yu et al. also reported a new development of the mesoporous oxide semiconductor gas sensor by introducing bio-metallic MOF-derived core-shell mesoporous Sn-doped NiO sensitivity materials via hydrothermal and ion exchange method, as shown in Figure 7d [92]. The sensor exhibited a high sensitivity of 100 ppm concentration of xylene at 250 °C with a low LOD of 63 ppb due to its high pore interconnectivity that increases gas diffusion on the surface.
In 2021, Ayyala et al. developed a chemoresistive NiO-based sensor for the detection of seven different common VOCs that include acetone, ethanol, toluene, hexane, methanol, 2-propanol, and isobutylene vapors between 5 ppm and 25 ppm at 350 °C and 40% relative humidity (RH) [93]. They deposited NiO using a spin-coating technique on the sensor with two different film thicknesses of 5 μm and 10 μm. The morphology of NiO film showed to be microporous structure and more compact due to the interconnections between the structure itself. The NiO sensor with two different thicknesses was reported to have a fast response and recovery time for all VOCs which were less than 80 s and 120 s, respectively. The highest response of 1.5 was also achieved and calculated from the ratio between the electrical resistance of air and the tested VOC, i.e., 5 ppm ethanol at 350 °C for both sensors. Recently, John et.al. investigated the crystalline NiO and manganese (Mn)-doped NiO nanostrcutures prepared by the co-precipitation method for the detection of VOCs including xylene, toluene, n-butyl, alcohol, 2-methoxyethanol, methanol, ethanol, acetone, ammonia, and formaldehyde [94]. These studies included the analysis of structural, elemental, morphological, and optical properties of the prepared gas sensor. The results showed that the Mn-doped NiO sensor exhibited high sensitivity, selectivity, and a rapid response time of 5 s, at 100 ppm concentration of formaldehyde along with a low LOD of 1 ppm at RT. This reported work also offered one of the cost-effective and facile methods for synthesizing NiO and Mn-doped NiO porous nanostuctures.
NiO has been studied to improve the sensitivity, selectivity, and stability of the VOC sensor by utilizing the sol-gel, solvothermal, hydrothermal, spin coating and doping methods. The main sensing mechanism of the NiO-based VOC sensors is generally based on the adsorption, desorption, and oxidation processes. The pore size of NiO can be modified for VOC’s detection that depends on the size of the targeted VOC molecules. Generally, the highly responsive sensors mostly have mesoporous structures due to their good volatile gas adsorption properties. However, there were reports showing that NiO-based sensors has a slow response time and low limit of detecting different clinical VOC biomarkers in real-time from patients suffering from bladder cancer and cystic fibrosis [95,96]. These sensors had a limitation in accurately detecting VOC concentrations as there exists a continuous change in the metabolic process in the human body that may give rise to multiple VOCs at a time.

5.3. Molybdenum Oxides (MoO3)

Molybdenum oxide (MoO3) is an n-type semiconductor that can detect several gases, e.g., carbon monoxide [97], hydrogen [98], methanol [99], etc. The main sensing mechanism of MoO3-based VOC sensors is the adsorption and desorption of gas molecules in the sensing area. The gas adsorption of the MoO3 sensor depends on its morphology which can be categorized into three types of structures. Firstly, a structure with a high aspect ratio is a one-dimensional structure (1D) such as nanofibers, nanorods, and nanoribbons. They have a limited surface area, but their reaction sites are mostly exposed to the environment. Next, is a two-dimensional structure (2D), such as nanoplates and nanosheets, which provides a larger surface area that can improve the sensitivity of gas detection. Lastly, three-dimensional structures (3D), such as nanoarrays, nanoflowers, and nanospheres, are assembled from low-dimensional structures providing the largest surface area and have more reaction sites for gas adsorption. A larger dimensional structure contains more gas diffusion channels, allowing gas molecules to pass into reaction sites. Therefore, it shows better gas absorption and sensing performance towards VOC molecules.
The MoO3-based sensor has a limitation in terms of selectivity. Therefore, the modification of the MoO3 surface to increase the specifically targeted gas adsorption ability is required. The MoO3 surface can be modified by doping with transition metals such as Cr, W, Ag, Au, Fe, Zn, Ni, or other nanomaterials to reduce the activation energy of chemosorption reaction. Furthermore, the composition of various materials to obtain hybrid structures, i.e., a mixed structure of two metals such as MoO3 and Au [100], α-MnO2, and h-MoO3 [101] affects the properties of sensing material in terms of its grain boundary barrier, energy band, carrier concentration, and depletion layer resulting in the sensing performance improvement [102]. In 2016, S. Yang, et al. developed the MoO3 nanobelt sensor synthesized from the hydrothermal method to detect trimethylamine (TMA) [103]. The sensing response was calculated from the ratio between the electrical resistance of the sensor in air and in testing gas. The sensor showed a response value of about 6, 36, and 582 towards 1 ppm, 5 ppm, and 50 ppm of TMA gas at 240 °C, respectively, with a response time of 15 s. In addition, this sensor showed a higher selectivity to TMA when tested with TMA, ethanol, methanol, acetone, ammonia, and toluene. It was also reported that not only does the surface morphology of MoO3 help to improve the sensing performance but the doping of MoO3 with noble material also plays a significant role in the improvement of sensing performance [104]. In 2017, Shen et al. reported the doping of Ni to α-MoO3 nanolamella for formaldehyde detection by using the solvothermal method [105]. The α-MoO3 was doped with various doping concentrations of 2.5 mol%, 5 mol%, and 10 mol%. The morphology of α-MoO3 doped with different concentrations of Ni is shown in Figure 8. They all have the same lamellar shape with a lateral length of 220 nm to 500 nm. With a 100 ppm concentration of formaldehyde at 255 °C, the 5 mol% doping concentration showed the best response with a response time of 37 s and a recovery time of 2 s.
In 2018, Wang et al. developed the multi-component structure of Au/α- MoO3 nanobelts for ethanol detection by applying gold nanoparticles onto α-MoO3 nanobelts using the hydrothermal method [106]. The rod structured morphology was observed to be homogeneous and uniform with an average rod length of 5 µm to 10 µm along with a 4.3 nm diameter of Au nanoparticle. With the exposure to 200 ppm concentration of ethanol, the Au/α- MoO3-coated sensor exhibited a rapid response with a recovery time of 4.3 s compared to a conventional α- MoO3 at 200 °C. The reported material has been investigated over ethanol, acetone, xylene, toluene, ammonia, and formaldehyde and observed to be highly selective for ethanol. In 2019 Shen et al. synthesized porous α- MoO3 nanosheets to detect TMA with a faster response and recovery time of 12 s and 200 s, respectively [107]. The highest response to TMA was achieved when the sensor was investigated with various VOCs such as acetone, formaldehyde, ethanol, ammonia, TMA, and dimethylamine. The sensor showed a response of 198 ppm to 50 ppm TMA concentration with a LOD of 20 ppb. In 2019, a flower-like MoO3 structure was assembled from nanosheets for ethanol detection by Ji et al. [108]. There were three MoO3 flower-like structures synthesized by the hydrothermal method including (i). sphere-like, (ii). rose-like, and (iii). plate flower as shown in Figure 9. The sensing performance of the materials was investigated at 50 °C to 350 °C under the controlled concentration of ethanol of 300 ppm. The obtained results showed that the rose-like nano flower provided the highest response of 37.1 compared to the sphere-like nanoflowers and plate flowers which were 30.9 and 27.3, respectively.
It can be observed that several MOs have been exploited on a large scale to develop a real-time VOC sensor. MoO3-based VOC sensors can selectively detect specific gases due to their different morphologies. To increase the sensitivity and selectivity of the MoO3-based sensor, a doping material with a transition metal element is recommended. This technique can increase the sensitivity by reducing the activation energy of the chemisorption reaction, leading to more interaction between reaction sites and targeted VOC. Moreover, adding metal components to obtain the hybrid component structure can also be considered for improving the response and recovery times.

5.4. Titanium Dioxide (TiO2)

Titanium dioxide (TiO2) is a high-resistance n-type semi-conductor with a band gap of 3 eV. It gains huge attention in various applications including solar cells [109], photocatalysis [110], chemical sensing [111], and gas sensing [112]. The major advantage of TiO2 is its chemical stability, catalytic properties, and easy modification of its optical, transport and structural properties which plays a significant role in gas-sensing applications [113]. Furthermore, TiO2 represents three crystal structures in nature which are (i) anatase, (ii) brookite, and (iii) rutile. The rutile structure is considered as the most stable, while the brookite and anatase structures are metastable and can be irreversibly converted to the rutile structure by annealing at 600 °C to 800 °C [114]. Anatase is widely used in gas-sensing applications due to its prominent gas reaction capacity and high oxygen vacancies [115]. However, the pristine TiO2 gas sensors still have limitations that include low selectivity, high operating temperatures, unstable repeatability, and stability when it is exposed to oxidizing gases. In past years, several TiO2 nanomaterials with new composites and structures have been reported and used in VOC monitoring [115]. The morphology geometry of TiO2 nanostructures can be synthesized in zero-, one-, two-, or even three-dimensional nanostructures [116]. The TiO2-based VOC sensors are mainly based on the adsorption phenomenon due to the presence of various active sites suitable for VOCs [5]. In this section, a detailed discussion has been provided on the various types of TiO2 structures and their role in the enhancement of the sensing performance of VOC gases. The TiO2 nanostructures are considered a potential sensing material for ultrasensitive and miniaturized vapor sensors due to their large specific surface area which provides various active sites for the adsorption of gases.
The 0D TiO2 nanostructures include nanospheres, quantum dots, and nanoparticles, which can be synthesized using various techniques such as the sol-gel method [117], hydrothermal [118], metal-organic decomposition (MOD) [119], and laser ablation in liquid [120]. In 2015, Li et al. reported hollow TiO2 microsphere-coated chemoresistive gas sensors for the selective detection of formaldehyde [121]. The sensor exhibited a high sensitivity and excellent selectivity to sub-ppm level of formaldehyde at RT along with an average response time of 40 s. The response time of the TiO2-based sensor was improved by other nanostructures [115]. In 2018, Navale et al. synthesized anatase TiO2 nanoparticle-based chemoresistive sensor for the real-time detection of acetone at ppb level at 270 °C [118]. One of the key features of the fabricated sensor was the rapid response time of 10 s and low LOD of 500 ppb of acetone. The sensor exhibited high selectivity towards acetone in the presence of ethanol, methanol, ammonia, nitrogen dioxide, and cyclohexanol. The high selectivity for acetone is attributed to the higher adsorption activities on the TiO2 surfaces. In 2020, Tohru et al. reported sphere-like anatase TiO2 nanostructures using the MOD technique, which is a process combining the coating and heat sintering or photosintering for VOC sensing, as shown in Figure 10 [119]. In this study, a thin film of TiO2 nanocrystalline structure was synthesized successfully. The fabricated sensor was able to detect 1-propanol with the highest response along with ethanol and methanol at concentrations varying from 50 ppm to 350 ppm. The sensor exhibited an excellent sensitivity of 40% to 60% with rapid response time of 1 s at 350 °C. The improved performance was achieved due to its large surface area and highly crystalline structure.
In 2021, Gakhar et al. synthesized fullerene-C60-encapsulated TiO2 nanoparticles for the selective and sensitive detection of formaldehyde [122]. The sensing material was synthesized via a chemical route using hydrated fullerene-C60 and sol-gel-derived undoped TiO2 nanoparticles. The average diameter of the C60-encapsulated TiO2 nanoparticles was observed to be 150 nm. The reported sensor exhibited a double response in the magnitude of 117% with high selectivity and a fast response time of 12 s for various concentrations of formaldehyde from 1 ppm to 1000 ppm. The efficient sensing performance towards the formaldehyde was achieved in C60-encapsulated-TiO2 nanoparticles due to (i) the extremely reactive surface provided by the oxygen-functionalized C60 and (ii) an electronic exchangeability between formaldehyde and the TiO2 nanoparticles via C60 layers.
The 1D TiO2 can be considered an ideal nanostructure for VOC detection due to its inherent properties of a strong adsorption capacity, large specific surface area, superior electron transport rate, and availability of the porous sites for faster diffusion of VOC molecules [123,124]. 1D TiO2 nanostructure includes nanorods, nanotubes, nanobelts, nanowires, and nanofibers. Several methods have been reported for the synthesis of 1D TiO2 including hydrothermal [125], electrochemical anodization [126], electrospinning [127], template-assisted synthesis [128], and matrix-assisted pulsed laser evaporation [129]. Among all these techniques, vertical TiO2 nanotube arrays synthesized by using electrochemical anodization contain more active sites for effective gas diffusion. The pristine 1D TiO2-based gas sensors have been successfully utilized to detect various VOCs, e.g., acetone [130], isopropanol [131], ethanol [132], formaldehyde [133], isopropanol [134], etc. In 2015, Dutta et al. performed electrochemically synthesis of TiO2 nanotubes for the real-time detection of benzene, toluene, and xylene (BTX) with enhanced sensitivity at a low temperature of 50 °C, as shown in Figure 11 [135]. Among these three VOC targets, benzene was observed to show the highest response magnitude followed by toluene and xylene at all concentrations varying from 20 ppm to 400 ppm.
In 2019, Bindra et al. fabricated a sandwich-structured sensor using Au/TiO2 nanotubes/Ti for the selective detection of four different VOCs including ethanol, methanol, acetone, and 2-propanol at RT [136]. The TiO2 nanotube array was grown via the electrochemical anodization technique which exhibited a maximum capacitive response towards the methanol compared to the other VOC test gases. However, it cannot completely recover its response for all other test gases. Researchers have also synthesized single-, double-, and triple-layered TiO2 nanotube arrays by using a voltage pulse-assisted anodization method [137]. In 2020, Zhao et al. synthesized a double-layered TiO2 nanorod array on fluorine-doped tin dioxide for the detection of low-concentration ammonia at RT [138]. The obtained results showed that the TiO2 nanorods have a growth time of nearly 6 h with a high gas-sensing response of 102% for 100 ppm to ammonia at RT. The most recent article reported by Cai and Park, in 2022, showed a synthesis of cobalt (II,III) oxide (Co3O4)-decorated porous TiO2 nanofibers by using the facile electrospinning technique to detect acetone [139]. The Co3O4 nanoparticle of diameter 30–50 nm was decorated on TiO2 nanofibers via the solvothermal process. The fabricated sensor exhibited a good response for acetone (71.88) compared to the other VOCs including ethanol (7.58), toluene (2.55), benzene (2.01), and p-xylene (3.01). In addition, the sensor exhibited an average response and recovery time of 122 s and 351 s, respectively, for 100 ppm acetone gas at 250 °C.
The 2D TiO2 exhibited several versatile features including high specific area, flexibility, quantum Hall effect, and superior mechanical strength [140]. The 2D TiO2 nanostructure includes thin film, nanoplates and nanosheets. The 2D TiO2 subunits synthesized by using crystal facet engineering can significantly improve the selectivity and surface reactivity [141]. Several strategies have been employed to synthesize the 2D TiO2 nanostructures such as the chemical decomposition method [142], the one-step annealing process [143], reactive evaporations [144], and the hydrothermal method [145]. Among all of these techniques, the hydrothermal approach is the most common technique for the synthesis of 2D TiO2 nanostructures. In the past decades, the pristine 2D TiO2 nanostructures have been employed for a wide range of VOCs including isopropanol [5], methanol [146], acetone [147], hydrogen sulfide [148], and ammonia [149]. In 2019, Wang et al. synthesized TiO2 nanoplate-based micro-electro-mechanical system (MEMS) to detect acetone at RT, as shown in Figure 12 [150]. The author synthesized TiO2 nanoplates with defective and complete {001} crystal facets by varying the concentration ratio of hydrofluoric and hydrochloric acid during the synthesis. The results showed that the sensing performance of TiO2 nanoplates with complete {001} facets was 70% higher compared to the defective TiO2 nanoplates. The poor sensing performance of the defective TiO2 nanoplates is attributed to fewer adsorption sites on the surface. The TiO2 nanoplates with a volume ratio of hydrofluoric acid to hydrochloric acid of 2/50 exhibited good sensing responses of 21.8, 8.46, 4.32, 4.08, 6.64, and 2.73, for acetone, ethanol, toluene, methanol, isopropanol, and benzene, respectively, at 400 °C.
In 2020, Ge et al. synthesized hierarchical porous TiO2 hexagonal nanosheets obtained from the layered TiSe2 nanosheet templates in a one-step annealing process [143]. The fabricated sensor exhibited a high selectivity towards acetone of 200 ppm concentration along with ultrafast response times of 0.75 s at 400 °C. The transportation of electrons between TiO2 and acetone was achieved quickly due to the unique porous hierarchical structure of TiO2 nanosheets, crystal facet engineering and strong interface coupling. Recently, Tian et al. developed a MEMS sensor to study the influence of multi-layer of TiO2/SnO2 heterojunction on the detection of ethanol [151]. In the reported work, SnO2 was proposed due to its smooth surface and small particle size which advances the diffusion of ethanol. Although the sensor exhibited a slower response time of 33 s, it can effectively detect a lower concentration (50 ppm) of ethanol at 260 °C along with the sensor response value of 7.54. The sensing response was calculated from the ratio between the electrical resistance of the sensor in the air and the testing gas.
Nowadays, 3D structured TiO2 nanoflower gains huge interest in VOC sensing. The TiO2 nanoflower structure comprises unique features allowing the diffusion and adsorption of VOC molecules onto its surface or penetrate into the interior part of these nanoflowers [152]. Moreover, some literature indicates that the TiO2 nanoflowers exhibit superior electronic properties compared to 0D, 1D, and 2D TiO2 nanostructures. This performance improvement comes from the structural defects, which act as active sites between the gas and TiO2 material [153]. The TiO2 nanoflowers can be synthesized by a facile hydrothermal method using titanium(III) chloride (TiCl3) [154], titanium tetrachloride (TiCl4) [155], Ti powders [156], titanium(iv) oxysulfate (TiOSO4) [157], Titanium(IV) isopropoxide (C12H28O4Ti) [158], or Ti foil [159] as raw material. In 2016, Bhowmik et al. synthesized 3D TiO2 nanoflowers consisting of 2D nanosheets utilizing a low-temperature hydrothermal technique [160]. The fabricated sensor exhibited a high selectivity and rapid response time towards acetone at 60 °C. In this work, the authors considered the effect of the bond dissociative activation energy, vapor pressure of the individual VOCs, acetone selectivity, and the diffusivity of the gas species at the TiO2 nanoflowers surface in order to describe its performance at low operating temperature. They investigated the dual effect of the low bond dissociative activation energy and the higher VOC vapor pressure of the acetone molecules compared to the other VOCs (toluene, propanol, methanol, and butanone). The result showed a maximum response towards acetone. In 2020, Wang et al. synthesized 3D hierarchical TiO2 nanoflowers to detect ethanol varying from 10 ppm to 500 ppm at RT, as shown in Figure 13 [156]. This nanoflower-like TiO2 nanostructure is composed of thin nanosheets and was synthesized by the hydrothermal technique. The sensor exhibited good sensing performance in terms of good stability, high selectivity, rapid response/recovery times, and good reproducibility against ethanol vapor at RT. The nanoflower structure and the spacing intervals provide more adsorption sites for the oxygen species and the VOC gases, leading to exceptional gas-sensing performance of the material.
In summary, TiO2 can be categorized as 0D, 1D, 2D, and 3D nanostructures. Each nanostructure exhibited an excellent sensing performance towards VOCs including selectivity, sensitivity, and rapid response time. From the above discussions, it can be seen that the higher dimension structure of TiO2 showed rapid response and recovery time. The lower dimension structure provided a poor response, but it is easy to synthesize. Even though TiO2 alone can successfully detect several VOCs, it fails to show a selective response in the mixed gaseous environment. Therefore, the noble metal loading/composition technique has been utilized by several researchers to enhance the selectivity for target-specific detection.

5.5. Advantages and Limitations of MO-Based VOC Sensors

The performance of MO (such as ZnO, NiO, MoO3, and TiO2)-based VOC sensors can be improved by reducing the particle sizes to a nanoscale level in the sensing layer. Thus, a larger surface area and a higher active site can be achieved leading to excellent sensitivity. Changing the property of MO-based sensors by doping and/or adding composite materials may offer an option for improving the selectivity and LODs. However, there still exist numerous challenges and limitations on VOC profiling for clinical samples in a multivariate environment and conditions. The MO-based sensors should also be able to overcome the major limitations of the VOC profiling approach such as the extremely high costs of the laboratory instruments based on accurate VOC detection, and standardized sample collection for clinical trials. In addition, the detection of ultralow concentrations of associated VOCs from te different persons, the development of generic MO-based sensors for their durability, robustness, selectivity, reliability, energy efficiency, and standard calibration of these sensors upon malfunction still remain a challenge.
In Table 3 we have summarized the various kinds of metal oxide sensors and their sensing performances.

6. Carbon Based Materials and Composites

6.1. Graphene Oxide (GO)

The usage of GO is considered another potential candidate for non-invasive VOC vapor detection due to its unique and desirable properties such as being thermally stable, large surface area, and highly electrically conductive. These properties of GO have been exploited over the past years by several researchers to detect VOC vapor. The mechanism of graphene and its derivative sensors is based on the change in the resistance that is associated with the adsorption/desorption of the VOC molecules on the graphene surface which leads to a change in its properties. These VOC molecules can easily adhere to the large surface area of the graphene, changing the conductivity of the graphene sheet [186,187,188,189]. The change in conductivity can be used to measure the concentration of the VOC vapor in the environment with a high level of sensitivity due to the large amounts of charge carriers and electron mobility. However, due to the current availability of fabrication techniques and technologies, it is extremely difficult to fabricate pure graphene using economically viable techniques. Therefore, most applications that require the usage of graphene-based structures rely on GO, which can be mass-produced by using chemical exfoliation. On the other hand, the usage of GO to detect the VOC vapor have several limitations such as a reduction in the level of conductivity due to oxygen and other oxygenated groups inside the graphene matrix which reduces the sensitivity of the VOC vapor sensor. This allows the graphene structure to be more vulnerable to a humid environment [190]. To overcome this limitation, reduced graphene oxide (rGO) is often used instead of pure GO. rGO contains a lower level of oxygen and oxygenated groups in the graphene matrix, allowing it to function more like pure graphene and minimizing the negative effects of the oxygen and other oxygenated molecules. However, the usage of rGO alone is not efficient because of the other types of VOCs that are attached to the sensor, resulting in low selectivity. Therefore, it is usually more efficient to utilize rGO with another compound or material to make the rGO structure more selective to the target VOCs.
One example of the integration of rGO can be seen in the works of Zito et al. reported in 2018, where a combination of rGO sheets and NiO composite in the hierarchical flower-like structure was used to detect methanol [191]. Integrating rGO increases the number of holes on the surface of NiO due to the transfer of the charged carriers to balance the difference in the Fermi levels of rGO and NiO. Oxygen is oxidized in the presence of methanol, allowing the electron to recombine with the hole. This increases the resistance, which is used to measure the concentration of methanol. Thus, the sensitivity and response of the sensor can be improved. In 2019, Modenes et al. worked on rGO combined with copper (II) oxide (CuO) hierarchical structure to create a VOC vapor sensor for ethanol detection [192] as shown in Figure 14. Figure 14a shows a FE-SEM image of the CuO hierarchical structure that was used as the basis of the VOC vapor sensor, while Figure 14b is the integration of CuO and rGO, where rGO is composed of 2% of the weight of the entire solution. Even though CuO has previously been used to detect gaseous particles such as VOC vapor, using it alone gives a low level of sensitivity. Therefore, integrating rGO with CuO allows the VOC sensor to take advantage of both materials to detect ethanol. With the presence of ethanol particles, the ionized oxygen adhered onto the surface will be oxidized, resulting in electron-hole pair recombination. This reduces the overall level of CuO conductivity. When oxygen atoms adhered onto the surface of CuO, oxygen is ionized into a negatively charged ion, which leaves a hole in the CuO matrix as CuO is also a p-type semiconductor. These holes accumulate on the surface of CuO, thus allowing the surface to be more conductive and sensitive to ethanol.
In the same year, Pargoletti et al. fabricated a novel composite material based on GO sheets that were decorated with tin dioxide (SnO2) nanoparticles by controlling chemical growth. They reported the composite materials, SnO2-GO (4:1), SnO2-GO (8:1), and SnO2-GO (16:1), and pristine SnO2 material for the detection of ethanol at low concentrations. Among them, the hybrid material SnO2-GO with a ratio of 16:1 was found to be superior for ethanol sensing at 1 ppm concentration both at RT (exploiting the UV light) and 150 °C with response and recovery times of 310 s/320 s and 70 s/75 s, respectively [193]. In 2020, Wu et al. reported a NO2 vapor sensor using a combination of rGO and SnS2 [194]. rGO was synthesized into a 3D porous structure to increase the surface area and the density of sites in such a way that NO2 particles can adhere to the surface. SnS2 is integrated into rGO to increase the selectivity as it is an n-type semiconductor that has a special affinity allowing the adsorption of NO2 molecules due to the high electronegativity of a semiconductor. However, SnS2-based sensors are not usable at RT when used alone due to a high electrical resistance value. Therefore, integrating rGO with SnS2 to develop a NO2 sensor can allow rGO to benefit from SnS2 by achieving a high level of selectivity for NO2 gas, while SnS2 can utilize the superior level of conductivity from rGO to overcome its inherent resistance value and operate within RT.
The sensitivity of GO and rGO VOC vapor sensors can also be improved by integrating them with nanostructured materials. In 2021, Zhao et al. reported the use of rGO in conjunction with SnO2 nanorods to develop a formaldehyde sensor [195]. The sensing efficiency of SnO2 depends on the morphology. In this report, the SnO2 nanorods structure is the most suitable structure integrating with rGO. The sensor showed a 150-times better response to formaldehyde due to larger available sites from the nanorods structure. Another research on Ag nanoparticles modified with Fe3O4/rGO composite through hydrothermal method for the detection of acetone was done by Jia et al. in 2022 [196]. The composite structure Ag/Fe3O4/rGO showed an ultra-high response (Ra/Rg = 35.81) at 50 ppm acetone concentration at 220 °C which was 2.5 times higher than pure Fe3O4/rGO composite structure. This is due to the chemical and electronic excitation effects of Ag nanoparticles which generate more oxygen species and active sites for excellent gas sensing performance. Moreover, introducing the noble metals along with the metal oxide and rGO offers excellent sensitivity, good response, and selectivity of the device for practical applications.
In 2019, Wang et al. developed a ternary nanocomposite NO2 sensor consisting of rGO, ZnO and SnO2 [197]. They reported that the NO2-sensing properties of the sensors on rGO improved massively from the formation of the heterostructures between the two metal oxides. With this nanocomposite, the electronic properties of the hybrids and the band structure change thereby promoting effective charge transfer among the interfaces of heterostructures which improves the gas-sensing performance. With this advantage of ternary structure-based sensors, a similar kind of work was also reported by Sen et al. in 2021 using the same materials for the detection of VOC vapor [198]. Therefore, these constructions of heterojunction have displayed the potential for enhanced VOC detection applications as well. ZnO is one of the building materials in a VOC sensor because it has a good level of response to gas molecules and has high stability in many operational environments. However, it must be integrated with other materials such as rGO to minimize a level of resistance. By adding ZnO to SnO2/rGO sensor, a p-n-n heterojunction is created allowing a larger surface area for the VOC vapor absorption. This can significantly increase the sensitivity and response of the sensor for multi-VOC vapor detection in mix gaseous environment.
Most rGO and GO-based VOC vapor sensors are usually fabricated to detect/target only one type of VOC vapor species which is the main limitation of this sensor. Therefore, it is highly desirable to develop a sensor that can be reliable, efficient, cost-effective, and possess a standard calibration for detecting multiple types of VOC vapor in terms of a clinical point of view in real-world applications. rGO is incorporated with other materials to take advantage of the desirable properties of each material to improve the VOCs sensing ability [10]. In addition, other types of rGO and GO-based sensors such as a ternary-based sensor also offer the ability to detect multiple VOCs.

6.2. Multiwall Carbon Nanotubes (MWCNTs)

Carbon is one of the most abundant elements on the planet. It is a promising material for nanoelectronics as it is able to form itself into various crystal structures, resulting in a different material property. Due to its superior properties from the four valence electrons, allows the mixing of s and p atomic orbitals to form several hybridization, i.e., sp2 and sp3 hybridization. The perfect hybridization provides a higher electrical conductivity, larger specific surface area and other improved electrochemical properties [199]. In addition, carbon can be formed into various nanostructures such as fullerene zero-dimensional structure, carbon nanotube one-dimensional structure, graphene two-dimensional structure, fullerite 3D structure, and carbon nanotube which have an extraordinary advantage in VOC sensing. The primary sensing mechanism of MWCNT-based VOC sensors is the adsorption of VOC molecules on the surface leading to a change in the output of the sensor.
A carbon nanotube (CNT) is a two-dimensional cylindrical structure with the thickness of one graphene arranged in a hexagonal lattice where each atom forms covalently with the sp2 hybridization that is even stronger than the diamond lattice formed in the sp3 hybridization [200]. CNTs have been widely reported with superior electrical, mechanical, and thermal properties that can be up to multiple times greater than typical steel or even some class-leading materials [201]. For this reason, there are many possibilities in several potential applications ranging from material enhancement, surface modification, biological applications, and electronic applications [202]. Many researchers have reported that the CNTs are able to transport electrons without any scattering phenomena due to their hollow shape in angstrom (Å) diameter. Therefore, the electrons can be quantized and confined along the circumferences and proceed only in the axial direction with a thousand times better mean free path compared to copper. Furthermore, the resistivity of the CNT can be as low as microohm per centimeter range [203]. The superior mechanical properties can be described as up to 63 GPa in tensile strength and 1 TPa in Young’s modulus which is greater than the 10-fold industrial-grade fibers [204]. The ideal CNTs have very high thermal conductivity, approximately 3000 W/m·K, which is eight times greater than copper and two- times greater than the copper diamond with sp3 hybridization [205]. In addition, Deng et al. reported in 2014 that the coefficient of thermal expansion of CNT can be as low as 1.9 × 10−5 K−1 at RT which hiders the change in shape allowing more possible applications [206].
Carbon atoms can be arranged in a variety of nanostructures including CNTs which can be categorized into two major types, single-walled carbon nanotubes (SWCNTs) and MWCNTs. CNTs can also be categorized according to their chirality to be armchair, zigzag, and chiral, which represent their electronic properties to be metallic-like or semiconductor-like properties. The MWCNTs have a large surface area, high conductivity, and excellent flexibility in chemical stability compared to the SWCNTs [207]. Therefore, it is considered the preferred candidate for the VOC gas-sensing applications. In 2017, Bahoumina et al. reported a microwave flexible VOC sensor based on poly (3,4-ethylenedioxythiophene) polystyrene sulfonate–multi-wall carbon nanotubes (PEDOT:PSS-MWCNTs) [208]. The sensing material was deposited using inkjet printing technology. The sensing response of the device was estimated at 4 min and 10 min exposure to ethanol vapor at 0 ppm, 500 ppm, 1000 ppm, and 2000 ppm concentration. The maximum sensitivity of the device has been estimated at −642.9 Hz/ppm and 648.1 Hz/ppm at 4 min and 10 min exposure of ethanol, respectively. In 2018, Liu et al. developed a RT VOCs vapor sensor utilizing layer-by-layer MWCNTs/poly-ethylene glycol (PEG) composite, as shown in Figure 15 [209]. The sensing performance of the device was calculated by measuring the relative resistance change when the sensor was exposed to the VOC’s environment. The sensor exhibited a high sensitivity to acetone, isopropanol, ethanol, and isoprene along with a rapid response time of 110 ± 5 s and a lower LOD of 9 ppm at RT.
In 2019, Janfaza et al. reported a highly selective chemoresistive sensor to detect hexanal composite layers of molecularly imprinted polymer (MIP)/MWCNTs on RT [210]. The result exhibited that the electrical resistance was increased with the concentration of hexanal due to the interaction of the vapor with functional groups on MIP leading to the swelling of the MIP. The MIP swelling allows moving MWCNTs away from each other and changing the conducting path created by MWCNTs in the sensitive film, leading to a change in the resistance of the sensor. Therefore, a significant increase in the resistance of the sensor occurs when interacting with hexanal. The sensor showed a linear sensing response with a low LOD of 10 ppm for the concentration varying from 10 ppm to 200 ppm. In 2020, Singh et al. synthesized molybdenum disulfide (MoS2)/MWCNTs composite layer using hydrothermal route to detect ammonia ranging from 12 ppm to 325 ppm, at RT [211]. In this work, a comparative study has been performed on two configurations Ag/MoS2/Ag and Ag/MoS2/MWCNTs/Ag. The device exhibited n-type semiconducting behavior and is capable of detecting ammonia as low as 12 ppm. The SEM image of MoS2 and MoS2/MWCNTs is shown in Figure 16. In the case of the MoS2 configuration, the sensor exhibited a slow response time of 400 s which was later significantly improved by the MoS2/MWCNTs composite layer to be 65 s. It has also been discussed in the report that these significant differences appeared due to the different adsorption energy for different VOC molecules on the surface of MoS2 or MoS2/MWCNTs composite.
The sensing performance of MWCNTs was later improved in 2022 by Yan et al. utilizing polypyrrole nanocomposite-coated with MWCNTs [212]. Thermoplastic polyurethane (TPU) conductive nanocomposites were fabricated using TPU as a matrix and polypyrrole (PPy) coating MWCNTs as a conductive filler. The fabricated TPU/MWCNTs-g-PPy nanocomposite comprised certain gas-sensitive response selectivity, especially for ammonia gas compared to the other VOC molecules. The sensor exhibited improved response values in the range of 90% to 100% with a rapid response time of 40 s to 50 s.
MWCNT can be considered one of the promising candidates in the VOC sensing application due to its superior properties, e.g., electrical, mechanical, and thermal properties. However, the major limitation of the MWCNTs sensor is the deterioration of the structure during long exposure sessions of the VOCs. This can cause a false output measurement which can be very crucial in some applications. Nonetheless, these challenges can be minimized by integrating with composite materials such as PEG, MoS2, PPy, MIP, etc. These composite materials not only solve the problem of deterioration of material but also improve performance in terms of sensitivity and response time.

6.3. Activated Carbon (AC)

With the rapid progress of industrialization and constant development of urbanization, VOCs and their toxicity have been a serious concern not only in the environment but also for human health [213]. With the constant increase in VOC emission, various techniques to degrade the VOCs are being investigated and implemented. One of the techniques is the add-on control where the VOC is absorbed by absorption materials due to physical and chemical interaction in exhaust gas molecules [214]. This technique allows effective, efficient, and economic measurements [215].
ACs are a type of absorption material that has been widely used due to their good physical and chemical properties with a large specific surface area, high chemical stability, high mechanical strength, and abundant functional groups along with alkali and acid resistance. Although ACs have these advantages, they also possess limitations due to their thermal stability and hydrophobic property. The ACs spontaneously ignite or collapse in a porous structure at high temperatures as it is a nonpolar absorbent. Therefore, it limits the adsorption towards hydrophilic VOCs [216,217,218]. ACs are generally produced in carbonization and activation process by carbon-rich materials such as coconut shells, wood, sawdust, lignin, palm shell, wood, petroleum pitch, and coal [219]. Figure 17 shows the SEM images of the precursor (raw palm shell AC) obtained under the optimum conditions developed by Arami et al. in 2012 [220].
There are several studies on ACs for the VOCs adsorption and its recovery such as alkane, alcohols, ethers, aldehydes, ketones, esters, aromatics, etc. In 2012, Li et al. investigated the influence of the ACs pore structure on the adsorption of VOCs [221]. They reported the fixed-bed thermostatic adsorption at certain conditions where the adsorption capacity of ACs to acetone, toluene, and 1,2-dichloroethane was conducted. The obtained results showed the selectivity of the adsorption of organic gases and the VOCs including toluene, acetone and 1, 2-dichloroethane adsorbed on the AC are in the different pore diameter range of (1.27 to 1.49) nm, (0.67 to 0.84) nm, and (1.39 to 1.75) nm. In 2019, Shen et al. prepared AC from rice husk (RH) and investigated the adsorption performance of toluene and phenol [222]. They reported the adsorption capacity for toluene (264 mg/g) to be higher than phenol (6.53 mg/g) in the gaseous phase. This is due to toluene being one of the strong volatile compounds that are greatly absorbed by RH char in ambient conditions. On the other hand, phenol is a semi-volatile organic compounds (SVOCs) that has a weaker volatility in the ambient, thus possessing a weaker adsorption capacity in the gas phase compared to the liquid phase. They also reported that AC can be reusable through thermal desorption. In 2020, Kang et al. proposed a gas concentrator for breath analysis of individual health to diagnose diseases by evaluating the adsorption and desorption performance from a commercial photoionization detector (PID) and a colorimetric tube detector [223].
The VOC investigated in this study was 1-ppm propionic acid. Its adsorption and desorption rates obtained were approximately 90% and 70%, respectively, making it an excellent gas concentrator which is benefited by low concentration for VOC detection. In the same year, Ma et al. studied the adsorption and adsorption mechanisms of toluene and chlorobenzene by two commercial ACs under medium-high-temperature (MHT) conditions [219]. The two ACs are coconut shell-based AC (CSAC) and wood-based AC (WAC) which were investigated in a fixed-bed reactor under medium-high temperature (MHT: 90 °C to 150 °C). The obtained results showed that the toluene and chlorobenzene adsorption characteristics at MHT conditions were different from those at low temperature adsorption. The adsorption values of toluene on CSAC and WAC at MHT conditions were 20.6 mg/g to 49.9 mg/g and 13.9 mg/g to 46.9 mg/g, while for chlorobenzene were 30.5 mg/g to 80.3 mg/g and 26 mg/g to 63.9 mg/g, respectively. The results provided the basic data and theoretical references for the development of adsorbent injection coupled with bag filtering technology to remove organic pollutants. From the above study, we can see that the mechanism of the AC-based VOC sensors is mainly reported in terms of adsorption rates, adsorption capacities, and their operation under the controlled temperature. However, the response and recovery times of the AC-based sensor are still limited. Moreover, limited studies can be found on the selective optimum adsorption capacity for VOC mixtures. Thus, tailoring the physical and chemical properties of AC, such as integrating with doping materials, can enhance the VOC adsorption and allow the detection in sub ppb/ppm level which is a key criterion for clinical VOC biomarkers detection. ACs have promising properties such as large surface area, variable surface chemical functionalization, and tunable micropore size and volume. Therefore, AC is considered to be a very good candidate for achieving high adsorption efficiency/capacity for the VOC detection ranging from dozen to several hundreds of mL/g.

6.4. Advantages and Limitations of Carbon-Based VOC Sensors

Nowadays, the development of the high-performance VOC gas sensor using carbon and carbon composite-based sensing materials offer great advancement and progress in terms of the structure design [224]. However, the developed VOC sensors using carbon-based materials have not yet succeeded in achieving and realizing their commercialization in environmental gas detection, diagnosis of non-invasive disease, and a portable health-monitoring device. The complexities of the sensing material require the fabrication of a high-performance VOC sensor. The field needs to pay a continuous effort for further improvement of the overall sensing properties of VOCs sensors employing carbon-based nanomaterials, including sensitivity, specificity, low LOD, RT operational, rapid response, fast recovery time, good reproducibility, long-term stability, and capability of intelligent readouts.
In Table 4 we have summarized various kinds of carbon-based VOC sensors and their sensing performances.

7. Polymers

A polymer is a substance or material consisting of very large molecules called macromolecules, made by linking up smaller repeating chemical units. In the past decades, polymers have gained huge interest in the field of the development of artificial sensors due to their special characteristics of swelling and shrinkage in the presence of gaseous molecules [243]. Rapid response time and better selectivity have been achieved by replacing classical sensing materials with polymers involving nanotechnology and exploiting either the extrinsic or intrinsic functions of polymers. Polymers play a crucial role in the detection of various chemical and biological fluids due to their swelling and shrinkage characteristics [244,245,246].

7.1. Poly (Dimethylsiloxane) (PDMS)

Poly (dimethylsiloxane) (PDMS) is a silicon-based elastomeric polymer which is formed by the cross-linked structure of the repeating units of -OSi (CH3)2- groups. PDMS has been considered an ideal material for the fabrication of integrated circuits and microfluidic channels [247,248]. Furthermore, it plays a crucial role in VOC sensing over the last decades due to its swelling property in the presence of VOC molecules [249]. The swelling and shrinkage of PDMS in presence of various VOC concentrations leads to a change in the output of sensing devices.
In 2012, Reddy et al. developed Fabry–Pérot interferometer (FPI) based optical vapor sensor array for micro-gas chromatography (μGC) applications [250]. The reported sensors were made by depositing a uniform thin layer of polymer on a silicon wafer. The polymer–silicon and air–polymer interfaces form the Fabry–Pérot (FP) cavity in which the resonance wavelengths varied with the VOC vapors allowing the rapid detection and quantification of VOCs. For the proof-of-concept, two polymers, namely PDMS and SU-8, were utilized independently and placed in a microfluidic channel. The sensor exhibited different sensitivities for the different vapors. It showed an excellent response to toluene and acetone for a concentration ranging from 0 ng to 20 ng and 0 ng to 150 ng, respectively. In 2016, Ning et al. reported a sensitive FPI and Sagnac Interferometer sensor coated with PDMS for the simultaneous detection of multiple VOC targets [251]. From the fabricated device, authors have successfully detected ethanol and 2-propanol by using second-order inverse matrix approach. The sensing head was fabricated using a single-mode hollow-core optical fiber coated with a 15 μm thick PDMS on the tip of fiber. The air between fiber and PDMS represented the FPI cavity (detail information on this interferometry technique is provided in our most recent review article [17]). The reported sensor exhibited a good sensitivity of 1.61 × 10−3 nm/ppm and 1.17 × 10−3 nm/ppm for 2-propanol and ethanol, respectively. Later in 2017, Kacik and Martincek utilized a similar approach to detect toluene using FPI formed by a micro air-cavity inside the PDMS matrix coated at the fiber-end facet [252], as shown in Figure 18. The interface between the fiber/air and air/PDMS served as two reflecting surfaces for reflected beams to form an interference pattern. The sensor exhibited an average sensitivity of 0.15 nm/g.m−3 to 1.4 nm/g.m−3 for the toluene concentration varying from 0.833 g·m−3 to 100 g·m−3.
In 2019, Zhao et al. developed PDMS coated optical fiber sensor to form a FPI structure, as shown in Figure 19 [253]. The 1 mm long fabricated sensor was made from a small section of large mode area fiber (LMAF) and a segment of hollow-core fiber (HCF). The sensitivity of the device was determined by monitoring the interference spectral due to the Vernier effect. The fabricated sensor exhibited an excellent sensitivity of 20 pm/ppm for isopropanol varied in the range from 0 ppm to 500 ppm.
Recently, Sappati et al. reported a printed acoustic wave sensor with PDMS as a sensing layer for the detection of the low concentration of VOCs [254]. In this work, an acoustic flexural plate wave sensor was printed with silver ink on a thin and flexible lead zirconate titanate-PDMS composite layer. The prototype resonator showed a resonant frequency of 22.65 MHz with an attenuation of −1.552 dBm. The gravimetric mass sensitivity of the device was monitored by introducing PDMS layers between the input and the output interdigital transducers. The sensor exhibited a high sensitivity of 0.66 kHz/ppm and 160.63 kHz/ppm and a low LOD of 10.9 ppm and 0.03 ppm for acetic acid and toluene, respectively.

7.2. Conducting Polymers (CPs)

Monitoring VOCs using conducting polymers (CPs) is at the core of attention for the fabrication of next-generation sensor applications. The CPs are recognized as a class of organic materials with unique optical and electrical features similar to those of metal and inorganic semiconductors. The major advantages of these polymers are their processability [255]. CPs can be synthesized using versatile, simple, and cost-effective approaches which makes them potential candidates in sensing applications. CPs are composed of functional groups which have pseudo capacitance features and hence exhibit conductivity as the material itself such as polyacetylene, polypyrrole, polyaniline, polythiophene, poly(phenylenevinylene), poly(para-phenylene), and polyfuran. In the next subsections, we briefly discussed two major CPs, polypyrrole and polythiophene, and their roles in VOC sensing. CP is an important sensing material based on swelling and shrinkage phenomena to monitor the changes in VOC concentrations.

7.2.1. Polypyrrole (PPy)

In recent years, PPy has been exploited in VOC sensing due to its salient features of high selectivity and sensitivity towards the detection of inorganic and organic gases [256]. PPy, is considered one of the emerging sensing materials with a wide range of applications in the field of electronic, optical, and electrochromic devices and sensors. It becomes a promising alternative in sensor development due to its stability in a different gaseous environment, ease of deposition from aqueous and non-aqueous media, adherence to diverse substrates, and high electrical conductivity.
The composite PPy-PVA film showed better sensing performance and rapid response time towards ethanol among other VOC vapors including ethanol, ammonia, chloroform, toluene, and acetone [257]. Among several conducting polymers, PPy has been exploited widely to detect VOC gases at RT [256]. PPy contains several unique and desirable characteristics such as its stability in different environments, high electrical conductivity, ease of deposition from aqueous and non-aqueous media, and its adhesion to various substrates makes it an ideal material for sensing applications that are required for the detection of various kinds of VOCs. With the unique properties of PPy, Li et al. developed a composite of SnO2 nanosheet coated with PPy to detect ammonia at RT, as shown in Figure 20 [258]. SnO2 is an inorganic semiconductor gas-sensing material that exhibits poor conductivity at RT with no significant response to ammonia gas [259,260]. However, there was a report showing a profound effect on the electrical properties of PPy and its sensing performance towards ammonia. Therefore, in the reported work authors utilized SnO2 with PPy to combine the features of both materials to improve the performance of sensors. The nanocomposite sensors exhibited a higher sensitivity of ∼6.2%/ppm for ammonia concentration varying from 1 ppm to 10.7 ppm, with a low LOD ∼257 ppb. The sensor coated with SnO2/PPy exhibited a higher response for ammonia compared to the sensor coated with PPy alone.
In 2019, Wu et al. reported a PPy-coated polarization-maintaining fiber sensor to detect isopropanol [261]. The sensing configuration was based on the combination of the Vernier effect of a single-fiber Sagnac loop and the isopropanol-sensitive material of PPy. The sensing mechanism is based on the monitoring of the strain effect due to the swelling of PPy in presence of different concentrations of isopropanol. The sensor exhibited a high sensitivity of 239 pm/ppm compared to the conventional Sagnac interferometer in the range of 0 ppm to 42 ppm. Later, in 2020, Setka et al. synthesized Cadmium telluride (CdTe)/PPy nanocomposites and integrated them into love wave (L-SAW) sensors to detect acetone at RT [262]. The sensing response was obtained in terms of frequency shift and calculated by monitoring the change in the resonant frequency in presence of air and the target VOC. The fabricated device exhibited a high sensitivity to acetone among ethanol and toluene along with a lower LOD of 5 ppb. In 2021, Shoeb et al. synthesized Gr/Ag–Ag2O/PPy nanocomposites for fast responsive and selective detection of ammonia at RT, as shown in Figure 21 [263]. The sensing response of the device was monitored by measuring the electrical conductivity responses of the device at RT. The composition of Gr/Ag–Ag2O/PPy exhibited a 40 times higher amplitude of conductivity variation compared to pristine PPy with the presence of 1000 ppm of ammonia vapor. The sensor also showed excellent selectivity towards ammonia compared to ethanol, acetaldehyde, formaldehyde, benzene toluene, and m-xylene.
Most recently, Byeon et al. synthesized SWCNT/PPy/phenyllactic acid nanocomposite as shown in Figure 22, to detect acetone vapor at RT [264]. The obtained sensing material had structural features similar to heptadecafluorooctanesulfonic acid (C8F)-doped-PPy layer surrounding a single-stranded SWCNT, and the phenyllactic acid layer surrounding the PPy work as a selective sensing layer for acetone. Here, the PLA bonded chemically with the PPy backbone, exhibited a sensing synergistic effect, and greatly improved the sensitivity to acetone. Finally, C8F-doped-PPy/PLA@SWCNT as an acetone sensor also provided reliable detection at a low concentration of 50 ppb at 25 °C. Additionally, the fabricated sensor was able to detect a clear signal with the humidity varying in the range of 0% to 80% at a low temperature, which is highly required for medical diagnosis during the breath analysis.
Nowadays, there is a sudden growth in PPy-based VOC sensors due their high sensitivity, selectivity, stability, and rapid response at RT. High sensitivity and significant selectivity sensors can be obtained by integrating or doping the PPy with other materials, e.g., SWCNT, CdTe, SnO2, etc. The only challenging task of the PPy-based VOC sensor is the mixing of PPy with other materials which requires good knowledge of chemistry and expertise in the synthesis.

7.2.2. Polythiophene (PTh)

Among the various conducting polymers, polythiophene (PTh), which is a p-type semiconductor, has been exploited very less for vapor detection compared to the other conducting polymers. However, the trend of using PTh nanocomposites in VOC-sensing technology has been rapidly increasing in the last few decades. In 2015, Tripathi et al. developed PTh/aluminium oxide (PTh/Al2O3) nanocomposite to study its potential for ammonia gas sensing [265]. The composite was synthesized in pellet form by chemical oxidation technique. The variations in the resistance of the pellets were studied with various concentrations of ammonia within the range of 25 ppm to 650 ppm in the closed chamber. The report showed that the response can be improved by increasing the concentration of Al2O3 doping as well as increasing the concentration of ammonia. Later in 2016, Bai et al. synthesized hybrid material to detect nitrogen dioxide (NO2) utilizing reduced graphene oxide (rGO) and PTh [266]. Graphene has shown great potential in various applications owing to its large surface area, excellent electron transport capacity, and outstanding mechanical strength which was further enhanced by the hybridization with conducting polymers. The results showed that the hybrid material sensor with 5 wt% rGO showed a four times higher sensitivity to 10 ppm of NO2 gas and excellent selectivity compared to a pure PTh sensor. Husain et al. reported a series of experiments on PTh and its potential in organic vapor sensing [267,268,269,270]. In 2019, they developed PTh/ZrO2 nanocomposite for ethane detection [267]. The synthesized materials were studied for comparative DC electrical conductivity retention under cyclic ageing and isothermal conditions. In this work, the high surface area and efficient conducting network allow the quantitative adsorption and desorption of ethane. Under the ethane gas environment, PTh/ZrO2 nanocomposite sensor exhibited a 19 times higher electrical conductivity, representing faster adsorption and desorption towards ethane compared to a pure PTh sensor due to its high surface area. In 2020, they proposed the utilization of PTh/SnO2 nanocomposite in a pellet-shaped for alcohol sensing including butan-1-ol (1° alcohol), butan-2-ol (2° alcohol), and 2-methyl propanol (3° alcohol) at RT [269]. The PTh/SnO2 pellet exhibited the highest sensing response in terms of the variation in DC electrical conductivity and maximal reproducibility for butan-1-ol. In the same year, Husain et al. developed PTh/GO nanocomposite to detect ethanol vapor at RT [270]. The PTh/GO nanocomposite significantly showed an improvement in DC electrical conductivity for ethanol vapor compared to pure PTh with a low LOD of 400 ppm. From all the research mentioned above, it is clearly observed that an appropriate amount of composite material in PTh may lead to a high sensitivity and selectivity towards the VOC vapors. Although this field is yet to be exploited, it can be widely applied to monitor VOCs at the sub-ppm range.

7.3. Molecularly Imprinted Polymer (MIP)

MIPs are polymeric materials synthesized with nanocavities of similar shape and size to the target molecules (also known as templates). The specificity is a distinctive feature of MIPs that makes them an ideal element in the sensing system. In 1894, an idea of molecular interactions was reported by Emil Fischer using the term “lock-and-key” [271]. The first article on the successful imprinting of amino acids in the polymer was reported by Andersson et al. in 1984 [272]. In 1985, the term ‘imprinted polymer’ was first used in history by G. Wulff [273,274]. He proposed the covalent-binding technique to create the imprinted polymer. The MIPs can be synthesized for almost every kind of target molecule despite its shape, size, and chemical structures. Generally, the synthesis of MIPs follows a similar outline as follows [275] (i) combination of monomers with template (ii) polymerization under UV or hot air (iii) extraction of template from the polymer to create a specific cavity for binding with target molecules. With the ability to do so with every kind of target molecule, MIP plays an outstanding role as an analytical tool in the development of tailor-made sensors and gains huge interest in VOC detection [276,277]. The MIP-based sensors exhibit a full range of properties including high selectivity, sensitivity, great reliability, low cost, and mechanical, thermal, and chemical stability [278].
In 2014, Mustafaa and Lieberzeit reported a MIP/Ag2S nanoparticle composite coated on quartz crystal microbalance (QCM) for aliphatic alcohol detection [279]. Mixing MIP with Ag2S nanoparticles results in a nanocomposite material with the collective properties of both materials. The MIP/Ag2S nanocomposite gave a normalized response of −70 Hz for 1-butanol vapor at 400 ppm, which was two times higher response compared to the pure Ag2S layer of similar thickness. The synergy produced by the strong affinity of Ag2S nanoparticles and the selective ability of MIPs toward 1-butanol vapor allowed fabricated sensors to have a relatively low detection limit (LOD) of 2 ppm. Later in 2016, Hussain et al. reported the QCM sensor coated with MIP for the detection of formaldehyde vapors in air streams [280]. The sensor achieved a low LOD of 500 ppb with good selectivity for formaldehyde compared to methanol, acetaldehyde, formic acid, and dichloromethane. In 2017, Tang et al. proposed another sensor for formaldehyde detection utilizing MIP/TiO2 nanotube array composite, as shown in Figure 23 [281]. In this work, the MIP/TiO2 nanotube array had a high surface-to-volume ratio, resulting in the improvement of sensing performance. As the humidity is neglected, the sensor can selectively detect formaldehyde in the ppm range at RT along in the long term.
In 2019, Abdelghani et al. developed highly sensitive SnO2 nanostructured electrochemical sensors utilizing MIP technique to detect acetone and ammonia, simultaneously [282]. Four SnO2 films were prepared using a spin coater under various conditions of introducing acetone and ammonium hydroxide as template molecules during the hydrothermal synthesis. A 500 nm thicker regular micro sheets have been developed from SnO2 nanoparticles of various sizes ranging from 20 nm to 50 nm (based on the synthesis condition). The sensing response of the device was tested against liquid petroleum gas (LPG), ammonium hydroxide, oxygen, acetone, and benzene. The result showed an increase in response for ammonia and acetone with the increment of the participation of the solvent during the synthesis. The results exhibited a promising response for ammonia gas with a highest sensitivity of 89% at 260 °C while for acetone the sensor exhibited a sensitivity of 77% at 230 °C. In a recent study published in 2022, Pathak et al. reported an optical sensor utilizing a MIP layer on a glass slide to detect isopropanol (IPA) vapor [283]. The work illustrated the surface morphology difference between non-imprinted and imprinted polymers, as shown in Figure 24. The regular texture for non-imprinted polymer appeared because of the absence of the IPA template during the synthesis leading to no specific binding sites for IPA molecules. However, the crack and/or porosity appeared on the imprinted polymer surface due to the successful removal of the IPA templates. The increase in the wrinkled structure and surface roughness in the imprinted polymer can be attributed to the increase in the surface area due to the accommodation of the IPA template. The fabricated device was investigated over IPA solutions with various concentrations varying from 50% to 100%. The sensor exhibited a maximum sensitivity of 0.63 nm/%IPA vapor at 120 min exposure, along with good selectivity among a similar class of VOCs.
In the same year, Mousazadeh et al. reported a chemorestive sensor with the use of Au/MIP nanocomposite to detect hexanol, which is a cancer biomarker [284]. The composite-sensing material was coated on an interdigitated electrode. The variation in electrical response and resistance of the device was analyzed by Fourier transform cyclic voltammetry with the presence of various concentrations of hexanal gas. The sensor exhibited a low LOD of 1.1 ppm for a linear range of hexanal concentration varying from 2.5 ppm to 300 ppm. Additionally, the device showed three times higher sensitivity to hexanal compared to the other VOCs with similar carbon atom numbers, exhibiting a superior selectivity sensor for hexanol.

7.4. Advantages and Limitations of Polymer-Based VOC Sensors

CP-based VOC sensors are nanomaterial-based sensors. This technology is commercialized under the name Cyranose 320, which comprises an array of 32 polymer composite sensors [285]. Although the CPs are highly sensitive to several VOCs along with the rapid response time, it is also sensitive towards humidity and temperature. Therefore, the device suffers from cross-sensitivity leading to an inaccurate readout. Whereas, the imprinted polymer is considered a promising candidate for selective detection applied under a mixed gaseous environment. However, the response time of the MIP sensors is longer due to the difficulty associated with diffusion [286]. The sensitivity and efficiency of MIP-based sensors are directly related to the number of imprinted cavities available on the sensor surface.
In Table 5 we have summarized the various kinds of metal oxide sensors and their sensing performances.

8. Other Sensing Materials

Aside from these renowned sensing materials, some other materials have also been exploited by researchers due to their superior adsorption capability towards the VOC gasses. Some of them are selected to be discussed in the following section due to their excellent performance which can be promising in the VOC-sensing area.

8.1. Metal Organic Frameworks (MOFs)

MOFs have emerged as an excellent sensing material for gas sensing due to its crystalline nature, extreme porosity and high surface areas [305]. These properties allow the MOFs to have the ability to sense both the inorganic and organic compounds. The MOFs used for the adsorption of VOCs are mainly Material of Institute Lavoisier (MILs) series and isoreticular metal-organic frameworks (IRMOFs) series which are discussed in detail in the following subsections. Over the past two decades, the MOFs have been exploited well and currently, there are more than twenty thousand MOFs materials available [306,307,308].

8.1.1. Material of Institute Lavoisier (MIL) Series

MIL-n is a series belonging to the class of materials in MOFs which was first synthesized by Ferey’s group [309]. The synthesized MIL-101 comprised a mesoporous structure with a pore diameter of 2.9 nm to 3.4 nm and an internal surface area of 6000 m2/g [309]. Due to the porous structure of MIL-101, it gains a huge interest in gas-sensing applications [310]. In 2011, Huang et al., firstly, developed a QCM-based VOCs sensor utilizing MIL-101 [311]. They reported that MIL-101 exhibits various affinities for the various VOCs molecules due to the energy heterogeneity of MIL-101. The adsorption isotherms of the VOCs on MIL-101 followed the Dubinin–Astakhov equation with the characteristic energy from 5.70 (methanol) to 9.13 kJ.mol−1 (n-butylamine), Astakhov exponent from 0.50 (n-butylamine) to 3.03 (n-hexane), and the limiting adsorption capacity from 0.08 (n-hexane) to 12.8 (n-butylamine) mmol.g−1. MIL-101 showed the weakest affinity for n-hexane and the strongest affinity to n-butylamine. The dynamic adsorption mechanism mainly relies on the metal sites of MIL-101 which leads to greater VOC adsorption capacity and higher affinity compared to the AC. Most recently in 2022, Mousavi and Zeinali reported VOC detection based on resistive gas nanosensor using MIL-101(Cr)/CNT nanocomposite [312]. The sensor was investigated over ethanol, methanol, isopropanol, formaldehyde, acetonitrile, acetone, tetrahydrofuran, dichloromethane, and n-hexane at RT. The sensor exhibited an excellent selectivity towards formaldehyde along with a low LOD of 0.07 ppm, and high sensitivity of 325.21%. The response time of the sensor was 6 min over a wide VOC concentration range varying from 10 ppm to 1000 ppm. The molecular size, chemical structures, and polarity of the material were the key factors which affect the sensing performance of the device.
Although the MIL-n series shows an excellent performance in gas separation and adsorption, there are still several limitations including poor stability, insufficient storage capacity, and high cost leading to limited research exploitation possibilities. However, the specific surface area, metal sites, pore structure, π-π binding, and modifiable groups are some of the key parameters which can affect the separation and adsorption processes of different VOCs. Currently, researchers are interested in designing and improving the structure of the MOFs materials, in order to develop a new MOFs material to attain specific functions and overcome their limitations.

8.1.2. Isoreticular Metal-Organic Frameworks (IRMOFs) Series

IRMOFs are a series of MOFs that own similar network topologies. Introducing a simple substitution of organic linkers of IRMOF-1 (i.e., MOF-5), other IRMOFs can be produced which shows unique property such as high chemical stability and a large Brunauer, Emmett and Teller (BET) surface areas [313]. In 1995, Yaghi et al. synthesized micropores MOF to selectively bind the organic guest molecules [314]. The building block was a symmetric organic molecule, which selectively binds metal ions in order to form a layer of the metal organic compound. This layer was alternatively formed with the layers whose composition was determined by the functionalization of the starting molecules creating channels which allow the selective binding of guest molecules. The crystal lattice of IRMOFs layers was stable at 350 °C even after the removal of the guest aromatic molecules and it selectively adsorbs gaseous molecules. The stability and rigidity of IRMOFs materials allow the size and shape-selective absorption of organic molecules and ions.
In 2015, Yaghi et al. proposed a high-absorptive material named MOF-5 (IRMOF-1) from benzene and octahedral Zn-O-C clusters links [306]. In this work, the author also showed that the synthesized porous material can be functionalized with organic groups such as NH2, Br, OC3H7, C2H4, OC5H11, and C4H4. In addition, its pore size can be expanded with a long molecular structure such as tetrahydropyrene, terphenyl, pyrene, and biphenyl. In order to improve the adsorption performance, pore texture and crystal structure of MOF-5, Saha et al. utilized diethylformamide as a solvent in the synthesis instead of the conventional solvent [315]. This synthesized MOF-5 comprised a large specific surface area of 3917 m2/g, with a pore size of 8.6 Å and pore volume of 1.39 cm3/g, respectively. Yang et al. developed MOF-177 at 85 °C by using solvent thermal technique and exhibited its superior adsorption capacity for VOC molecules [316]. In the reported work, the specific surface area of MOF-177 was 4170 m2/g, with an average pore size of 0.94 nm, and pore volume of 1.11 cm3/g, respectively. The material exhibited great adsorption towards benzene, toluene, acetone, ethylbenzene, o-xylene, m-xylene, and p-xylene in the air. From the above investigation, it is quite clear that the MOFs show an excellent adsorption capacity due to their large specific surface areas and may become good candidate materials in the VOC sensing.

8.2. Zeolites

Zeolites are micropores crystalline aluminosilicates with a highly ordered porous structure. Zeolites are widely utilized in various sensing applications including heavy metal detection [317], biosensing [318], VOC gas sensing [310] due to their tunable pore size, acceptable thermal stability, high hydrophobicity, and ease of modification of their surface. According to the previous research work, the average specific surface area of zeolites has been determined to be 800 m2/g, which can be maximized up to 1769 m2/g [215].
In 2009, Bhatia et al. reported the adsorption of butyl acetate using silver-exchanged yttrium (AgY (IE)) and silver-loaded zeolites (AgZSM-5 (IE)) at 28 °C [319]. The Ag was loaded in the zeolites via ion exchange and impregnation techniques. The adsorption capacity of AgY (IE) against the butyl acetate was reduced by 42% due to the increase in the RHof the VOCs. However, the adsorption capacity of AgZSM-5 (IE) for butyl acetate was negligibly affected by the humidity, and the adsorption capacity was only reduced by 7%. In 2016, Xiangping et al. utilized the optical fiber as the sensing platform for the detection of VOC traces [320]. In the reported work, the author utilized the feature of nanoporous zeolites for the selective adsorption of isopropanol. The device comprised of the spherical-shaped fiber coated with thin zeolite film at the edge, as shown in Figure 25a–d. The zeolite layer and spherical shaped fiber formed an arc-shaped inline FP cavity, which improved the sensing performance, shown in Figure 25e,f. The proposed sensor exhibited enhanced sensitivity of 0.91 nm/ppm for IPA concentrations ranging from 0 ppm to 70 ppm.
An electrochemical sensor developed from SnO2/ZSM-5 zeolite for the detection of formaldehyde was reported in 2021 by Sun et al. [321]. The composite material was highly selective and sensitive towards the formaldehyde due to the large number of oxygen vacancies existing on the surface of the composite allowing the rapid adsorption of formaldehyde. The result obtained from SnO2/ZSM-5 composite had a better sensing performance compared to the pure SnO2. This improvement was attributed to the composite structure by integrating ZSM-5 zeolite.
Zeolites are generally used in commercial separation processes and VOCs recovery due to their molecular sieving ability to preferentially adsorb certain molecules and to exclude others [322]. The size and the structure of both the micropores and the molecular adsorbate determine the selectivity and adsorption capacity. The sorbed VOC molecules may interact with (a). the zeolite lattice, (b). extra-framework cations, and (c). other sorbed molecules. The polarity of the probe molecule, its dimensions, and its partial pressure define the kind and the magnitude of interaction with a given adsorbent. On the other hand, for a given adsorbate, the number and the place of the free cations, the topology, the size of the channels, and the possible pre-adsorption of other molecules (e.g., humidity) also affect this adsorbate/adsorbent interaction. MCM-41 and SBA-15 are the types of zeolites which are widely used in VOC detection [323,324].

9. Future Prospects

This review highlights the potential use of active materials including metal oxides, polymers, conductive polymers, carbon-based materials, composites, and other adsorption materials as a VOC biomarker sensor due to their excellent sensing property and wide applicability. These materials are representatives for sensing clinical VOC biomarkers. The enhanced properties of these materials such as sensitivity, thermal stability, conductivity, selectivity, and fast responses generally include the control of their composition, size, and morphology. Some of the modified structures of these materials such as hierarchical structures, bilayer structure formation by the introduction of tunable attractive elements, such as by doping, have displayed an efficient improvement in VOC-sensing performances. VOC has increasingly affected the ecological environment and human health due to its high toxicity, carcinogenic, and photocatalytic in nature. Therefore, there is a need for their precise detection, monitoring, and controlling adverse effects of these VOCs effectively by combing the aforementioned materials in real time. Metal oxides such as ZnO, NiO, TiO2, and MoO3 offer admissible responses to a different variety of VOCs due to their nature of absorbing gases/vapors and converting them into conductivity variations. Moreover, the usage of carbon-based materials, polymers, and other adsorption materials also have similar properties to metal oxides such as charge transfer between the gases/vapor species, oxidation, and the reduction of the dopants (n- or p-type) which are widely accepted as a promising candidate in the VOC-sensing applications. Their different modified morphologies at the nanometric scale with the introduction of different tunable elements have attracted the attention of many researchers because of their increase in a surface-to-volume ratio that gives rise to high sensitivity, fast response speed, lower operating temperatures, and good stabilities. However, the advances in terms of selectivity in a mixed volatile environment still remain to be a challenging task.
Although outstanding advantages were achieved by implementing those materials into the VOC sensors, there are still major challenges involved in the clinical diagnosis of the VOC biomarkers. The thorough investigation in the functionality of materials to a larger group of VOCs as well as the implementation of other sensing mechanisms such as optical, electrical, and mechanical-based gas/vapor sensors are in high demand. In terms of materials, the accurate detection, and the sensitivity for the low concentrations (ppb-ppt levels) of the VOC biomarkers is a prime concern. Therefore, there is a strong need in exploring and utilizing new hybrid materials, i.e., ( i ) by the combination of metal oxide and carbon-based materials, ( ii ) metal oxide and polymers, and ( iii ) carbon-based materials and polymers towards VOC detection to overcome the issue of poor reproducibility in routine use. It is also worth mentioning that by combining varieties of these materials, it is expected to provide promising results when operating them together, thus offering strong stability in abrasive variations of humidity and/or temperature.
Further, numerous reports can be found on the use of doping and composite materials for tailoring the sensor responses towards specific VOCs. However, valid information on efficacy and the role of such materials of VOC detection are limited. Moreover, a large number of reports are available for chemoresistive sensors that work under the resistive principle but not on electrical, optical, and mechanical detection of VOCs. Thus, in the future, the properties of materials for VOC detection should be explored more by utilizing the various physical changes. As these physical changes in the properties of materials may have the possibility to offer better selectivity for VOC-sensing devices.
In summary, it can be seen that various challenges and concerns associated with VOC sensors are mostly based on the type of materials synthesis method discussed in this review. From the materials discussed earlier, incorporating them into developing gas sensors for the detection of associated clinical VOC biomarkers is known to influence its sensing performances. These materials are therefore gaining a huge interest in designing clinical and diagnostic tools.
Thus, the clinical use of the VOCs from the human exhaled breath can be implemented by developing a versatile breath gas sensor with specific requirements depending on the applications to achieve a high response value, exceptional selectivity, lower LOD, great stability/reproducibility, stable calibration of the sensor system upon malfunction and high response/recovery speeds with a lower detection limit in real time. Moreover, from the economic point of view, these gas sensors based on the materials discussed should be universally available and should provide ease of fabrication, wider compatibility, and simple and low cost for easy implementation in different transducing platforms.

10. Clinical Challenges

Human breath analysis is an interdisciplinary field of research work which includes analytical techniques, data processing, medical science, materials chemistry, and engineering. Health monitoring and disease detection via blood analysis are invasive techniques which are painful to the patient. In the past decades, exhaled breath analysis has emerged as a better alternative to diagnose the disease at an early stage because of its non-invasive nature. However, the method is fraught with several challenges. Direct exhaled breath cannot be used in all breath analysis methods. Breath collection and its storage became a major issue. The collected breath which is stored for a long time is often degraded and hence, their original compositions are changed. Researchers are still not sure whether mouth-exhaled or nose-exhaled breath should be utilized for the VOC analysis. Currently, it is advised that dead space air should not be considered for breath analysis; only end-tidal breath should be used. However, for highly water-soluble breath biomarkers, such as isoprene and acetone, there is no anatomic dead space to be defined. Most of the sensor responses are dependent on the exhaled breath flowrate onto the sensor head. In the direct breath analysis, patients are allowed to directly exhale onto the sensing head. However, different patients would exhale at different flow-speeds, leading to inconsistent measurements. The exhaled breath concentration of various VOC gases varies significantly according to the patient’s gender, age, weight, race, food habits, pregnancy, lifestyle, etc. In addition, not all exhaled breath components are endogenous; rather, most of them are exogenous in origin. Many of the endogenous gases are not even systemic in origin. Therefore, locality becomes another confounding factor. In this section, a comparison of the performance is drawn between the gold-standard devices and other devices based on non-biological sensing materials. The selective detection of diseases from exhaled breath is limited to a small number of diseases, e.g., chronic obstructive pulmonary disease, diabetes, etc., which can be monitored using a single biomarker. In most diseases, especially various cancers, the concentrations of breath VOCs varied simultaneously. Therefore, the use of target-specific detection demands the development of highly selective sensing materials which is a paramount challenge in itself. Other sensors with poor selectivity can suffer problems from low to medium sensitivity. Therefore, obtaining sufficient discrimination between diseased and healthy groups might become difficult.
Gold standards: GC–MS, SIFT-MS, and PTR-MS are the most accurate detection techniques for exhaled breath analysis with an analysis time of millisecond. These instruments require a fraction of ml of liquid or gas samples for monitoring. These classical instruments are expensive and require a trained person for handling, data analysis, and data interpretation. These instruments are not like any household devices and can only be available in hospitals and diagnostic clinics. This impedes the use of these instruments as breath analyzers on a day-to-day basis at our homes. One of the major goals of breath analysis is early detection of disease by regular health monitoring and for that to be possible the device used must be financially affordable for the common man, easy to use even for a layman and handheld.
MO sensing materials: MO-based VOC sensors are cost-effective, easy to use, compact and capable of detecting several VOC biomarkers at ppm, ppb, and even, at ppt levels. Although most MO sensors exhibit excellent performance, a lack of specificity limits them to be used as a medical diagnostic tool. MO-based VOC sensors are also sensitive to humidity. Human breath contains almost saturated moisture which acts as a major cross-sensitive agent impeding the response of the sensor to target VOC molecules. Currently, thick film MO sensors are available in the market. The thick MO film-based sensors suffer high power consumption and poor sensitivity due to a low diffusion rate. MEMS-based sensor arrays can resolve the issue of power and size to some extent; however, they comprise other problems. The microsystems generally utilize moisture traps, which adsorb moisture along with the target gases. Humidity dependence of gas adsorption can significantly reduce the reliability and reproducibility of the pre-concentrator. Pre-concentration requires longer time; thereby, making real-time analysis difficult. Further, the small volume of pre-concentrator materials present in such microsystems might not be sufficient for the detection of trace gases. Selective and reversible pre-concentration remain challenges.
Carbon-based VOC sensors: By coupling MO, e.g., ZnO, NiO, MoO3, and TiO2 with innovative carbon-based sensing materials such as GO, rGO, MWCNTs, AC, etc., the performance of the device can be improved in terms of sensitivity, LOD, response and recovery times towards the specific gaseous targets. Although most of the sensors show excellent performance, many of them fail to fulfill the requirement of medical diagnosis due to poor selective performance in a mixed gaseous environment. Even though GO, CNT, AC, etc., are considered novel VOC-sensing materials, they still require further study and improvement for commercialization.
Polymer-based VOC sensors: Polymer-based VOC sensors are considered selective tools for breath biomarkers detection. Among various polymers, the MIP can be considered an excellent material to selectively detect VOC in very low-level concentrations of ppm/ppb. Although the MIP-based sensor provides a good selective response towards the target gas it suffers from the issue of rapid response time. The field is still growing; therefore, the commercialization of polymer-coated VOC sensors has not been explored yet.

11. Conclusions

It has been reported previously that each human generates unique VOC profiles in their exhaled breath which can be utilized as biomarkers in order to monitor and diagnose various disease conditions. In the past decades, several sensing materials have been reported to achieve high selectivity and sensitivity in the sub-ppm range. In this review, a detailed literature survey of the recent progress of VOCs sensing materials and the key factors to improve the sensing response and to make it reliable for real-time utilization are provided. Each sensing material comprises some advantages and limitations which are briefly outlined at the end of each subheading. The new idea and approaches to improve the sensing performance of materials concerning adhesion, stability, and selectivity are also suggested.

Author Contributions

Conceptualization, C.V. and A.K.P.; methodology, C.V., A.K.P. and K.S.; formal analysis, A.K.P. and K.S.; writing original draft preparation, A.K.P., K.S., N.K., P.J., N.A. and J.U.; writing—review and editing, C.V., A.K.P. and K.S.; supervision, C.V.; project administration, C.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Second Century Fund (C2F), Chulalongkorn University, and Rachadapisek Sompote Fund for Intelligent Control Automation of Process Systems Research Unit, Chulalongkorn University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Minh, T.D.C.; Blake, D.R.; Galassetti, P.R. The clinical potential of exhaled breath analysis for diabetes mellitus. Diabetes Res. Clin. Pract. 2012, 97, 195–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Phillips, M.; Herrera, J.; Krishnan, S.; Zain, M.; Greenberg, J.; Cataneo, R.N. Variation in volatile organic compounds in the breath of normal humans. J. Chromatogr. B Biomed. Sci. Appl. 1999, 729, 75–88. [Google Scholar] [CrossRef] [PubMed]
  3. Miekisch, W.; Schubert, J.K.; Noeldge-Schomburg, G.F. Diagnostic potential of breath analysis—Focus on volatile organic compounds. Clin. Chim. Acta 2004, 347, 25–39. [Google Scholar] [CrossRef]
  4. Zhou, J.; Al Husseini, D.; Li, J.; Lin, Z.; Sukhishvili, S.; Coté, G.L.; Gutierrez-Osuna, R.; Lin, P.T. Detection of volatile organic compounds using mid-infrared silicon nitride waveguide sensors. Sci. Rep. 2022, 12, 5572. [Google Scholar] [CrossRef] [PubMed]
  5. Pathak, A.K.; Viphavakit, C. VOC Biomarker Monitoring for Diabetes through Exhaled Breath Using Ag/P-TiO2 Composite Plasmonic Sensor. IEEE Sens. J. 2021, 21, 22631–22637. [Google Scholar] [CrossRef]
  6. He, J.; Zheng, F.; Zhou, Y.; Li, X.; Wang, Y.; Xiao, J.; Li, Y.; Chen, D.; Lu, J. Catalytic oxidation of VOCs over 3D@2D Pd/CoMn2O4 nanosheets supported on hollow Al2O3 microspheres. J. Colloid Interface Sci. 2022, 613, 155–167. [Google Scholar] [CrossRef] [PubMed]
  7. Kaluza, D.; Węgrzyn, K.; Kaczmarczyk, B.; Gniadek, M.; Maksymiuk, K.; Michalska, A. An Electrochemical Approach to Quantification of Volatile Organic Solvents Dispersed in Solution—Towards Bipolar Electrode Sensors. Electroanalysis 2022, 34, 25–32. [Google Scholar] [CrossRef]
  8. Qu, C.; Wang, S.; Liu, L.; Bai, Y.; Li, L.; Sun, F.; Hao, M.; Li, T.; Lu, Q.; Li, L.; et al. Bioinspired Flexible Volatile Organic Compounds Sensor Based on Dynamic Surface Wrinkling with Dual-Signal Response. Small 2019, 15, 1900216. [Google Scholar] [CrossRef] [PubMed]
  9. Huang, Y.; Lin, W.; Huang, T.; Li, Z.; Zhang, Z.; Xiao, R.; Yang, X.; Lian, S.; Pan, J.; Ma, J.; et al. Ultrafast Response Optical Microfiber Interferometric VOC Sensor Based on Evanescent Field Interaction with ZIF-8/Graphene Oxide Nanocoating. Adv. Opt. Mater. 2022, 10, 2101561. [Google Scholar] [CrossRef]
  10. Tomić, M.; Šetka, M.; Vojkůvka, L.; Vallejos, S. VOCs Sensing by Metal Oxides, Conductive Polymers, and Carbon-Based Materials. Nanomaterials 2021, 11, 552. [Google Scholar] [CrossRef]
  11. Tricoli, A.; Righettoni, M.; Teleki, A. Semiconductor Gas Sensors: Dry Synthesis and Application. Angew. Chem. Int. Ed. 2010, 49, 7632–7659. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, Z.; Bu, M.; Hu, N.; Zhao, L. An overview on room-temperature chemiresistor gas sensors based on 2D materials: Research status and challenge. Compos. Part B Eng. 2023, 248, 110378. [Google Scholar] [CrossRef]
  13. Abokifa, A.A.; Haddad, K.; Fortner, J.; Lo, C.S.; Biswas, P. Sensing mechanism of ethanol and acetone at room temperature by SnO2 nano-columns synthesized by aerosol routes: Theoretical calculations compared to experimental results. J. Mater. Chem. A 2018, 6, 2053–2066. [Google Scholar] [CrossRef]
  14. Elger, A.; Hess, C. Elucidating the Mechanism of Working SnO2 Gas Sensors Using Combined Operando UV/Vis, Raman, and IR Spectroscopy. Angew. Chem. Int. Ed. 2019, 58, 15057–15061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Ruzsanyi, V.; Lochmann, F.; Jürschik, S.; Mochalski, P.; Unterkofler, K.; Mayhew, C.A. Chapter 1. Origin and Emission of Volatile Biomarkers in Breath: End-tidal Perspective. In Volatile Biomarkers for Human Health; Royal Society of Chemistry: Cambridge, UK, 2022; pp. 1–21. [Google Scholar]
  16. Available online: https://www.owlstonemedical.com/science-technology/voc-biomarkers/ (accessed on 16 December 2022).
  17. Pathak, A.K.; Viphavakit, C. A Review on All-optical Fiber-based VOC sensors: Heading towards the Development of Promising Technology. Sens. Actuators A Phys. 2022, 338, 113455. [Google Scholar] [CrossRef]
  18. Cuiting, Y.; Miao, G.; Pi, Y.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Abatement of Various Types of VOCs by Adsorption/Catalytic Oxidation: A review. J. Chem. Eng. 2019, 370, 1128–1153. [Google Scholar] [CrossRef]
  19. McLawrence, L.T.; Patrick, R.; Silke, B. Breath isoprene in patients with heart failure. Eur. J. Heart Fail. 2001, 3, 423–427. [Google Scholar] [CrossRef] [PubMed]
  20. Bajtarevic, A.; Ager, C.; Pienz, M.; Klieber, M.; Schwarz, K.; Ligor, M.; Ligor, T.; Filipiak, W.; Denz, H.; Fiegl, M.; et al. Noninvasive detection of lung cancer by analysis of exhaled breath. BMC Cancer 2009, 9, 348. [Google Scholar] [CrossRef] [Green Version]
  21. Trefz, P.; Obermeier, J.; Lehbrink, R.; Schubert, J.K.; Miekisch, W.; Fischer, D.-C. Exhaled volatile substances in children suffering from type 1 diabetes mellitus: Results from a cross-sectional study. Sci. Rep. 2019, 9, 15707. [Google Scholar] [CrossRef] [Green Version]
  22. Crofford, O.B.; Mallard, R.E.; Winton, R.E.; Rogers, N.L.; Jackson, J.C.; Keller, U. Acetone in breath and blood. Trans. Am. Clin. Climatol. Assoc. 1977, 88, 128–139. [Google Scholar]
  23. Fan, X.; Xu, Y.; Ma, C.; He, W. In-situ growth of Co3O4 nanoparticles based on electrospray for an acetone gas sensor. J. Alloys Compd. 2021, 854, 157234. [Google Scholar] [CrossRef]
  24. Fearing, M.K.; Shih, V.E. Urea Cycle, Inborn Defects of. In Encyclopedia of Biological Chemistry; Elsevier: Amsterdam, The Netherlands, 2004; pp. 326–330. [Google Scholar]
  25. Essiet, I.O. Diagnosis of kidney failure by analysis of the concentration of ammonia in exhaled human breath. J. Emerg. Trends Eng. Appl. Sci. 2013, 4, 859. [Google Scholar]
  26. Butterworth, R.F. Hepatic encephalopathy. Alcohol Res. Health 2003, 27, 240. [Google Scholar] [PubMed]
  27. Chaudhary, P.P.; Conway, P.L.; Schlundt, J. Methanogens in humans: Potentially beneficial or harmful for health. Appl. Microbiol. Biotechnol. 2018, 102, 3095–3104. [Google Scholar] [CrossRef] [PubMed]
  28. Polag, D.; Keppler, F. Global methane emissions from the human body: Past, present and future. Atmos. Environ. 2019, 214, 116823. [Google Scholar] [CrossRef]
  29. Dharmawardana, N.; Goddard, T.; Woods, C.; Watson, D.I.; Butler, R.; Ooi, E.H.; Yazbeck, R. Breath methane to hydrogen ratio as a surrogate marker of intestinal dysbiosis in head and neck cancer. Sci. Rep. 2020, 10, 15010. [Google Scholar] [CrossRef]
  30. Ayala, A.; Muñoz, M.F.; Argüelles, S. Lipid Peroxidation: Production, Metabolism, and Signaling Mechanisms of Malondialdehyde and 4-Hydroxy-2-Nonenal. Oxid. Med. Cell. Longev. 2014, 2014, 360438. [Google Scholar] [CrossRef] [Green Version]
  31. Pizzimenti, S.; Ciamporcero, E.; Daga, M.; Pettazzoni, P.; Arcaro, A.; Cetrangolo, G.; Minelli, R.; Dianzani, C.; Lepore, A.; Gentile, F.; et al. Interaction of aldehydes derived from lipid peroxidation and membrane proteins. Front. Physiol. 2013, 4, 242. [Google Scholar] [CrossRef] [Green Version]
  32. Skoumalová, A.; Hort, J. Blood markers of oxidative stress in Alzheimer’s disease. J. Cell. Mol. Med. 2012, 16, 2291–2300. [Google Scholar] [CrossRef]
  33. Wang, R. Two’s company, three’s a crowd: Can H2S be the third endogenous gaseous transmitter? FASEB J. 2002, 16, 1792–1798. [Google Scholar] [CrossRef] [Green Version]
  34. Zaichko, N.V. Hydrogen sulfide: Metabolism, biological and medical role. Ukr. Biochem. J. 2014, 86, 5–25. [Google Scholar] [CrossRef] [PubMed]
  35. Toombs, C.F.; Insko, M.A.; Wintner, E.A.; Deckwerth, T.L.; Usansky, H.; Jamil, K.; Goldstein, B.; Cooreman, M.; Szabo, C. Detection of exhaled hydrogen sulphide gas in healthy human volunteers during intravenous administration of sodium sulphide. Br. J. Clin. Pharmacol. 2010, 69, 626–636. [Google Scholar] [CrossRef] [PubMed]
  36. Rosenberg, M.; McCulloch, C.A.G. Measurement of Oral Malodor: Current Methods and Future Prospects. J. Periodontol. 1992, 63, 776–782. [Google Scholar] [CrossRef] [PubMed]
  37. Suarez, F.L.; Furne, J.K.; Springfield, J.; Levitt, M.D. Morning Breath Odor: Influence of Treatments on Sulfur Gases. J. Dent. Res. 2000, 79, 1773–1777. [Google Scholar] [CrossRef] [PubMed]
  38. Kushch, I.; Arendacká, B.; Štolc, S.; Mochalski, P.; Filipiak, W.; Schwarz, K.; Schwentner, L.; Schmid, A.; Dzien, A.; Lechleitner, M.; et al. Breath isoprene—Aspects of normal physiology related to age, gender and cholesterol profile as determined in a proton transfer reaction mass spectrometry study. Clin. Chem. Lab. Med. 2008, 46, 1011–1018. [Google Scholar] [CrossRef] [PubMed]
  39. Deng, C.; Zhang, J.; Yu, X.; Zhang, W.; Zhang, X. Determination of acetone in human breath by gas chromatography–mass spectrometry and solid-phase microextraction with on-fiber derivatization. J. Chromatogr. B 2004, 810, 269–275. [Google Scholar] [CrossRef]
  40. Kearney, D.J.; Hubbard, T.; Putnam, D. Breath Ammonia Measurement in Helicobacter pylori Infection. Dig. Dis. Sci. 2002, 47, 2523–2530. [Google Scholar] [CrossRef]
  41. Velusamy, P.; Su, C.-H.; Ramasamy, P.; Arun, V.; Rajnish, N.; Raman, P.; Baskaralingam, V.; Senthil Kumar, S.M.; Gopinath, S.C.B. Volatile Organic Compounds as Potential Biomarkers for Noninvasive Disease Detection by Nanosensors: A Comprehensive Review. Crit. Rev. Anal. Chem. 2022, 1–12. [Google Scholar] [CrossRef]
  42. Rezaie, A.; Buresi, M.; Lembo, A.; Lin, H.; McCallum, R.; Rao, S.; Schmulson, M.; Valdovinos, M.; Zakko, S.; Pimentel, M. Hydrogen and Methane-Based Breath Testing in Gastrointestinal Disorders: The North American Consensus. Am. J. Gastroenterol. 2017, 112, 775–784. [Google Scholar] [CrossRef] [Green Version]
  43. Müller-Wirtz, L.M.; Kiefer, D.; Ruffing, S.; Brausch, T.; Hüppe, T.; Sessler, D.I.; Volk, T.; Fink, T.; Kreuer, S.; Maurer, F. Quantification of Volatile Aldehydes Deriving from In Vitro Lipid Peroxidation in the Breath of Ventilated Patients. Molecules 2021, 26, 3089. [Google Scholar] [CrossRef]
  44. Das, S.; Pal, M. Review—Non-Invasive Monitoring of Human Health by Exhaled Breath Analysis: A Comprehensive Review. J. Electrochem. Soc. 2020, 167, 037562. [Google Scholar] [CrossRef]
  45. Kistenev, Y.V.; Karapuzikov, A.I.; Kostyukova, N.Y.; Starikova, M.K.; Boyko, A.A.; Bukreeva, E.B.; Bulanova, A.A.; Kolker, D.B.; Kuzmin, D.A.; Zenov, K.G.; et al. Screening of patients with bronchopulmonary diseases using methods of infrared laser photoacoustic spectroscopy and principal component analysis. J. Biomed. Opt. 2015, 20, 065001. [Google Scholar] [CrossRef] [PubMed]
  46. Kistenev, Y.V.; Bukreeva, E.B.; Bulanova, A.A.; Kuzmin, D.A.; Tuzikov, S.A.; Yumov, E.L. Laser spectroscopy and chemometric study of the specific features of air exhaled by patients with lung cancer and chronic obstructive pulmonary disease. Phys. Wave Phenom. 2014, 22, 210–215. [Google Scholar] [CrossRef]
  47. Selvaraj, R.; Vasa, N.J.; Nagendra, S.M.S.; Mizaikoff, B. Advances in Mid-Infrared Spectroscopy-Based Sensing Techniques for Exhaled Breath Diagnostics. Molecules 2020, 25, 2227. [Google Scholar] [CrossRef] [PubMed]
  48. Haines, A.; Dilawari, J.; Metz, G.; Blendis, L.; Wiggins, H. Breath-Methane in Patients with Cancer of the Large Bowel. Lancet 1977, 310, 481–483. [Google Scholar] [CrossRef]
  49. Dryahina, K.; Smith, D.; Bortlík, M.; Machková, N.; Lukáš, M.; Španěl, P. Pentane and other volatile organic compounds, including carboxylic acids, in the exhaled breath of patients with Crohn’s disease and ulcerative colitis. J. Breath Res. 2017, 12, 016002. [Google Scholar] [CrossRef]
  50. Vasthare, R.; Kumar, S.; Arron, L.R. Carbon monoxide breath analyzers and its role in tobacco cessation: A narrative review of literature. J. Int. Oral Health 2018, 10, 71. [Google Scholar] [CrossRef]
  51. Ryter, S.W.; Choi, A.M.K. Carbon monoxide in exhaled breath testing and therapeutics. J. Breath Res. 2013, 7, 017111. [Google Scholar] [CrossRef] [Green Version]
  52. Carpagnano, G.E.; Kharitonov, S.A.; Resta, O.; Foschino-Barbaro, M.P.; Gramiccioni, E.; Barnes, P.J. Increased 8-Isoprostane and Interleukin-6 in Breath Condensate of Obstructive Sleep Apnea Patients. Chest 2002, 122, 1162–1167. [Google Scholar] [CrossRef]
  53. Carpagnano, G.E.; Resta, O.; Foschino-Barbaro, M.P.; Gramiccioni, E.; Carpagnano, F. Interleukin-6 is Increased in Breath Condensate of Patients with Non-Small Cell Lung Cancer. Int. J. Biol. Markers 2002, 17, 141–145. [Google Scholar] [CrossRef]
  54. Sanchez, J.M.; Sacks, R.D. GC Analysis of Human Breath with A Series-Coupled Column Ensemble and a Multibed Sorption Trap. Anal. Chem. 2003, 75, 2231–2236. [Google Scholar] [CrossRef] [PubMed]
  55. Lord, H.; Yu, Y.; Segal, A.; Pawliszyn, J. Breath Analysis and Monitoring by Membrane Extraction with Sorbent Interface. Anal. Chem. 2002, 74, 5650–5657. [Google Scholar] [CrossRef]
  56. Giardina, M.; Olesik, S.V. Application of Low-Temperature Glassy Carbon-Coated Macrofibers for Solid-Phase Microextraction Analysis of Simulated Breath Volatiles. Anal. Chem. 2003, 75, 1604–1614. [Google Scholar] [CrossRef] [PubMed]
  57. Schnabel, R.; Fijten, R.; Smolinska, A.; Dallinga, J.; Boumans, M.-L.; Stobberingh, E.; Boots, A.; Roekaerts, P.; Bergmans, D.; van Schooten, F.J. Analysis of volatile organic compounds in exhaled breath to diagnose ventilator-associated pneumonia. Sci. Rep. 2015, 5, 17179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Durán-Acevedo, C.M.; Jaimes-Mogollón, A.L.; Gualdrón-Guerrero, O.E.; Welearegay, T.G.; Martinez-Marín, J.D.; Caceres-Tarazona, J.M.; Sánchez-Acevedo, Z.C.; de Jesus Beleño-Saenz, K.; Cindemir, U.; Österlund, L.; et al. Exhaled breath analysis for gastric cancer diagnosis in Colombian patients. Oncotarget 2018, 9, 28805–28817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Adams, N.G.; Smith, D. The selected ion flow tube (SIFT); A technique for studying ion-neutral reactions. Int. J. Mass Spectrom. Ion Phys. 1976, 21, 349–359. [Google Scholar] [CrossRef]
  60. Smith, D.; Spanel, P. The Novel Selected-ion Flow Tube Approach to Trace Gas Analysis of Air and Breath. Rapid Commun. Mass Spectrom. 1996, 10, 1183–1198. [Google Scholar] [CrossRef]
  61. Španěl, P.; Davies, S.; Smith, D. Quantification of breath isoprene using the selected ion flow tube mass spectrometric analytical method. Rapid Commun. Mass Spectrom. 1999, 13, 1733–1738. [Google Scholar] [CrossRef]
  62. Diskin, A.M.; Španěl, P.; Smith, D. Time variation of ammonia, acetone, isoprene and ethanol in breath: A quantitative SIFT-MS study over 30 days. Physiol. Meas. 2003, 24, 107–119. [Google Scholar] [CrossRef]
  63. Abbott, S.M.; Elder, J.B.; Španěl, P.; Smith, D. Quantification of acetonitrile in exhaled breath and urinary headspace using selected ion flow tube mass spectrometry. Int. J. Mass Spectrom. 2003, 228, 655–665. [Google Scholar] [CrossRef]
  64. Walton, C.; Patel, M.; Pitts, D.; Knight, P.; Hoashi, S.; Evans, M.; Turner, C. The use of a portable breath analysis device in monitoring type 1 diabetes patients in a hypoglycaemic clamp: Validation with SIFT-MS data. J. Breath Res. 2014, 8, 037108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Alkhouri, N.; Cikach, F.; Eng, K.; Moses, J.; Patel, N.; Yan, C.; Hanouneh, I.; Grove, D.; Lopez, R.; Dweik, R. Analysis of breath volatile organic compounds as a noninvasive tool to diagnose nonalcoholic fatty liver disease in children. Eur. J. Gastroenterol. Hepatol. 2014, 26, 82–87. [Google Scholar] [CrossRef] [PubMed]
  66. Samara, M.A.; Tang, W.H.W.; Cikach, F.; Gul, Z.; Tranchito, L.; Paschke, K.M.; Viterna, J.; Wu, Y.; Laskowski, D.; Dweik, R.A. Single Exhaled Breath Metabolomic Analysis Identifies Unique Breathprint in Patients with Acute Decompensated Heart Failure. J. Am. Coll. Cardiol. 2013, 61, 1463–1464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Hansel, A.; Jordan, A.; Holzinger, R.; Prazeller, P.; Vogel, W.; Lindinger, W. Proton transfer reaction mass spectrometry: On-line trace gas analysis at the ppb level. Int. J. Mass Spectrom. Ion Process. 1995, 149–150, 609–619. [Google Scholar] [CrossRef]
  68. Amann, A.; Poupart, G.; Telser, S.; Ledochowski, M.; Schmid, A.; Mechtcheriakov, S. Applications of breath gas analysis in medicine. Int. J. Mass Spectrom. 2004, 239, 227–233. [Google Scholar] [CrossRef]
  69. Karl, T.; Prazeller, P.; Mayr, D.; Jordan, A.; Rieder, J.; Fall, R.; Lindinger, W. Human breath isoprene and its relation to blood cholesterol levels: New measurements and modeling. J. Appl. Physiol. 2001, 91, 762–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Schmutzhard, J.; Rieder, J.; Deibl, M.; Schwentner, I.M.; Schmid, S.; Lirk, P.; Abraham, I.; Gunkel, A.R. Pilot study: Volatile organic compounds as a diagnostic marker for head and neck tumors. Head Neck 2008, 30, 743–749. [Google Scholar] [CrossRef]
  71. Lukman Hekiem, N.L.; Md Ralib, A.A.; bt Mat Hattar, M.A.; Ahmad, F.B.; Nordin, A.N.; Rahim, R.A.; Za’bah, N.F. Advanced vapour sensing materials: Existing and latent to acoustic wave sensors for VOCs detection as the potential exhaled breath biomarkers for lung cancer. Sens. Actuators A Phys. 2021, 329, 112792. [Google Scholar] [CrossRef]
  72. Yang, X.; Salles, V.; Kaneti, Y.V.; Liu, M.; Maillard, M.; Journet, C.; Jiang, X.; Brioude, A. Fabrication of highly sensitive gas sensor based on Au functionalized WO3 composite nanofibers by electrospinning. Sens. Actuators B Chem. 2015, 220, 1112–1119. [Google Scholar] [CrossRef]
  73. Mirzaei, A.; Leonardi, S.G.; Neri, G. Detection of hazardous volatile organic compounds (VOCs) by metal oxide nanostructures-based gas sensors: A review. Ceram. Int. 2016, 42, 15119–15141. [Google Scholar] [CrossRef]
  74. Zhu, L.; Zeng, W.; Ye, H.; Li, Y. Volatile organic compound sensing based on coral rock-like ZnO. Mater. Res. Bull. 2018, 100, 259–264. [Google Scholar] [CrossRef]
  75. Al-Hardan, N.H.; Abdullah, M.J.; Abdul Aziz, A.; Ahmad, H.; Low, L.Y. ZnO thin films for VOC sensing applications. Vacuum 2010, 85, 101–106. [Google Scholar] [CrossRef]
  76. 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. [Google Scholar] [CrossRef]
  77. Pathak, A.K.; Chaudhary, D.K.; Singh, V.K. Broad range and highly sensitive optical pH sensor based on Hierarchical ZnO microflowers over tapered silica fiber. Sens. Actuators A Phys. 2018, 280, 399–405. [Google Scholar] [CrossRef]
  78. Swargiary, K.; Metem, P.; Kulatumyotin, C.; Thaneerat, S.; Ajchareeyasoontorn, N.; Jitpratak, P.; Bora, T.; Mohammed, W.S.; Dutta, J.; Viphavakit, C. ZnO Nanorods Coated Single-Mode–Multimode–Single-Mode Optical Fiber Sensor for VOC Biomarker Detection. Sensors 2022, 22, 6273. [Google Scholar] [CrossRef]
  79. Teimoori, F.; Khojier, K.; Dehnavi, N.Z. Investigation of sensitivity and selectivity of ZnO thin film to volatile organic compounds. J. Theor. Appl. Phys. 2017, 11, 157–163. [Google Scholar] [CrossRef] [Green Version]
  80. Peng, C.; Guo, J.; Yang, W.; Shi, C.; Liu, M.; Zheng, Y.; Xu, J.; Chen, P.; Huang, T.; Yang, Y. Synthesis of three-dimensional flower-like hierarchical ZnO nanostructure and its enhanced acetone gas sensing properties. J. Alloys Compd. 2016, 654, 371–378. [Google Scholar] [CrossRef]
  81. Renganathan, B.; Sastikumar, D.; Gobi, G.; Rajeswari Yogamalar, N.; Chandra Bose, A. Nanocrystalline ZnO coated fiber optic sensor for ammonia gas detection. Opt. Laser Technol. 2011, 43, 1398–1404. [Google Scholar] [CrossRef]
  82. Muthukrishnan, K.; Vanaraja, M.; Boomadevi, S.; Karn, R.K.; Singh, V.; Singh, P.K.; Pandiyan, K. Studies on acetone sensing characteristics of ZnO thin film prepared by sol–gel dip coating. J. Alloys Compd. 2016, 673, 138–143. [Google Scholar] [CrossRef]
  83. Hosseini, Z.S.; Mortezaali, A.; Iraji zad, A.; Fardindoost, S. Sensitive and selective room temperature H2S gas sensor based on Au sensitized vertical ZnO nanorods with flower-like structures. J. Alloys Compd. 2015, 628, 222–229. [Google Scholar] [CrossRef]
  84. Bora, T.; Fallah, H.; Chaudhari, M.; Apiwattanadej, T.; Harun, S.W.; Mohammed, W.S.; Dutta, J. Controlled side coupling of light to cladding mode of ZnO nanorod coated optical fibers and its implications for chemical vapor sensing. Sens. Actuators B Chem. 2014, 202, 543–550. [Google Scholar] [CrossRef]
  85. Wang, J.; Yang, J.; Han, N.; Zhou, X.; Gong, S.; Yang, J.; Hu, P.; Chen, Y. Highly sensitive and selective ethanol and acetone gas sensors based on modified ZnO nanomaterials. Mater. Des. 2017, 121, 69–76. [Google Scholar] [CrossRef]
  86. Ma, L.; Ma, S.Y.; Kang, H.; Shen, X.F.; Wang, T.T.; Jiang, X.H.; Chen, Q. Preparation of Ag-doped ZnO-SnO2 hollow nanofibers with an enhanced ethanol sensing performance by electrospinning. Mater. Lett. 2017, 209, 188–192. [Google Scholar] [CrossRef]
  87. Mandal, B.; Aaryashree; Singh, R.; Mukherjee, S. Highly Selective and Sensitive Methanol Sensor Using Rose-Like ZnO Microcube and MoO3 Micrograss-Based Composite. IEEE Sens. J. 2018, 18, 2659–2666. [Google Scholar] [CrossRef]
  88. Dong, Z.-M.; Xia, Q.; Ren, H.; Shang, X.; Lu, X.; Joo, S.W.; Huang, J. Preparation of hollow SnO2/ZnO cubes for the high-performance detection of VOCs. Ceram. Int. 2022, 49, 4650–4658. [Google Scholar] [CrossRef]
  89. Nakate, U.T.; Ahmad, R.; Patil, P.; Yu, Y.T.; Hahn, Y.-B. Ultra thin NiO nanosheets for high performance hydrogen gas sensor device. Appl. Surf. Sci. 2020, 506, 144971. [Google Scholar] [CrossRef]
  90. Li, C.; Choi, P.G.; Kim, K.; Masuda, Y. High performance acetone gas sensor based on ultrathin porous NiO nanosheet. Sens. Actuators B Chem. 2022, 367, 132143. [Google Scholar] [CrossRef]
  91. Yin, M.; Zhu, Z. Mesoporous NiO as an ultra-highly sensitive and selective gas sensor for sensing of trace ammonia at room temperature. J. Alloys Compd. 2019, 789, 941–947. [Google Scholar] [CrossRef]
  92. Yu, Q.; Gong, X.; Jiang, Y.; Zhao, L.; Wang, T.; Liu, F.; Yan, X.; Liang, X.; Liu, F.; Sun, P.; et al. Bimetallic MOFs-derived core-shell structured mesoporous Sn-doped NiO for conductometric ppb-level xylene gas sensors. Sens. Actuators B Chem. 2022, 372, 132620. [Google Scholar] [CrossRef]
  93. Ayyala, S.K.; Covington, J.A. Nickel-Oxide Based Thick-Film Gas Sensors for Volatile Organic Compound Detection. Chemosensors 2021, 9, 247. [Google Scholar] [CrossRef]
  94. John, R.A.B.; Shruthi, J.; Ramana Reddy, M.V.; Ruban Kumar, A. Manganese doped nickel oxide as room temperature gas sensor for formaldehyde detection. Ceram. Int. 2022, 48, 17654–17667. [Google Scholar] [CrossRef]
  95. Montuschi, P.; Paris, D.; Melck, D.; Lucidi, V.; Ciabattoni, G.; Raia, V.; Calabrese, C.; Bush, A.; Barnes, P.J.; Motta, A. NMR spectroscopy metabolomic profiling of exhaled breath condensate in patients with stable and unstable cystic fibrosis. Thorax 2012, 67, 222–228. [Google Scholar] [CrossRef] [PubMed]
  96. Jian, Y.; Zhang, N.; Liu, T.; Zhu, Y.; Wang, D.; Dong, H.; Guo, L.; Qu, D.; Jiang, X.; Du, T.; et al. Artificially Intelligent Olfaction for Fast and Noninvasive Diagnosis of Bladder Cancer from Urine. ACS Sens. 2022, 7, 1720–1731. [Google Scholar] [CrossRef] [PubMed]
  97. Yang, Z.; Zhang, D.; Wang, D. Carbon monoxide gas sensing properties of metal-organic frameworks-derived tin dioxide nanoparticles/molybdenum diselenide nanoflowers. Sens. Actuators B Chem. 2020, 304, 127369. [Google Scholar] [CrossRef]
  98. Zhang, D.; Sun, Y.; Jiang, C.; Zhang, Y. Room temperature hydrogen gas sensor based on palladium decorated tin oxide/molybdenum disulfide ternary hybrid via hydrothermal route. Sens. Actuators B Chem. 2017, 242, 15–24. [Google Scholar] [CrossRef]
  99. Comini, E.; Yubao, L.; Brando, Y.; Sberveglieri, G. Gas sensing properties of MoO3 nanorods to CO and CH3OH. Chem. Phys. Lett. 2005, 407, 368–371. [Google Scholar] [CrossRef]
  100. Angiola, M.; Alsaif, M.M.; Kalantar-zadeh, K.; Wisitsoraat, A.; Wlodarski, W.; Martucci, A. Optical Hydrogen Sensing Based on Hybrid 2D MoO3/Au Nanoparticles. Procedia Eng. 2015, 120, 1141–1144. [Google Scholar] [CrossRef]
  101. Shafi, P.M.; Dhanabal, R.; Chithambararaj, A.; Velmathi, S.; Bose, A.C. α-MnO2/h-MoO3 Hybrid Material for High Performance Supercapacitor Electrode and Photocatalyst. ACS Sustain. Chem. Eng. 2017, 5, 4757–4770. [Google Scholar] [CrossRef]
  102. Wang, J.; Zhou, Q.; Peng, S.; Xu, L.; Zeng, W. Volatile Organic Compounds Gas Sensors Based on Molybdenum Oxides: A Mini Review. Front. Chem. 2020, 8, 339. [Google Scholar] [CrossRef]
  103. Yang, S.; Liu, Y.; Chen, W.; Jin, W.; Zhou, J.; Zhang, H.; Zakharova, G.S. High sensitivity and good selectivity of ultralong MoO3 nanobelts for trimethylamine gas. Sens. Actuators B Chem. 2016, 226, 478–485. [Google Scholar] [CrossRef]
  104. Liu, A.; Liu, T.; Fu, H.; Yin, X.; Liu, K.; Yu, J. Enhanced performance of Zn and Co co-doped MoO3 nanosheets as gas sensor for n-butylamine. Ceram. Int. 2022, 48, 32986–32993. [Google Scholar] [CrossRef]
  105. Shen, J.; Guo, S.; Chen, C.; Sun, L.; Wen, S.; Chen, Y.; Ruan, S. Synthesis of Ni-doped α-MoO3 nanolamella and their improved gas sensing properties. Sens. Actuators B Chem. 2017, 252, 757–763. [Google Scholar] [CrossRef]
  106. Wang, L.; Wang, S.; Fu, H.; Wang, Y.; Yu, K. Synthesis of Au Nanoparticles Functionalized 1D α-MoO3 Nanobelts and Their Gas Sensing Properties. Nano 2018, 13, 1850115. [Google Scholar] [CrossRef]
  107. Shen, S.; Zhang, X.; Cheng, X.; Xu, Y.; Gao, S.; Zhao, H.; Zhou, X.; Huo, L. Oxygen-Vacancy-Enriched Porous α-MoO3 Nanosheets for Trimethylamine Sensing. ACS Appl. Nano Mater. 2019, 2, 8016–8026. [Google Scholar] [CrossRef]
  108. Ji, H.; Zeng, W.; Li, Y. Assembly of 2D nanosheets into flower-like MoO3: New insight into the petal thickness affect on gas-sensing properties. Mater. Res. Bull. 2019, 118, 110476. [Google Scholar] [CrossRef]
  109. Li, T.; Rui, Y.; Zhang, X.; Shi, J.; Wang, X.; Wang, Y.; Yang, J.; Zhang, Q. Anatase TiO2 nanorod arrays as high-performance electron transport layers for perovskite solar cells. J. Alloys Compd. 2020, 849, 156629. [Google Scholar] [CrossRef]
  110. He, X.; Wu, M.; Ao, Z.; Lai, B.; Zhou, Y.; An, T.; Wang, S. Metal–organic frameworks derived C/TiO2 for visible light photocatalysis: Simple synthesis and contribution of carbon species. J. Hazard. Mater. 2021, 403, 124048. [Google Scholar] [CrossRef]
  111. Pathak, A.K.; Bhardwaj, V.; Gangwar, R.K.; De, M.; Singh, V.K. Fabrication and characterization of TiO2 coated cone shaped nano-fiber pH sensor. Optics Commun. 2016, 386, 43–48. [Google Scholar] [CrossRef]
  112. Chen, N.; Li, Y.; Deng, D.; Liu, X.; Xing, X.; Xiao, X.; Wang, Y. Acetone sensing performances based on nanoporous TiO2 synthesized by a facile hydrothermal method. Sens. Actuators B Chem. 2017, 238, 491–500. [Google Scholar] [CrossRef]
  113. Pozos, H.G.; Krishna, K.T.V.; de la Luz Olvera Amador, M.; Kudriavtsev, Y.; Alvarez, A.M. TiO2 thin film based gas sensors for CO-detection. J. Mater. Sci. Mater. Electron. 2018, 29, 15829–15837. [Google Scholar] [CrossRef]
  114. Liao, Y.; Que, W.; Jia, Q.; He, Y.; Zhang, J.; Zhong, P. Controllable synthesis of brookite/anatase/rutile TiO2 nanocomposites and single-crystalline rutile nanorods array. J. Mater. Chem. 2012, 22, 7937. [Google Scholar] [CrossRef]
  115. Tian, X.; Cui, X.; Lai, T.; Ren, J.; Yang, Z.; Xiao, M.; Wang, B.; Xiao, X.; Wang, Y. Gas sensors based on TiO2 nanostructured materials for the detection of hazardous gases: A review. Nano Mater. Sci. 2021, 3, 390–403. [Google Scholar] [CrossRef]
  116. Reghunath, S.; Pinheiro, D.; KR, S.D. A review of hierarchical nanostructures of TiO2: Advances and applications. Appl. Surf. Sci. Adv. 2021, 3, 100063. [Google Scholar] [CrossRef]
  117. Dubey, R.S. Temperature-dependent phase transformation of TiO2 nanoparticles synthesized by sol-gel method. Mater. Lett. 2018, 215, 312–317. [Google Scholar] [CrossRef]
  118. Navale, S.T.; Yang, Z.B.; Liu, C.; Cao, P.J.; Patil, V.B.; Ramgir, N.S.; Mane, R.S.; Stadler, F.J. Enhanced acetone sensing properties of titanium dioxide nanoparticles with a sub-ppm detection limit. Sens. Actuators B Chem. 2018, 255, 1701–1710. [Google Scholar] [CrossRef]
  119. Sugahara, T.; Alipour, L.; Hirose, Y.; Ekubaru, Y.; Nakamura, J.; Ono, H.; Harada, N.; Suganuma, K. Formation of Metal-Organic Decomposition Derived Nanocrystalline Structure Titanium Dioxide by Heat Sintering and Photosintering Methods for Advanced Coating Process, and Its Volatile Organic Compounds’ Gas-Sensing Properties. ACS Appl. Electron. Mater. 2020, 2, 1670–1678. [Google Scholar] [CrossRef]
  120. Lim, C.-J.; Park, J.-W. Luminescent oxygen-sensing films with improved sensitivity based on light scattering by TiO2 particles. Sens. Actuators B Chem. 2017, 253, 934–941. [Google Scholar] [CrossRef]
  121. Li, X.; Li, X.; Wang, J.; Lin, S. Highly sensitive and selective room-temperature formaldehyde sensors using hollow TiO2 microspheres. Sens. Actuators B Chem. 2015, 219, 158–163. [Google Scholar] [CrossRef]
  122. Gakhar, T.; Hazra, A. C60-encapsulated TiO2 nanoparticles for selective and ultrahigh sensitive detection of formaldehyde. Nanotechnology 2021, 32, 505505. [Google Scholar] [CrossRef]
  123. Lee, K.; Kirchgeorg, R.; Schmuki, P. Role of Transparent Electrodes for High Efficiency TiO2 Nanotube Based Dye-Sensitized Solar Cells. J. Phys. Chem. C 2014, 118, 16562–16566. [Google Scholar] [CrossRef]
  124. Bindra, P.; Gangopadhyay, S.; Hazra, A. 1-D TiO2 Nanorods Array-Based Parallel Electrode Sensor for Selective and Stable Detection of Organic Vapors. IEEE Sens. J. 2020, 20, 664–671. [Google Scholar] [CrossRef]
  125. Wang, Y.; Ge, S.; Zhang, L.; Yu, J.; Yan, M.; Huang, J. Visible photoelectrochemical sensing platform by in situ generated CdS quantum dots decorated branched-TiO2 nanorods equipped with Prussian blue electrochromic display. Biosens. Bioelectron. 2017, 89, 859–865. [Google Scholar] [CrossRef] [PubMed]
  126. Rho, W.-Y.; Jeon, H.; Kim, H.-S.; Chung, W.-J.; Suh, J.S.; Jun, B.-H. Recent Progress in Dye-Sensitized Solar Cells for Improving Efficiency: TiO2 Nanotube Arrays in Active Layer. J. Nanomater. 2015, 2015, 247689. [Google Scholar] [CrossRef] [Green Version]
  127. Landau, O.; Rothschild, A. Fibrous TiO2 gas sensors produced by electrospinning. J. Electroceram. 2015, 35, 148–159. [Google Scholar] [CrossRef]
  128. Xu, C.; Gao, D. Two-Stage Hydrothermal Growth of Long ZnO Nanowires for Efficient TiO2 Nanotube-Based Dye-Sensitized Solar Cells. J. Phys. Chem. C 2012, 116, 7236–7241. [Google Scholar] [CrossRef]
  129. Caricato, A.P.; Buonsanti, R.; Catalano, M.; Cesaria, M.; Cozzoli, P.D.; Luches, A.; Manera, M.G.; Martino, M.; Taurino, A.; Rella, R. Films of brookite TiO2 nanorods/nanoparticles deposited by matrix-assisted pulsed laser evaporation as NO2 gas-sensing layers. Appl. Phys. A 2011, 104, 963–968. [Google Scholar] [CrossRef]
  130. Kumar, K.G.; Avinash, B.S.; Rahimi-Gorji, M.; Majdoubi, J. Photocatalytic activity and smartness of TiO2 nanotube arrays for room temperature acetone sensing. J. Mol. Liq. 2020, 300, 112418. [Google Scholar] [CrossRef]
  131. Sennik, E.; Kilinc, N.; Ozturk, Z.Z. Electrical and VOC sensing properties of anatase and rutile TiO2 nanotubes. J. Alloys Compd. 2014, 616, 89–96. [Google Scholar] [CrossRef]
  132. Gakhar, T.; Hazra, A. Oxygen vacancy modulation of titania nanotubes by cathodic polarization and chemical reduction routes for efficient detection of volatile organic compounds. Nanoscale 2020, 12, 9082–9093. [Google Scholar] [CrossRef] [PubMed]
  133. Lin, S.; Li, D.; Wu, J.; Li, X.; Akbar, S.A. A selective room temperature formaldehyde gas sensor using TiO2 nanotube arrays. Sens. Actuators B Chem. 2011, 156, 505–509. [Google Scholar] [CrossRef]
  134. Alev, O.; Şennik, E.; Kılınç, N.; Öztürk, Z.Z. Gas Sensor Application of Hydrothermally Growth TiO2 Nanorods. Procedia Eng. 2015, 120, 1162–1165. [Google Scholar] [CrossRef] [Green Version]
  135. Dutta, K.; Chattopadhyay, P.P.; Lu, C.-W.; Ho, M.-S.; Bhattacharyya, P. A highly sensitive BTX sensor based on electrochemically derived wall connected TiO2 nanotubes. Appl. Surf. Sci. 2015, 354, 353–361. [Google Scholar] [CrossRef]
  136. Bindra, P.; Hazra, A. Selective detection of organic vapors using TiO2 nanotubes based single sensor at room temperature. Sens. Actuators B Chem. 2019, 290, 684–690. [Google Scholar] [CrossRef]
  137. Bindra, P.; Hazra, A. Multi-Layered TiO2 Nanotubes Array-Based Highly Sensitive Room-Temperature Vapor Sensors. IEEE Trans. Nanotechnol. 2019, 18, 13–20. [Google Scholar] [CrossRef]
  138. Zhao, G.; Xuan, J.; Gong, Q.; Wang, L.; Ren, J.; Sun, M.; Jia, F.; Yin, G.; Liu, B. In Situ Growing Double-Layer TiO2 Nanorod Arrays on New-Type FTO Electrodes for Low-Concentration NH3 Detection at Room Temperature. ACS Appl. Mater. Interfaces 2020, 12, 8573–8582. [Google Scholar] [CrossRef] [PubMed]
  139. Cai, Z.; Park, S. Highly selective acetone sensor based on Co3O4-decorated porous TiO2 nanofibers. J. Alloys Compd. 2022, 919, 165875. [Google Scholar] [CrossRef]
  140. Kumari, A.; Shanbogh, S.M.; Udachyan, I.; Kandaiah, S.; Roy, A.; Varade, V.; Ponnam, A. Interface-Driven Multifunctionality in Two-Dimensional TiO2 Nanosheet/Poly(Dimercaptothiadiazole-Triazine) Hybrid Resistive Random Access Memory Device. ACS Appl. Mater. Interfaces 2020, 12, 56568–56578. [Google Scholar] [CrossRef]
  141. Dral, A.P.; ten Elshof, J.E. 2D metal oxide nanoflakes for sensing applications: Review and perspective. Sens. Actuators B Chem. 2018, 272, 369–392. [Google Scholar] [CrossRef] [Green Version]
  142. Sonker, R.K.; Hitkari, G.; Sabhajeet, S.R.; Sikarwar, S.; Rahul; Singh, S. Green synthesis of TiO2 nanosheet by chemical method for the removal of Rhodamin B from industrial waste. Mater. Sci. Eng. B 2020, 258, 114577. [Google Scholar] [CrossRef]
  143. Ge, W.; Jiao, S.; Chang, Z.; He, X.; Li, Y. Ultrafast Response and High Selectivity toward Acetone Vapor Using Hierarchical Structured TiO2 Nanosheets. ACS Appl. Mater. Interfaces 2020, 12, 13200–13207. [Google Scholar] [CrossRef]
  144. Horprathum, M.; Eiamchai, P.; Chindaudom, P.; Pokaipisit, A.; Limsuwan, P. Oxygen Partial Pressure Dependence of the Properties of TiO2 Thin Films Deposited by DC Reactive Magnetron Sputtering. Procedia Eng. 2012, 32, 676–682. [Google Scholar] [CrossRef] [Green Version]
  145. Hayashi, K.; Nozaki, K.; Tan, Z.; Fujita, K.; Nemoto, R.; Yamashita, K.; Miura, H.; Itaka, K.; Ohara, S. Enhanced Antibacterial Property of Facet-Engineered TiO2 Nanosheet in Presence and Absence of Ultraviolet Irradiation. Materials 2019, 13, 78. [Google Scholar] [CrossRef]
  146. Bhowmik, B.; Bhattacharyya, P. Efficient Gas Sensor Devices Based on Surface Engineered Oxygen Vacancy Controlled TiO2 Nanosheets. IEEE Trans. Electron Devices 2017, 64, 2357–2363. [Google Scholar] [CrossRef]
  147. Wang, B.; Deng, L.; Sun, L.; Lei, Y.; Wu, N.; Wang, Y. Growth of TiO2 nanostructures exposed {001} and {110} facets on SiC ultrafine fibers for enhanced gas sensing performance. Sens. Actuators B Chem. 2018, 276, 57–64. [Google Scholar] [CrossRef]
  148. Chinh, N.D.; Kim, C.; Kim, D. UV-light-activated H2S gas sensing by a TiO2 nanoparticulate thin film at room temperature. J. Alloys Compd. 2019, 778, 247–255. [Google Scholar] [CrossRef]
  149. Wilson, R.; Simion, C.; Blackman, C.; Carmalt, C.; Stanoiu, A.; Di Maggio, F.; Covington, J. The Effect of Film Thickness on the Gas Sensing Properties of Ultra-Thin TiO2 Films Deposited by Atomic Layer Deposition. Sensors 2018, 18, 735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Wang, W.; Liu, F.; Wang, B.; Wang, Y. Effect of defects in TiO2 nanoplates with exposed {001} facets on the gas sensing properties. Chin. Chem. Lett. 2019, 30, 1261–1265. [Google Scholar] [CrossRef]
  151. Tian, X.; Hu, Z.; Wang, T.; Wang, H.; Zhang, Q.; Wei, X. Influence of multi-layer TiO2/SnO2 heterojunctions on fast and sensitive ethanol detection. Vacuum 2023, 207, 111620. [Google Scholar] [CrossRef]
  152. Lee, J.-H. Gas sensors using hierarchical and hollow oxide nanostructures: Overview. Sens. Actuators B Chem. 2009, 140, 319–336. [Google Scholar] [CrossRef]
  153. Prabakar, S.; Bumby, C.W.; Tilley, R.D. Liquid-Phase Synthesis of Flower-like and Flake-like Titanium Disulfide Nanostructures. Chem. Mater. 2009, 21, 1725–1730. [Google Scholar] [CrossRef]
  154. Gao, X.; Li, Y.; Zeng, W.; Zhang, C.; Wei, Y. Hydrothermal synthesis of agglomerating TiO2 nanoflowers and its gas sensing. J. Mater. Sci. Mater. Electron. 2017, 28, 18781–18786. [Google Scholar] [CrossRef]
  155. Akhtar, M.; Umar, A.; Sood, S.; Jung, I.; Hegazy, H.; Algarni, H. Rapid Growth of TiO2 Nanoflowers via Low-Temperature Solution Process: Photovoltaic and Sensing Applications. Materials 2019, 12, 566. [Google Scholar] [CrossRef] [PubMed]
  156. Wang, M.; Zhu, Y.; Meng, D.; Wang, K.; Wang, C. A novel room temperature ethanol gas sensor based on 3D hierarchical flower-like TiO2 microstructures. Mater. Lett. 2020, 277, 128372. [Google Scholar] [CrossRef]
  157. Liu, M.; Lu, W.; Zhao, L.; Zhou, C.; Li, H.; Wang, W. Fabrication and photocatalytical properties of flower-like TiO2 nanostructures. Trans. Nonferrous Met. Soc. China 2010, 20, 2299–2302. [Google Scholar] [CrossRef]
  158. Wategaonkar, S.B.; Pawar, R.P.; Parale, V.G.; Pakhare, K.S.; Sargar, B.M.; Mane, R.K. Synthesis and Characterizations of 3D TiO2 Nanoflowers Thin Film: Hydrothermal Method. Macromol. Symp. 2020, 393, 2000040. [Google Scholar] [CrossRef]
  159. Siva Prasad, M.; Chen, R.; Ni, H.; Kiran Kumar, K. Directly grown of 3D-nickel oxide nano flowers on TiO2 nanowire arrays by hydrothermal route for electrochemical determination of naringenin flavonoid in vegetable samples. Arab. J. Chem. 2020, 13, 1520–1531. [Google Scholar] [CrossRef]
  160. Bhowmik, B.; Manjuladevi, V.; Gupta, R.K.; Bhattacharyya, P. Highly Selective Low-Temperature Acetone Sensor Based on Hierarchical 3-D TiO2 Nanoflowers. IEEE Sens. J. 2016, 16, 3488–3495. [Google Scholar] [CrossRef]
  161. Wang, C.; Wang, Z.-G.; Xi, R.; Zhang, L.; Zhang, S.-H.; Wang, L.-J.; Pan, G.-B. In situ synthesis of flower-like ZnO on GaN using electrodeposition and its application as ethanol gas sensor at room temperature. Sens. Actuators B Chem. 2019, 292, 270–276. [Google Scholar] [CrossRef]
  162. Liu, C.; Wang, B.; Wang, T.; Liu, J.; Sun, P.; Chuai, X.; Lu, G. Enhanced gas sensing characteristics of the flower-like ZnFe2O4/ZnO microstructures. Sens. Actuators B Chem. 2017, 248, 902–909. [Google Scholar] [CrossRef]
  163. Jagannathan, M.; Dhinasekaran, D.; Rajendran, A.R.; Subramaniam, B. Selective room temperature ammonia gas sensor using nanostructured ZnO/CuO@graphene on paper substrate. Sens. Actuators B Chem. 2022, 350, 130833. [Google Scholar] [CrossRef]
  164. Park, Y.; Yoo, R.; Park, S.R.; Lee, J.H.; Jung, H.; Lee, H.-S.; Lee, W. Highly sensitive and selective isoprene sensing performance of ZnO quantum dots for a breath analyzer. Sens. Actuators B Chem. 2019, 290, 258–266. [Google Scholar] [CrossRef]
  165. Wang, H.; Luo, Y.; Liu, B.; Gao, L.; Duan, G. CuO nanoparticle loaded ZnO hierarchical heterostructure to boost H2S sensing with fast recovery. Sens. Actuators B Chem. 2021, 338, 129806. [Google Scholar] [CrossRef]
  166. Li, X.; Li, Y.; Sun, G.; Zhang, B.; Wang, Y.; Zhang, Z. Enhanced CH4 sensitivity of porous nanosheets-assembled ZnO microflower by decoration with Zn2SnO4. Sens. Actuators B Chem. 2020, 304, 127374. [Google Scholar] [CrossRef]
  167. Lou, C.; Yang, C.; Zheng, W.; Liu, X.; Zhang, J. Atomic layer deposition of ZnO on SnO2 nanospheres for enhanced formaldehyde detection. Sens. Actuators B Chem. 2021, 329, 129218. [Google Scholar] [CrossRef]
  168. Mun, Y.; Park, S.; An, S.; Lee, C.; Kim, H.W. NO2 gas sensing properties of Au-functionalized porous ZnO nanosheets enhanced by UV irradiation. Ceram. Int. 2013, 39, 8615–8622. [Google Scholar] [CrossRef]
  169. Patil, N.B.; Nimbalkar, A.R.; Patil, M.G. ZnO thin film prepared by a sol-gel spin coating technique for NO2 detection. Mater. Sci. Eng. B 2018, 227, 53–60. [Google Scholar] [CrossRef]
  170. Gagaoudakis, E.; Michail, G.; Katerinopoulou, D.; Moschovis, K.; Iliopoulos, E.; Kiriakidis, G.; Binas, V.; Aperathitis, E. Transparent p-type NiO:Al thin films as room temperature hydrogen and methane gas sensors. Mater. Sci. Semicond. Process. 2020, 109, 104922. [Google Scholar] [CrossRef]
  171. Wu, Z.; Zhang, H.; Ji, H.; Yuan, Z.; Meng, F. Novel combined waveform temperature modulation method of NiO-In2O3 based gas sensor for measuring and identifying VOC gases. J. Alloys Compd. 2022, 918, 165510. [Google Scholar] [CrossRef]
  172. Fu, X.; Yang, P.; Xiao, X.; Zhou, D.; Huang, R.; Zhang, X.; Cao, F.; Xiong, J.; Hu, Y.; Tu, Y.; et al. Ultra-fast and highly selective room-temperature formaldehyde gas sensing of Pt-decorated MoO3 nanobelts. J. Alloys Compd. 2019, 797, 666–675. [Google Scholar] [CrossRef]
  173. Cong, S.; Sugahara, T.; Wei, T.; Jiu, J.; Hirose, Y.; Nagao, S.; Suganuma, K. Diverse Adsorption/Desorption Abilities Originating from the Nanostructural Morphology of VOC Gas Sensing Devices Based on Molybdenum Trioxide Nanorod Arrays. Adv. Mater. Interfaces 2016, 3, 1600252. [Google Scholar] [CrossRef]
  174. Mo, Y.; Tan, Z.; Sun, L.; Lu, Y.; Liu, X. Ethanol-sensing properties of α-MoO3 nanobelts synthesized by hydrothermal method. J. Alloys Compd. 2020, 812, 152166. [Google Scholar] [CrossRef]
  175. Yang, S.; Liu, Y.; Chen, T.; Jin, W.; Yang, T.; Cao, M.; Liu, S.; Zhou, J.; Zakharova, G.S.; Chen, W. Zn doped MoO3 nanobelts and the enhanced gas sensing properties to ethanol. Appl. Surf. Sci. 2017, 393, 377–384. [Google Scholar] [CrossRef]
  176. Qu, F.; Zhou, X.; Zhang, B.; Zhang, S.; Jiang, C.; Ruan, S.; Yang, M. Fe2O3 nanoparticles-decorated MoO3 nanobelts for enhanced chemiresistive gas sensing. J. Alloys Compd. 2019, 782, 672–678. [Google Scholar] [CrossRef]
  177. Jiang, W.; Meng, L.; Zhang, S.; Chuai, X.; Zhou, Z.; Hu, C.; Sun, P.; Liu, F.; Yan, X.; Lu, G. Design of highly sensitive and selective xylene gas sensor based on Ni-doped MoO3 nano-pompon. Sens. Actuators B Chem. 2019, 299, 126888. [Google Scholar] [CrossRef]
  178. Gardon, M.; Monereo, O.; Dosta, S.; Vescio, G.; Cirera, A.; Guilemany, J.M. New procedures for building-up the active layer of gas sensors on flexible polymers. Surf. Coat. Technol. 2013, 235, 848–852. [Google Scholar] [CrossRef]
  179. Mei, H.; Zhou, S.; Lu, M.; Zhao, Y.; Cheng, L. Construction of pine-branch-like α-Fe2O3/TiO2 hierarchical heterostructure for gas sensing. Ceram. Int. 2020, 46, 18675–18682. [Google Scholar] [CrossRef]
  180. Chen, N.; Deng, D.; Li, Y.; Xing, X.; Liu, X.; Xiao, X.; Wang, Y. The xylene sensing performance of WO3 decorated anatase TiO2 nanoparticles as a sensing material for a gas sensor at a low operating temperature. RSC Adv. 2016, 6, 49692–49701. [Google Scholar] [CrossRef]
  181. Wang, Z.; Ali Haidry, A.; Xie, L.; Zavabeti, A.; Li, Z.; Yin, W.; Lontio Fomekong, R.; Saruhan, B. Acetone sensing applications of Ag modified TiO2 porous nanoparticles synthesized via facile hydrothermal method. Appl. Surf. Sci. 2020, 533, 147383. [Google Scholar] [CrossRef]
  182. Tong, X.; Shen, W.; Zhang, X.; Corriou, J.-P.; Xi, H. Synthesis and density functional theory study of free-standing Fe-doped TiO2 nanotube array film for H2S gas sensing properties at low temperature. J. Alloys Compd. 2020, 832, 155015. [Google Scholar] [CrossRef]
  183. Xu, Y.; Zheng, L.; Yang, C.; Zheng, W.; Liu, X.; Zhang, J. Chemiresistive sensors based on core-shell ZnO@TiO2 nanorods designed by atomic layer deposition for n-butanol detection. Sens. Actuators B Chem. 2020, 310, 127846. [Google Scholar] [CrossRef]
  184. Feng, Z.; Zhang, L.; Chen, W.; Peng, Z.; Li, Y. A strategy for supportless sensors: Fluorine doped TiO2 nanosheets directly grown onto Ti foam enabling highly sensitive detection toward acetone. Sens. Actuators B Chem. 2020, 322, 128633. [Google Scholar] [CrossRef]
  185. Chen, K.; Pi, M.; Zhang, D. Ultraviolet irradiation enhanced formaldehyde-sensing performance based on SnO2@TiO2 nanofiber heteroarchitectures. J. Phys. D Appl. Phys. 2020, 53, 125301. [Google Scholar] [CrossRef]
  186. Tung, T.T.; Tran, M.T.; Feller, J.-F.; Castro, M.; Van Ngo, T.; Hassan, K.; Nine, M.J.; Losic, D. Graphene and metal organic frameworks (MOFs) hybridization for tunable chemoresistive sensors for detection of volatile organic compounds (VOCs) biomarkers. Carbon 2020, 159, 333–344. [Google Scholar] [CrossRef]
  187. Ravi Kumar, Y.; Deshmukh, K.; Kovářík, T.; Khadheer Pasha, S.K. A systematic review on 2D materials for volatile organic compound sensing. Coord. Chem. Rev. 2022, 461, 214502. [Google Scholar] [CrossRef]
  188. Pathak, A.; Viphavakit, C.; Rahman, B.M.A.; Singh, V. A Highly Sensitive SPR Refractive Index Sensor Based on Microfluidic Channel Assisted with Graphene-Ag Composite Nanowire. IEEE Photonics J. 2021, 13, 1–8. [Google Scholar] [CrossRef]
  189. Pargoletti, E.; Verga, S.; Chiarello, G.L.; Longhi, M.; Cerrato, G.; Giordana, A.; Cappelletti, G. Exploring SnxTi1−xO2 Solid Solutions Grown onto Graphene Oxide (GO) as Selective Toluene Gas Sensors. Nanomaterials 2020, 10, 761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Wu, J.; Li, Z.; Xie, X.; Tao, K.; Liu, C.; Khor, K.A.; Miao, J.; Norford, L.K. 3D superhydrophobic reduced graphene oxide for activated NO2 sensing with enhanced immunity to humidity. J. Mater. Chem. A 2018, 6, 478–488. [Google Scholar] [CrossRef]
  191. Zito, C.A.; Perfecto, T.M.; Fonseca, C.S.; Volanti, D.P. Effective reduced graphene oxide sheets/hierarchical flower-like NiO composites for methanol sensing under high humidity. New J. Chem. 2018, 42, 8638–8645. [Google Scholar] [CrossRef]
  192. Modenes-Junior, M.A.; Zito, C.A.; Perfecto, T.M.; Volanti, D.P. Ethanol detection using composite based on reduced graphene oxide and CuO hierarchical structure under wet atmosphere. Mater. Sci. Eng. B 2019, 248, 114385. [Google Scholar] [CrossRef]
  193. Pargoletti, E.; Tricoli, A.; Pifferi, V.; Orsini, S.; Longhi, M.; Guglielmi, V.; Cerrato, G.; Falciola, L.; Derudi, M.; Cappelletti, G. An electrochemical outlook upon the gaseous ethanol sensing by graphene oxide-SnO2 hybrid materials. Appl. Surf. Sci. 2019, 483, 1081–1089. [Google Scholar] [CrossRef]
  194. Wu, J.; Wu, Z.; Ding, H.; Wei, Y.; Huang, W.; Yang, X.; Li, Z.; Qiu, L.; Wang, X. Flexible, 3D SnS2/Reduced graphene oxide heterostructured NO2 sensor. Sens. Actuators B Chem. 2020, 305, 127445. [Google Scholar] [CrossRef]
  195. Zhao, Y.; Li, H.; Li, Y.; Ma, Y.; Yang, H.; Liu, H.; Ren, X.; Zhao, H. Layered SnO2 nanorods arrays anchored on reduced graphene oxide for ultra-high and ppb level formaldehyde sensing. Sens. Actuators B Chem. 2021, 346, 130452. [Google Scholar] [CrossRef]
  196. Jia, X.; Yu, S.; Cheng, C.; Yang, J.; Li, Y.; Wang, S.; Song, H. Ag nanoparticles modified Fe3O4/reduced graphene oxide and their acetone sensing properties. Mater. Chem. Phys. 2022, 276, 125378. [Google Scholar] [CrossRef]
  197. Wang, Z.; Gao, S.; Fei, T.; Liu, S.; Zhang, T. Construction of ZnO/SnO2 Heterostructure on Reduced Graphene Oxide for Enhanced Nitrogen Dioxide Sensitive Performances at Room Temperature. ACS Sens. 2019, 4, 2048–2057. [Google Scholar] [CrossRef]
  198. Sen, S.; Kundu, S. Reduced graphene oxide (rGO) decorated ZnO-SnO2: A ternary nanocomposite towards improved low concentration VOC sensing performance. J. Alloys Compd. 2021, 881, 160406. [Google Scholar] [CrossRef]
  199. Liang, S.; Wang, H.; Tao, X. Effects of multiwalled carbon nanotubes and reduced graphene oxide of different proportions on the electrothermal properties of cationic cellulose nanofibril-based composites. J. Mater. Res. Technol. 2022, 17, 2388–2399. [Google Scholar] [CrossRef]
  200. Yang, G.; Li, L.; Lee, W.B.; Ng, M.C. Structure of graphene and its disorders: A review. Sci. Technol. Adv. Mater. 2018, 19, 613–648. [Google Scholar] [CrossRef] [Green Version]
  201. Ibrahim, K.S. Carbon nanotubes-properties and applications: A review. Carbon Lett. 2013, 14, 131–144. [Google Scholar] [CrossRef] [Green Version]
  202. De Volder, M.F.L.; Tawfick, S.H.; Baughman, R.H.; Hart, A.J. Carbon Nanotubes: Present and Future Commercial Applications. Science 2013, 339, 535–539. [Google Scholar] [CrossRef] [Green Version]
  203. Lekawa-Raus, A.; Patmore, J.; Kurzepa, L.; Bulmer, J.; Koziol, K. Electrical Properties of Carbon Nanotube Based Fibers and Their Future Use in Electrical Wiring. Adv. Funct. Mater. 2014, 24, 3661–3682. [Google Scholar] [CrossRef]
  204. Ruoff, R.S.; Qian, D.; Liu, W.K. Mechanical properties of carbon nanotubes: Theoretical predictions and experimental measurements. Comptes Rendus Phys. 2003, 4, 993–1008. [Google Scholar] [CrossRef]
  205. Gulotty, R.; Castellino, M.; Jagdale, P.; Tagliaferro, A.; Balandin, A.A. Effects of Functionalization on Thermal Properties of Single-Wall and Multi-Wall Carbon Nanotube–Polymer Nanocomposites. ACS Nano 2013, 7, 5114–5121. [Google Scholar] [CrossRef]
  206. Deng, L.; Young, R.J.; Kinloch, I.A.; Sun, R.; Zhang, G.; Noé, L.; Monthioux, M. Coefficient of thermal expansion of carbon nanotubes measured by Raman spectroscopy. Appl. Phys. Lett. 2014, 104, 051907. [Google Scholar] [CrossRef] [Green Version]
  207. Tao, F.; Liu, Y.; Ren, X.; Jiang, A.; Wei, H.; Zhai, X.; Wang, F.; Stock, H.-R.; Wen, S.; Ren, F. Carbon nanotube-based nanomaterials for high-performance sodium-ion batteries: Recent advances and perspectives. J. Alloys Compd. 2021, 873, 159742. [Google Scholar] [CrossRef]
  208. Bahoumina, P.; Hallil, H.; Lachaud, J.L.; Abdelghani, A.; Frigui, K.; Bila, S.; Baillargeat, D.; Ravichandran, A.; Coquet, P.; Paragua, C.; et al. Microwave flexible gas sensor based on polymer multi wall carbon nanotubes sensitive layer. Sens. Actuators B Chem. 2017, 249, 708–714. [Google Scholar] [CrossRef]
  209. Liu, Z.; Yang, T.; Dong, Y.; Wang, X. A Room Temperature VOCs Gas Sensor Based on a Layer by Layer Multi-Walled Carbon Nanotubes/Poly-ethylene Glycol Composite. Sensors 2018, 18, 3113. [Google Scholar] [CrossRef] [Green Version]
  210. Janfaza, S.; Banan Nojavani, M.; Nikkhah, M.; Alizadeh, T.; Esfandiar, A.; Ganjali, M.R. A selective chemiresistive sensor for the cancer-related volatile organic compound hexanal by using molecularly imprinted polymers and multiwalled carbon nanotubes. Microchim. Acta 2019, 186, 137. [Google Scholar] [CrossRef]
  211. Singh, S.; Sharma, S.; Singh, R.C.; Sharma, S. Hydrothermally synthesized MoS2-multi-walled carbon nanotube composite as a novel room-temperature ammonia sensing platform. Appl. Surf. Sci. 2020, 532, 147373. [Google Scholar] [CrossRef]
  212. Yan, L.; Xiong, T.; Zhang, Z.; Yang, H.; Zhang, X.; He, Y.; Bian, J.; Lin, H.; Chen, D. Facile preparation of TPU conductive nanocomposites containing polypyrrole-coated multi-walled carbon nanotubes for a rapid and selective response in volatile organic compounds applications. Compos. Part A Appl. Sci. Manuf. 2022, 157, 106913. [Google Scholar] [CrossRef]
  213. Gałęzowska, G.; Chraniuk, M.; Wolska, L. In vitro assays as a tool for determination of VOCs toxic effect on respiratory system: A critical review. TrAC Trends Anal. Chem. 2016, 77, 14–22. [Google Scholar] [CrossRef]
  214. Della Pina, C.; De Gregorio, M.A.; Clerici, L.; Dellavedova, P.; Falletta, E. Polyaniline (PANI): An innovative support for sampling and removal of VOCs in air matrices. J. Hazard. Mater. 2018, 344, 308–315. [Google Scholar] [CrossRef] [PubMed]
  215. Li, X.; Zhang, L.; Yang, Z.; Wang, P.; Yan, Y.; Ran, J. Adsorption materials for volatile organic compounds (VOCs) and the key factors for VOCs adsorption process: A review. Sep. Purif. Technol. 2020, 235, 116213. [Google Scholar] [CrossRef]
  216. Twumasi, E.; Forslund, M.; Norberg, P.; Sjöström, C. Carbon–silica composites prepared by the precipitation method. Effect of the synthesis parameters on textural characteristics and toluene dynamic adsorption. J. Porous Mater. 2012, 19, 333–343. [Google Scholar] [CrossRef]
  217. Rakić, V.; Rac, V.; Krmar, M.; Otman, O.; Auroux, A. The adsorption of pharmaceutically active compounds from aqueous solutions onto activated carbons. J. Hazard. Mater. 2015, 282, 141–149. [Google Scholar] [CrossRef]
  218. Zhang, X.; Gao, B.; Creamer, A.E.; Cao, C.; Li, Y. Adsorption of VOCs onto engineered carbon materials: A review. J. Hazard. Mater. 2017, 338, 102–123. [Google Scholar] [CrossRef] [PubMed]
  219. Ma, X.; Zhang, Z.; Wu, H.; Li, J.; Yang, L. Adsorption of Volatile Organic Compounds at Medium-High Temperature Conditions by Activated Carbons. Energy Fuels 2020, 34, 3679–3690. [Google Scholar] [CrossRef]
  220. Arami-Niya, A.; Wan Daud, W.M.A.; Mjalli, F.S.; Abnisa, F.; Shafeeyan, M.S. Production of microporous palm shell based activated carbon for methane adsorption: Modeling and optimization using response surface methodology. Chem. Eng. Res. Des. 2012, 90, 776–784. [Google Scholar] [CrossRef]
  221. Li, L.; Song, J.; Yao, X.; Huang, G.; Liu, Z.; Tang, L. Adsorption of volatile organic compounds on three activated carbon samples: Effect of pore structure. J. Cent. South Univ. 2012, 19, 3530–3539. [Google Scholar] [CrossRef]
  222. Shen, Y.; Zhang, N. Facile synthesis of porous carbons from silica-rich rice husk char for volatile organic compounds (VOCs) sorption. Bioresour. Technol. 2019, 282, 294–300. [Google Scholar] [CrossRef]
  223. Kang, H.-L.; Hong, D.; Yoon, S.; Song, S.; Kim, Y.J.; Kim, W.-H.; Cho, H.; Kim, H.-W.; Seong, W.-K.; Shin, D.-S.; et al. Gas concentrator with an activated carbon adsorbent for breath analysis. Sens. Actuators B Chem. 2021, 330, 129341. [Google Scholar] [CrossRef]
  224. Yin, F.; Yue, W.; Li, Y.; Gao, S.; Zhang, C.; Kan, H.; Niu, H.; Wang, W.; Guo, Y. Carbon-based nanomaterials for the detection of volatile organic compounds: A review. Carbon 2021, 180, 274–297. [Google Scholar] [CrossRef]
  225. Vessalli, B.A.; Zito, C.A.; Perfecto, T.M.; Volanti, D.P.; Mazon, T. ZnO nanorods/graphene oxide sheets prepared by chemical bath deposition for volatile organic compounds detection. J. Alloys Compd. 2017, 696, 996–1003. [Google Scholar] [CrossRef]
  226. Tang, N.; Jiang, Y.; Qu, H.; Duan, X. Graphene Oxide-Doped Conducting Polymer Nanowires Fabricated by Soft Lithography for Gas Sensing Applications. IEEE Sens. J. 2018, 18, 7765–7771. [Google Scholar] [CrossRef]
  227. Pargoletti, E.; Hossain, U.H.; Di Bernardo, I.; Chen, H.; Tran-Phu, T.; Chiarello, G.L.; Lipton-Duffin, J.; Pifferi, V.; Tricoli, A.; Cappelletti, G. Engineering of SnO2–Graphene Oxide Nanoheterojunctions for Selective Room-Temperature Chemical Sensing and Optoelectronic Devices. ACS Appl. Mater. Interfaces 2020, 12, 39549–39560. [Google Scholar] [CrossRef] [PubMed]
  228. Duong, V.T.; Nguyen, C.T.; Luong, H.B.; Nguyen, D.C.; Nguyen, H.L. Ultralow-detection limit ammonia gas sensors at room temperature based on MWCNT/WO3 nanocomposite and effect of humidity. Solid State Sci. 2021, 113, 106534. [Google Scholar] [CrossRef]
  229. Roy, N.; Sinha, R.; Daniel, T.T.; Nemade, H.B.; Mandal, T.K. Highly Sensitive Room Temperature CO Gas Sensor Based on MWCNT-PDDA Composite. IEEE Sens. J. 2020, 20, 13245–13252. [Google Scholar] [CrossRef]
  230. Jafari, N.; Zeinali, S. Highly Rapid and Sensitive Formaldehyde Detection at Room Temperature Using a ZIF-8/MWCNT Nanocomposite. ACS Omega 2020, 5, 4395–4402. [Google Scholar] [CrossRef] [Green Version]
  231. Ahmad, Z.; Naseem; Manzoor, S.; Talib, M.; Islam, S.S.; Mishra, P. Self-standing MWCNTs based gas sensor for detection of environmental limit of CO2. Mater. Sci. Eng. B 2020, 255, 114528. [Google Scholar] [CrossRef]
  232. Kohli, N.; Hastir, A.; Kumari, M.; Singh, R.C. Hydrothermally synthesized heterostructures of In2O3/MWCNT as acetone gas sensor. Sens. Actuators A Phys. 2020, 314, 112240. [Google Scholar] [CrossRef]
  233. Liu, B.; Liu, X.; Yuan, Z.; Jiang, Y.; Su, Y.; Ma, J.; Tai, H. A flexible NO2 gas sensor based on polypyrrole/nitrogen-doped multiwall carbon nanotube operating at room temperature. Sens. Actuators B Chem. 2019, 295, 86–92. [Google Scholar] [CrossRef]
  234. Zhang, D.; Wu, Z.; Li, P.; Zong, X.; Dong, G.; Zhang, Y. Facile fabrication of polyaniline/multi-walled carbon nanotubes/molybdenum disulfide ternary nanocomposite and its high-performance ammonia-sensing at room temperature. Sens. Actuators B Chem. 2018, 258, 895–905. [Google Scholar] [CrossRef]
  235. Maity, D.; Rajavel, K.; Kumar, R.T.R. Polyvinyl alcohol wrapped multiwall carbon nanotube (MWCNTs) network on fabrics for wearable room temperature ethanol sensor. Sens. Actuators B Chem. 2018, 261, 297–306. [Google Scholar] [CrossRef]
  236. Abdulla, S.; Mathew, T.L.; Pullithadathil, B. Highly sensitive, room temperature gas sensor based on polyaniline-multiwalled carbon nanotubes (PANI/MWCNTs) nanocomposite for trace-level ammonia detection. Sens. Actuators B Chem. 2015, 221, 1523–1534. [Google Scholar] [CrossRef]
  237. Chen, N.; Li, Q.; Li, Y.; Deng, D.; Xiao, X.; Wang, Y. Facile synthesis and gas sensing performances based on nickel oxide nanoparticles/multi-wall carbon nanotube composite. J. Mater. Sci. Mater. Electron. 2015, 26, 8240–8248. [Google Scholar] [CrossRef]
  238. Kumar, V.; Tyagi, P.K. Potential application of multi-walled carbon nanotubes/activated carbon/bamboo charcoal for efficient alcohol sensing. J. Alloys Compd. 2018, 767, 215–222. [Google Scholar] [CrossRef]
  239. Lee, M.; Jang, Y.; Oh, J.; Jung, J. A silver-nanoparticle-embedded activated-carbon–PEDOT: PSS composite conductor for enhancing electrical conductivity of a gas sensor module. Dig. J. Nanomater. Biostruct. 2018, 13, 527–534. [Google Scholar]
  240. Mirzaie, M.; Talebizadeh, A.R.; Hashemipour, H. Mathematical modeling and experimental study of VOC adsorption by Pistachio shell–based activated carbon. Environ. Sci. Pollut. Res. 2021, 28, 3737–3747. [Google Scholar] [CrossRef]
  241. Oh, K.-J.; Park, D.-W.; Kim, S.-S.; Park, S.-W. Breakthrough data analysis of adsorption of volatile organic compounds on granular activated carbon. Korean J. Chem. Eng. 2010, 27, 632–638. [Google Scholar] [CrossRef]
  242. Li, M.; Lu, B.; Ke, Q.-F.; Guo, Y.-J.; Guo, Y.-P. Synergetic effect between adsorption and photodegradation on nanostructured TiO2/activated carbon fiber felt porous composites for toluene removal. J. Hazard. Mater. 2017, 333, 88–98. [Google Scholar] [CrossRef]
  243. Adhikari, B.; Majumdar, S. Polymers in sensor applications. Prog. Polym. Sci. 2004, 29, 699–766. [Google Scholar] [CrossRef]
  244. Swargiary, K.; Jolivot, R.; Mohammed, W.S. Demonstration of a Polymer-Based Single Step Waveguide by 3D Printing Digital Light Processing Technology for Isopropanol Alcohol-Concentration Sensor. Photonic Sens. 2022, 12, 10–22. [Google Scholar] [CrossRef]
  245. Pathak, A.K.; Singh, V.K. A wide range and highly sensitive optical fiber pH sensor using polyacrylamide hydrogel. Opt. Fiber Technol. 2017, 39, 43–48. [Google Scholar] [CrossRef]
  246. Viphavakit, C.; Komodromos, M.; Themistos, C.; Mohammed, W.S.; Kalli, K.; Azizur Rahman, B.M. Optimization of a horizontal slot waveguide biosensor to detect DNA hybridization. Appl. Opt. 2015, 54, 4881. [Google Scholar] [CrossRef] [PubMed]
  247. Victor, A.; Ribeiro, J.; Araújo, F.S. Study of PDMS characterization and its applications in biomedicine: A review. J. Mech. Eng. Biomech. 2019, 4, 1–9. [Google Scholar] [CrossRef]
  248. Raj, M.K.; Chakraborty, S. PDMS microfluidics: A mini review. J. Appl. Polym. Sci. 2020, 137, 48958. [Google Scholar] [CrossRef]
  249. Sidek, O.; Bin Afzal, M.H. A review paper on fiber-optic sensors and application of PDMS materials for enhanced performance. In Proceedings of the 2011 IEEE Symposium on Business, Engineering and Industrial Applications (ISBEIA), Langkawi, Malaysia, 25–28 September 2011; IEEE: Piscataway, NJ, USA, 2011; pp. 458–463. [Google Scholar]
  250. Reddy, K.; Guo, Y.; Liu, J.; Lee, W.; Khaing Oo, M.K.; Fan, X. Rapid, sensitive, and multiplexed on-chip optical sensors for micro-gas chromatography. Lab Chip 2012, 12, 901. [Google Scholar] [CrossRef]
  251. Ning, X.; Yang, J.; Zhao, C.L.; Chan, C.C. PDMS-coated fiber volatile organic compounds sensors. Appl. Opt. 2016, 55, 3543. [Google Scholar] [CrossRef]
  252. Kacik, D.; Martincek, I. Toluene optical fibre sensor based on air microcavity in PDMS. Opt. Fiber Technol. 2017, 34, 70–73. [Google Scholar] [CrossRef]
  253. Zhao, C.; Han, F.; Li, Y.; Mao, B.; Kang, J.; Shen, C.; Dong, X. Volatile Organic Compound Sensor Based on PDMS Coated Fabry–Perot Interferometer with Vernier Effect. IEEE Sens. J. 2019, 19, 4443–4450. [Google Scholar] [CrossRef]
  254. Sappati, K.K.; Bhadra, S. Printed Acoustic Sensor for Low Concentration Volatile Organic Compound Monitoring. IEEE Sens. J. 2021, 21, 9808–9818. [Google Scholar] [CrossRef]
  255. Snook, G.A.; Kao, P.; Best, A.S. Conducting-polymer-based supercapacitor devices and electrodes. J. Power Sources 2011, 196, 1–12. [Google Scholar] [CrossRef]
  256. Miah, M.R.; Yang, M.; Khandaker, S.; Bashar, M.M.; Alsukaibi, A.K.D.; Hassan, H.M.A.; Znad, H.; Awual, M.R. Polypyrrole-based sensors for volatile organic compounds (VOCs) sensing and capturing: A comprehensive review. Sens. Actuators A Phys. 2022, 347, 113933. [Google Scholar] [CrossRef]
  257. Das, M.; Sarkar, D. Development of room temperature ethanol sensor from polypyrrole (PPy) embedded in polyvinyl alcohol (PVA) matrix. Polym. Bull. 2018, 75, 3109–3125. [Google Scholar] [CrossRef]
  258. Li, Y.; Ban, H.; Yang, M. Highly sensitive NH3 gas sensors based on novel polypyrrole-coated SnO2 nanosheet nanocomposites. Sens. Actuators B Chem. 2016, 224, 449–457. [Google Scholar] [CrossRef]
  259. Lin, Q.; Li, Y.; Yang, M. Tin oxide/graphene composite fabricated via a hydrothermal method for gas sensors working at room temperature. Sens. Actuators B Chem. 2012, 173, 139–147. [Google Scholar] [CrossRef]
  260. Lu, G.; Ocola, L.E.; Chen, J. Room-Temperature Gas Sensing Based on Electron Transfer between Discrete Tin Oxide Nanocrystals and Multiwalled Carbon Nanotubes. Adv. Mater. 2009, 21, 2487–2491. [Google Scholar] [CrossRef]
  261. Wu, B.; Chen, L.; Zhao, C.; Chan, C.C. Polypyrrole-Coated Polarization Maintaining Fiber-Based Vernier Effect for Isopropanol Measurement. J. Light. Technol. 2019, 37, 2768–2772. [Google Scholar] [CrossRef]
  262. Šetka, M.; Bahos, F.A.; Chmela, O.; Matatagui, D.; Gràcia, I.; Drbohlavová, J.; Vallejos, S. Cadmium telluride/polypyrrole nanocomposite based Love wave sensors highly sensitive to acetone at room temperature. Sens. Actuators B Chem. 2020, 321, 128573. [Google Scholar] [CrossRef]
  263. Shoeb, M.; Mobin, M.; Ahmad, S.; Naqvi, A.H. Facile synthesis of polypyrrole coated graphene Gr/Ag–Ag2O/PPy nanocomposites for a rapid and selective response towards ammonia sensing at room temperature. J. Sci. Adv. Mater. Devices 2021, 6, 223–233. [Google Scholar] [CrossRef]
  264. Byeon, J.-H.; Kim, J.-S.; Kang, H.-K.; Kang, S.; Kim, J.-Y. Acetone Gas Sensor Based on SWCNT/Polypyrrole/Phenyllactic Acid Nanocomposite with High Sensitivity and Humidity Stability. Biosensors 2022, 12, 354. [Google Scholar] [CrossRef]
  265. Tripathi, A.; Mishra, S.K.; Bahadur, I.; Shukla, R.K. Optical properties of regiorandom polythiophene/Al2O3 nanocomposites and their application to ammonia gas sensing. J. Mater. Sci. Mater. Electron. 2015, 26, 7421–7430. [Google Scholar] [CrossRef]
  266. Bai, S.; Guo, J.; Sun, J.; Tang, P.; Chen, A.; Luo, R.; Li, D. Enhancement of NO2-Sensing Performance at Room Temperature by Graphene-Modified Polythiophene. Ind. Eng. Chem. Res. 2016, 55, 5788–5794. [Google Scholar] [CrossRef]
  267. Husain, A.; Ahmad, S.; Mohammad, F. Thermally stable and highly sensitive ethene gas sensor based on polythiophene/zirconium oxide nanocomposites. Mater. Today Commun. 2019, 20, 100574. [Google Scholar] [CrossRef]
  268. Husain, A.; Shariq, M.U.; Mohammad, F. DC electrical conductivity and liquefied petroleum gas sensing application of polythiophene/zinc oxide nanocomposite. Materialia 2020, 9, 100599. [Google Scholar] [CrossRef]
  269. Husain, A.; Ahmad, S.; Mohammad, F. Electrical conductivity and alcohol sensing studies on polythiophene/tin oxide nanocomposites. J. Sci. Adv. Mater. Devices 2020, 5, 84–94. [Google Scholar] [CrossRef]
  270. Husain, A.; Ahmad, S.; Mohammad, F. Synthesis, characterisation and ethanol sensing application of polythiophene/graphene nanocomposite. Mater. Chem. Phys. 2020, 239, 122324. [Google Scholar] [CrossRef]
  271. Fischer, E. Einfluss der Configuration auf die Wirkung der Enzyme. Ber. Dtsch. Chem. Ges. 1894, 27, 2985–2993. [Google Scholar] [CrossRef] [Green Version]
  272. Andersson, L.; Sellergren, B.; Mosbach, K. Imprinting of amino acid derivatives in macroporous polymers. Tetrahedron Lett. 1984, 25, 5211–5214. [Google Scholar] [CrossRef]
  273. GWulff, G.; Sarhan, A.; Sarhan, A.-W.; Sarhan, H. Use of Polymers with Enzyme-Analogous Structures for the Resolution of Racemates. Angew. Chem. Int. Ed. 1972, 11, 341–344. [Google Scholar]
  274. Wulff, G.; Oberkobusch, D.; Minárik, M. Enzyme-analogue built polymers, 18 chiral cavities in polymer layers coated on wide-pore silica. React. Polym. Ion Exch. Sorbents 1985, 3, 261–275. [Google Scholar] [CrossRef]
  275. BelBruno, J.J. Molecularly Imprinted Polymers. Chem. Rev. 2019, 119, 94–119. [Google Scholar] [CrossRef] [PubMed]
  276. Zhang, Y.; Zhang, J.; Liu, Q. Gas Sensors Based on Molecular Imprinting Technology. Sensors 2017, 17, 1567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Hua, Y.; Ahmadi, Y.; Kim, K.-H. Molecularly imprinted polymers for sensing gaseous volatile organic compounds: Opportunities and challenges. Environ. Pollut. 2022, 311, 119931. [Google Scholar] [CrossRef] [PubMed]
  278. Carballido, L.; Karbowiak, T.; Cayot, P.; Gerometta, M.; Sok, N.; Bou-Maroun, E. Applications of molecularly imprinted polymers and perspectives for their use as food quality trackers. Chem 2022, 8, 2330–2341. [Google Scholar] [CrossRef]
  279. Mustafa, G.; Lieberzeit, P.A. Molecularly imprinted polymer–Ag2S nanoparticle composites for sensing volatile organics. RSC Adv. 2014, 4, 12723–12728. [Google Scholar] [CrossRef] [Green Version]
  280. Hussain, M.; Kotova, K.; Lieberzeit, P. Molecularly Imprinted Polymer Nanoparticles for Formaldehyde Sensing with QCM. Sensors 2016, 16, 1011. [Google Scholar] [CrossRef] [Green Version]
  281. Tang, X.; Raskin, J.-P.; Lahem, D.; Krumpmann, A.; Decroly, A.; Debliquy, M. A Formaldehyde Sensor Based on Molecularly-Imprinted Polymer on a TiO2 Nanotube Array. Sensors 2017, 17, 675. [Google Scholar] [CrossRef] [Green Version]
  282. Abdelghani, R.; Shokry Hassan, H.; Morsi, I.; Kashyout, A.B. Nano-architecture of highly sensitive SnO2–based gas sensors for acetone and ammonia using molecular imprinting technique. Sens. Actuators B Chem. 2019, 297, 126668. [Google Scholar] [CrossRef]
  283. Pathak, A.K.; Limprapassorn, P.; Kongruttanachok, N.; Viphavakit, C. Molecularly Imprinted Polymer-Based Optical Sensor for Isopropanol Vapor. J. Sens. Actuator Netw. 2022, 11, 46. [Google Scholar] [CrossRef]
  284. Mousazadeh, M.; Jahangiri-Manesh, A.; Nikkhah, M.; Abbasian, S.; Moshaii, A.; Masroor, M.J.; Norouzi, P. Detection of hexanal gas as a volatile organic compound cancer biomarker using a nanocomposite of gold nanoparticles and selective polymers. J. Electroanal. Chem. 2022, 905, 115962. [Google Scholar] [CrossRef]
  285. Queralto, N.; Berliner, A.N.; Goldsmith, B.; Martino, R.; Rhodes, P.; Lim, S.H. Detecting cancer by breath volatile organic compound analysis: A review of array-based sensors. J. Breath Res. 2014, 8, 027112. [Google Scholar] [CrossRef] [PubMed]
  286. Neto, J.d.R.M.; Santos, W.d.J.R.; Lima, P.R.; Tanaka, S.M.C.N.; Tanaka, A.A.; Kubota, L.T. A hemin-based molecularly imprinted polymer (MIP) grafted onto a glassy carbon electrode as a selective sensor for 4-aminophenol amperometric. Sens. Actuators B Chem. 2011, 152, 220–225. [Google Scholar] [CrossRef] [Green Version]
  287. St-Gelais, R.; Mackey, G.; Saunders, J.; Zhou, J.; Leblanc-Hotte, A.; Poulin, A.; Barnes, J.A.; Loock, H.-P.; Brown, R.S.; Peter, Y.-A. Gas sensing using polymer-functionalized deformable Fabry–Perot interferometers. Sens. Actuators B Chem. 2013, 182, 45–52. [Google Scholar] [CrossRef] [Green Version]
  288. Reddy, K.; Guo, Y.; Liu, J.; Lee, W.; Khaing Oo, M.K.; Fan, X. On-chip Fabry–Pérot interferometric sensors for micro-gas chromatography detection. Sens. Actuators B Chem. 2011, 159, 60–65. [Google Scholar] [CrossRef]
  289. Martínez-Hipatl, C.; Muñoz-Aguirre, S.; Beltrán-Pérez, G.; Castillo-Mixcóatl, J.; Rivera-De la Rosa, J. Detection of volatile organic compounds by an interferometric sensor. Sens. Actuators B Chem. 2010, 147, 37–42. [Google Scholar] [CrossRef]
  290. Chakraborty, N.; Mondal, S. Chemiresistive NH3 detection at sub-zero temperatures by polypyrrole- loaded Sn1−xSbxO2 nanocubes. Mater. Horiz. 2022, 9, 1750–1762. [Google Scholar] [CrossRef]
  291. Xu, Y.; Lee, H.; Hu, Y.; Huang, J.; Kim, S.; Yun, M. Detection and Identification of Breast Cancer Volatile Organic Compounds Biomarkers Using Highly-Sensitive Single Nanowire Array on a Chip. J. Biomed. Nanotechnol. 2013, 9, 1164–1172. [Google Scholar] [CrossRef] [Green Version]
  292. Hosono, K.; Matsubara, I.; Murayama, N.; Woosuck, S.; Izu, N. Synthesis of Polypyrrole/MoO3 Hybrid Thin Films and Their Volatile Organic Compound Gas-Sensing Properties. Chem. Mater. 2005, 17, 349–354. [Google Scholar] [CrossRef]
  293. Campos, M.; Simões, F.R.; Pereira, E.C. Influence of methane in the electrical properties of polypyrrole films doped with dodecylbenzene sulfonic acid. Sens. Actuators B Chem. 2007, 125, 158–166. [Google Scholar] [CrossRef]
  294. Lee, J.-S.; Yoon, N.-R.; Kang, B.-H.; Lee, S.-W.; Gopalan, S.-A.; Kim, S.-W.; Lee, S.-H.; Kwon, D.-H.; Kang, S.-W. Au-Polypyrrole Framework Nanostructures for Improved Localized Surface Plasmon Resonance Volatile Organic Compounds Gas Sensing. J. Nanosci. Nanotechnol. 2015, 15, 7738–7742. [Google Scholar] [CrossRef]
  295. Liu, X.; Zheng, W.; Kumar, R.; Kumar, M.; Zhang, J. Conducting polymer-based nanostructures for gas sensors. Coord. Chem. Rev. 2022, 462, 214517. [Google Scholar] [CrossRef]
  296. Husain, A. Electrical conductivity based ammonia, methanol and acetone vapour sensing studies on newly synthesized polythiophene/molybdenum oxide nanocomposite. J. Sci. Adv. Mater. Devices 2021, 6, 528–537. [Google Scholar] [CrossRef]
  297. Kong, F.; Wang, Y.; Zhang, J.; Xia, H.; Zhu, B.; Wang, Y.; Wang, S.; Wu, S. The preparation and gas sensitivity study of polythiophene/SnO2 composites. Mater. Sci. Eng. B 2008, 150, 6–11. [Google Scholar] [CrossRef]
  298. Huang, J.; Kang, Y.; Yang, T.; Wang, Y.; Wang, S. Preparation of polythiophene/WO3 organic-inorganic hybrids and their gas sensing properties for NO2 detection at low temperature. J. Nat. Gas Chem. 2011, 20, 403–407. [Google Scholar] [CrossRef]
  299. Jahangiri-Manesh, A.; Mousazadeh, M.; Nikkhah, M. Fabrication of chemiresistive nanosensor using molecularly imprinted polymers for acetone detection in gaseous state. Iran. Polym. J. 2022, 31, 883–891. [Google Scholar] [CrossRef]
  300. Rong, Q.; Zhang, Y.; Wang, C.; Zhu, Z.; Zhang, J.; Liu, Q. A high selective methanol gas sensor based on molecular imprinted Ag-LaFeO3 fibers. Sci. Rep. 2017, 7, 12110. [Google Scholar] [CrossRef] [Green Version]
  301. Antwi-Boampong, S.; Peng, J.S.; Carlan, J.; BelBruno, J.J. A Molecularly Imprinted Fluoral-P/Polyaniline Double Layer Sensor System for Selective Sensing of Formaldehyde. IEEE Sens. J. 2014, 14, 1490–1498. [Google Scholar] [CrossRef]
  302. Fu, Y.; Finklea, H.O. Quartz Crystal Microbalance Sensor for Organic Vapor Detection Based on Molecularly Imprinted Polymers. Anal. Chem. 2003, 75, 5387–5393. [Google Scholar] [CrossRef]
  303. Jha, S.K.; Hayashi, K. A quick responding quartz crystal microbalance sensor array based on molecular imprinted polyacrylic acids coating for selective identification of aldehydes in body odor. Talanta 2015, 134, 105–119. [Google Scholar] [CrossRef]
  304. Hirayama, K.; Sakai, Y.; Kameoka, K.; Noda, K.; Naganawa, R. Preparation of a sensor device with specific recognition sites for acetaldehyde by molecular imprinting technique. Sens. Actuators B Chem. 2002, 86, 20–25. [Google Scholar] [CrossRef]
  305. Linnane, E.; Haddad, S.; Melle, F.; Mei, Z.; Fairen-Jimenez, D. The uptake of metal–organic frameworks: A journey into the cell. Chem. Soc. Rev. 2022, 51, 6065–6086. [Google Scholar] [CrossRef] [PubMed]
  306. Furukawa, H.; Müller, U.; Yaghi, O.M. “Heterogeneity within Order” in Metal-Organic Frameworks. Angew. Chem. Int. Ed. 2015, 54, 3417–3430. [Google Scholar] [CrossRef] [PubMed]
  307. Tayfuroglu, O.; Kocak, A.; Zorlu, Y. A neural network potential for the IRMOF series and its application for thermal and mechanical behaviors. Phys. Chem. Chem. Phys. 2022, 24, 11882–11897. [Google Scholar] [CrossRef] [PubMed]
  308. Thomas, K.M. Adsorption and desorption of hydrogen on metal–organic framework materials for storage applications: Comparison with other nanoporous materials. Dalt. Trans. 2009, 9, 1487–1505. [Google Scholar] [CrossRef] [PubMed]
  309. Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science 2005, 309, 2040–2042. [Google Scholar] [CrossRef] [PubMed]
  310. Yin, T.; Meng, X.; Wang, S.; Yao, X.; Liu, N.; Shi, L. Study on the adsorption of low-concentration VOCs on zeolite composites based on chemisorption of metal-oxides under dry and wet conditions. Sep. Purif. Technol. 2022, 280, 119634. [Google Scholar] [CrossRef]
  311. Huang, C.-Y.; Song, M.; Gu, Z.-Y.; Wang, H.-F.; Yan, X.-P. Probing the Adsorption Characteristic of Metal–Organic Framework MIL-101 for Volatile Organic Compounds by Quartz Crystal Microbalance. Environ. Sci. Technol. 2011, 45, 4490–4496. [Google Scholar] [CrossRef]
  312. Mousavi, S.; Zeinali, S. VOC s detection using resistive gas nanosensor based on MIL-101(Cr) as a metal organic framework. Sens. Actuators A Phys. 2022, 346, 113810. [Google Scholar] [CrossRef]
  313. Yin, D.; Hu, X.; Cai, M.; Wang, K.; Peng, H.; Bai, J.; Xv, Y.; Fu, T.; Dong, X.; Ni, J.; et al. Preparation, Characterization, and In Vitro Release of Curcumin-Loaded IRMOF-10 Nanoparticles and Investigation of Their Pro-Apoptotic Effects on Human Hepatoma HepG2 Cells. Molecules 2022, 27, 3940. [Google Scholar] [CrossRef]
  314. Yaghi, O.M.; Li, G.; Li, H. Selective binding and removal of guests in a microporous metal–organic framework. Nature 1995, 378, 703–706. [Google Scholar] [CrossRef]
  315. Saha, D.; Wei, Z.; Deng, S. Hydrogen adsorption equilibrium and kinetics in metal–organic framework (MOF-5) synthesized with DEF approach. Sep. Purif. Technol. 2009, 64, 280–287. [Google Scholar] [CrossRef]
  316. Yang, K.; Xue, F.; Sun, Q.; Yue, R.; Lin, D. Adsorption of volatile organic compounds by metal-organic frameworks MOF-177. J. Environ. Chem. Eng. 2013, 1, 713–718. [Google Scholar] [CrossRef]
  317. Elboughdiri, N. The use of natural zeolite to remove heavy metals Cu (II), Pb (II) and Cd (II), from industrial wastewater. Cogent Eng. 2020, 7, 1782623. [Google Scholar] [CrossRef]
  318. Ozansoy Kasap, B.; Marchenko, S.V.; Soldatkin, O.O.; Dzyadevych, S.V.; Akata Kurc, B. Biosensors Based on Nano-Gold/Zeolite-Modified Ion Selective Field-Effect Transistors for Creatinine Detection. Nanoscale Res. Lett. 2017, 12, 162. [Google Scholar] [CrossRef] [PubMed]
  319. Bhatia, S.; Abdullah, A.Z.; Wong, C.T. Adsorption of butyl acetate in air over silver-loaded Y and ZSM-5 zeolites: Experimental and modelling studies. J. Hazard. Mater. 2009, 163, 73–81. [Google Scholar] [CrossRef]
  320. Ning, X.; Zhao, C.L.; Yang, J.; Chan, C.C. Zeolite thin film-coated spherical end-face fiber sensors for detection of trace organic vapors. Opt. Commun. 2016, 364, 55–59. [Google Scholar] [CrossRef]
  321. Sun, Y.; Wang, J.; Du, H.; Li, X.; Wang, C.; Hou, T. Formaldehyde gas sensors based on SnO2/ZSM-5 zeolite composite nanofibers. J. Alloys Compd. 2021, 868, 159140. [Google Scholar] [CrossRef]
  322. Baimpos, T.; Gora, L.; Nikolakis, V.; Kouzoudis, D. Selective detection of hazardous VOCs using zeolite/Metglas composite sensors. Sens. Actuators A Phys. 2012, 186, 21–31. [Google Scholar] [CrossRef]
  323. Hu, Q.; Li, J.J.; Hao, Z.P.; Li, L.D.; Qiao, S.Z. Dynamic adsorption of volatile organic compounds on organofunctionalized SBA-15 materials. Chem. Eng. J. 2009, 149, 281–288. [Google Scholar] [CrossRef]
  324. Serrano, D.P.; Calleja, G.; Botas, J.A.; Gutierrez, F.J. Adsorption and Hydrophobic Properties of Mesostructured MCM-41 and SBA-15 Materials for Volatile Organic Compound Removal. Ind. Eng. Chem. Res. 2004, 43, 7010–7018. [Google Scholar] [CrossRef]
Figure 1. A summary of the sensing materials used for detecting VOC biomarkers.
Figure 1. A summary of the sensing materials used for detecting VOC biomarkers.
Biosensors 13 00114 g001
Figure 2. Block schematic diagram of a GC–MS.
Figure 2. Block schematic diagram of a GC–MS.
Biosensors 13 00114 g002
Figure 3. Block schematic diagram of a SIFT-MS.
Figure 3. Block schematic diagram of a SIFT-MS.
Biosensors 13 00114 g003
Figure 4. Block schematic diagram of a PTR-MS.
Figure 4. Block schematic diagram of a PTR-MS.
Biosensors 13 00114 g004
Figure 5. Scanning electron microscope (SEM) micrographs of ZnO nanorods on CSF with growth times of (a) 3 h; (b) 4 h; (c) 5 h; and (d) 7 h. (Reprinted/adapted from [78]).
Figure 5. Scanning electron microscope (SEM) micrographs of ZnO nanorods on CSF with growth times of (a) 3 h; (b) 4 h; (c) 5 h; and (d) 7 h. (Reprinted/adapted from [78]).
Biosensors 13 00114 g005
Figure 6. (a) SEM and (b) transmission electron microscopy (TEM) images of hollow SnO2/ZnS cubes. (c) SEM, (d) TEM, (e) high-resolution TEM images, and selected area diffraction (SAED) pattern of hollow SnO2/ZnO cubes (f). diffraction rings in the SAED pattern for different crystalline planes of SnO2 and ZnO (Reprinted/adapted from [88]).
Figure 6. (a) SEM and (b) transmission electron microscopy (TEM) images of hollow SnO2/ZnS cubes. (c) SEM, (d) TEM, (e) high-resolution TEM images, and selected area diffraction (SAED) pattern of hollow SnO2/ZnO cubes (f). diffraction rings in the SAED pattern for different crystalline planes of SnO2 and ZnO (Reprinted/adapted from [88]).
Biosensors 13 00114 g006
Figure 7. Field emission scanning electron microscopy (FESM) image of (a) NiO nanosheets [90] (b) NiO-75 (75% ethanol to distilled water ratio) thin nanosheets with porous structure ([89] (c) mesoporous NiO [91], and (d) 2.64% Sn-doped NiO [92].
Figure 7. Field emission scanning electron microscopy (FESM) image of (a) NiO nanosheets [90] (b) NiO-75 (75% ethanol to distilled water ratio) thin nanosheets with porous structure ([89] (c) mesoporous NiO [91], and (d) 2.64% Sn-doped NiO [92].
Biosensors 13 00114 g007
Figure 8. SEM image of Ni-doped α- MoO3 (a) 0 mol%, (b) 2.5 mol%, (c) 5 mol%, (d) 10 mol%. (Reprinted/adapted from [105]).
Figure 8. SEM image of Ni-doped α- MoO3 (a) 0 mol%, (b) 2.5 mol%, (c) 5 mol%, (d) 10 mol%. (Reprinted/adapted from [105]).
Biosensors 13 00114 g008
Figure 9. SEM image of MoO3 samples (a) sphere-like, (b) rose-like, (c) plate flower. (Reprinted/adapted from [108]).
Figure 9. SEM image of MoO3 samples (a) sphere-like, (b) rose-like, (c) plate flower. (Reprinted/adapted from [108]).
Biosensors 13 00114 g009
Figure 10. (a) Photograph and top view of the SEM images of TiOx nanostructure thin film on PI substrate after photosintering at 3.2 J/cm once on the PI substrate. Top view/cross-sectional view of the SEM images of TiO2 nanostructures on the rigid SiO2 glass substrate after (b) heat sintering at 500 °C and (c) photosintering at 3.2 J/cm one time. (Reprinted/adapted from [119]).
Figure 10. (a) Photograph and top view of the SEM images of TiOx nanostructure thin film on PI substrate after photosintering at 3.2 J/cm once on the PI substrate. Top view/cross-sectional view of the SEM images of TiO2 nanostructures on the rigid SiO2 glass substrate after (b) heat sintering at 500 °C and (c) photosintering at 3.2 J/cm one time. (Reprinted/adapted from [119]).
Biosensors 13 00114 g010
Figure 11. FESEM images of six anodically grown TiO2 samples (a,b) shows TiO2 nanotubes buried under fibrous coverage, (c) fibrous coverage is almost absent (d) nanotubular structures appeared (e) nanotubes were grown upward (f) free-standing ordered TiO2 nanotubes. (Reprinted/adapted from [119,135]).
Figure 11. FESEM images of six anodically grown TiO2 samples (a,b) shows TiO2 nanotubes buried under fibrous coverage, (c) fibrous coverage is almost absent (d) nanotubular structures appeared (e) nanotubes were grown upward (f) free-standing ordered TiO2 nanotubes. (Reprinted/adapted from [119,135]).
Biosensors 13 00114 g011
Figure 12. (a,c) The SEM images of C-TiO2 and D-TiO2, respectively; (b,d) the corresponding enlarged SEM images. The insets in (b,d) are the schematic illustration of different crystalline planes. (Reprinted/adapted from [150]).
Figure 12. (a,c) The SEM images of C-TiO2 and D-TiO2, respectively; (b,d) the corresponding enlarged SEM images. The insets in (b,d) are the schematic illustration of different crystalline planes. (Reprinted/adapted from [150]).
Biosensors 13 00114 g012
Figure 13. (a) XRD patterns, (bd) FESEM images, (e) TEM image and HRTEM image (insert), and (f) BET results of TiO2 flower-like microstructures. (Reprinted/adapted from [156]).
Figure 13. (a) XRD patterns, (bd) FESEM images, (e) TEM image and HRTEM image (insert), and (f) BET results of TiO2 flower-like microstructures. (Reprinted/adapted from [156]).
Biosensors 13 00114 g013
Figure 14. SEM image of (a) CuO hierarchical structure and (b) integration of rGO and the CuO hierarchical structure. (Reprinted/adapted from [192]).
Figure 14. SEM image of (a) CuO hierarchical structure and (b) integration of rGO and the CuO hierarchical structure. (Reprinted/adapted from [192]).
Biosensors 13 00114 g014
Figure 15. FESEM images of (a) MWCNTs on silicon dioxide; (b) PEG/MWCNTs on silicon dioxide; (c) the cross-section of MWCNTs and PEG composite layer. (d) Optical microscope images of MWCNTs and PEG composite layer. (Reprinted/adapted from [209]).
Figure 15. FESEM images of (a) MWCNTs on silicon dioxide; (b) PEG/MWCNTs on silicon dioxide; (c) the cross-section of MWCNTs and PEG composite layer. (d) Optical microscope images of MWCNTs and PEG composite layer. (Reprinted/adapted from [209]).
Biosensors 13 00114 g015
Figure 16. SEM images of two different samples (a) MoS2 only (b,c) MoS2 and MWCNTs hybrid. (Reprinted/adapted from [211]).
Figure 16. SEM images of two different samples (a) MoS2 only (b,c) MoS2 and MWCNTs hybrid. (Reprinted/adapted from [211]).
Biosensors 13 00114 g016
Figure 17. SEM image of palm shell-based AC prepared under optimum conditions. (Reprinted/adapted from [220]).
Figure 17. SEM image of palm shell-based AC prepared under optimum conditions. (Reprinted/adapted from [220]).
Biosensors 13 00114 g017
Figure 18. Microscope image of the fabricated microcavity inside PDMS. (Reprinted/adapted from [252]).
Figure 18. Microscope image of the fabricated microcavity inside PDMS. (Reprinted/adapted from [252]).
Biosensors 13 00114 g018
Figure 19. Schematic diagram of the fabricated FPI sensor coated with PDMS.
Figure 19. Schematic diagram of the fabricated FPI sensor coated with PDMS.
Biosensors 13 00114 g019
Figure 20. SEM images of SnO2/PPy nanocomposite prepared with different doping acids: (a) PSSA; (b) HCl; (c) TSA; and (d) CSA (polymerization time: 1 h). (Reprinted/adapted from [258]).
Figure 20. SEM images of SnO2/PPy nanocomposite prepared with different doping acids: (a) PSSA; (b) HCl; (c) TSA; and (d) CSA (polymerization time: 1 h). (Reprinted/adapted from [258]).
Biosensors 13 00114 g020
Figure 21. SEM images of (a) PPy, (c) Ag–Ag2O/PPy, and (e) Gr/Ag–Ag2O/PPy on the scale of 1 μm, respectively; (b,df) EDAX spectrum proves the elemental composition of as-synthesized, respectively (inset: presence of C, Ag, N, and O). (Corresponding inset) EDAX map indicates the presence C, Ag, N, and O). (Reprinted/adapted from [263]).
Figure 21. SEM images of (a) PPy, (c) Ag–Ag2O/PPy, and (e) Gr/Ag–Ag2O/PPy on the scale of 1 μm, respectively; (b,df) EDAX spectrum proves the elemental composition of as-synthesized, respectively (inset: presence of C, Ag, N, and O). (Corresponding inset) EDAX map indicates the presence C, Ag, N, and O). (Reprinted/adapted from [263]).
Biosensors 13 00114 g021
Figure 22. (a) The SEM and (b) transmission core–shell-shaped nanorods (sample). (Reprinted/adapted from [264]).
Figure 22. (a) The SEM and (b) transmission core–shell-shaped nanorods (sample). (Reprinted/adapted from [264]).
Biosensors 13 00114 g022
Figure 23. SEM images of (a) top view of TiO2 nanotube array, (b) cross-section view of TiO2 nanotube array, (c) thin layer of molecularly imprinted polypyrrole synthesized on a TiO2 nanotube array, (d) zoom of (c); and (e) thick polypyrrole film on the flat substrate (Reprinted/adapted from [281]).
Figure 23. SEM images of (a) top view of TiO2 nanotube array, (b) cross-section view of TiO2 nanotube array, (c) thin layer of molecularly imprinted polypyrrole synthesized on a TiO2 nanotube array, (d) zoom of (c); and (e) thick polypyrrole film on the flat substrate (Reprinted/adapted from [281]).
Biosensors 13 00114 g023
Figure 24. (a). Non-imprinted polymer (b). IPA-MIP surface. (Reprinted/adapted from [283]).
Figure 24. (a). Non-imprinted polymer (b). IPA-MIP surface. (Reprinted/adapted from [283]).
Biosensors 13 00114 g024
Figure 25. Zeolite film-coated fiber. (a). side view, (b). front view the, (c). cross-section, (d): enlarged view, (e). zeolite layer on fiber, and (f). spherical-shaped fiber end. (Reprinted/adapted from [320]).
Figure 25. Zeolite film-coated fiber. (a). side view, (b). front view the, (c). cross-section, (d): enlarged view, (e). zeolite layer on fiber, and (f). spherical-shaped fiber end. (Reprinted/adapted from [320]).
Biosensors 13 00114 g025
Table 1. Summary of most important VOCs and associated disorder along with their concentrations.
Table 1. Summary of most important VOCs and associated disorder along with their concentrations.
Exhaled VOCMolecular FormulaAssociated DisorderConcentration
(Healthy Person)
Concentration
(Diseased)
Refs.
IsopreneC5H8Diabetes, heart failure, lung cancer5.8 ppb to 274.9 ppb [19,20,21,38]
AcetoneC3H6ODiabetes,0.39 ppm to 0.85 ppmT1D: >2.2 ppm, typically >10 ppm[22,39]
AmmoniaNH3Liver dysfunction, kidney failure, hepatic encephalopathy, type-II Alzheimer, swelling of the brain0.25 ppm to 2.9 ppmNA[25,26,40]
MethaneCH4Oxidative stress, heart disease, breast cancer4 ppm to 10 ppm8 ppm to 50 ppm[41,42]
AldehydeR−CH=OLiver cancer, Alzheimer, Parkinson’s Disease13 ± 5 ppbNA[41,43]
Hydrogen sulfideH2SAsthma, oral, and dental health8 ppb to 16 ppbNA[41,44]
Nitric OxideNOAsthma, acute lung, injury, lung infection, lung cancer<35 ppbNA[45,46]
EthaneC2H6Scleroderma, cystic fibrosis0 ppb–12 ppbNA[47,48]
PentaneC5H12Myocardial infarction0.3 nmol/L to 0.8 nmol/LNA[49]
Carbon monoxideCOSmoking0.4 ppm to 0.8 ppm
(non-smokers)
2 ppm to 20 ppm[50,51]
8-isoprostane-Obstructive sleep apnea4 pg/mL to 5 pg/mL6.7 to 7.1 pg/mL[52]
Interleukin-6IL-6Non-small cell lung cancer3.3 pg/mL to 3.7 pg/mL9.3 pg/mL to 9.9 pg/mL[53]
Table 2. Advantage and limitations of classical detection techniques.
Table 2. Advantage and limitations of classical detection techniques.
TechniqueAdvantageLimitations
GC–MS
  • Identification of organic components through separating complex mixtures
  • Quantitative analysis
  • Trace detection of organic contamination as low as ppb level for liquid matrices and nanogram level for solid matrices
  • Non-volatile matrices require additional preparation such as extraction, outgassing, etc.
  • Target compounds must either be volatile
  • Atmospheric gases such as N2, CO2, O2, CO, Ar, H2O are challenging
PTR-MS
  • Quantitative monitoring
  • Absolute concentrations can be determined without previous calibration
  • Fast analysis, i.e., within milliseconds and online, which allow the real-time monitoring of exhaled breath
  • Only molecules with a proton affinity higher than water can be detected
  • Cannot provide as much information as GC–MS
SIFT-MS
  • Quantitative monitoring
  • Absolute concentrations can be determined without previous calibration
  • Fast analysis, i.e., within milliseconds and online, which allow the real-time monitoring of exhaled breath
  • This technique uses three reagent ions, i.e., NO+, H3O+, and O2+, which makes it suitable for detection of volatiles with lower proton affinities than water.
  • Cannot provide as much information as GC–MS
Table 3. Summary of all MO based VOC sensors.
Table 3. Summary of all MO based VOC sensors.
MaterialMorphologyVOCSensing RangeResponse TimeOperating TemperatureRefs.
ZnO based VOC sensors
ZnOFlower-like microstructureEthanol50 ppm12 sRT[161]
ZnFe2O4/ZnOFlower-like microstructuresAcetone50 ppm2 s250 °C[162]
ZnO/CuO on graphene substrateNanoflowerAmmonia5 ppm4.1 sRT[163]
ZnO QDsNanoparticlesIsoprene1 ppm8 s150 °C[164]
ZnO@CuOPorous hollow sphere nanoparticleH2S10 ppm33 sRT[165]
ZnO/Zn2SnO4MicroflowersCH4400 ppm10 s250 °C[166]
SnO2/ZnONanosphereFormaldehyde20 ppm27 s200 °C[167]
ZnO/AuPorous nanosheetsNO21 ppm1 sRT[168]
ZnOThin filmNO2100 ppm3 s200 °C[169]
NiO based VOC sensors
NiO:AlThin filmsHydrogen100 ppm388 sRT[170]
Methane100 ppm1737 s
NiO/ZnOhollow microspheresToluene100 ppm2 s300 °C[92]
NiO-In2O3nanosphereN-propanol100 ppm541 s210 °C[171]
Formic acid100 ppm215 s200 °C
Ethanol100 ppm263 s240 °C
Methanol100 ppm38 s300 °C
Isopropanol100 ppm139 s250 °C
Acetone100 ppm141 s250 °C
NiOMesoporousAmmonia0.4 ppm54 sRT[91]
MoO3 based VOC sensor
Pt-MoO3NanobeltsFormaldehyde200 ppm8.8 s27 °C[172]
α-MoO3NanorodMethanol500 pm32 s300 °C[173]
α-MoO3NanobeltsEthanol500 ppm14 s300 °C[174]
Zn-MoO3NanobeltsEthanol250 ppm121 s240 °C[175]
Fe2O3/MoO3NanobeltsXylene100 ppm4 s233.5 °C[176]
Ni-MoO3Nano-pomponXylene100 ppm1 s250 °C[177]
TiO2 based VOC sensors
TiO2 (rutile)Thin filmNH312.5–100 ppm2 min210 °C[178]
α-Fe2O3-TiO2Heterostructure Thin filmEthanol100 ppm46 s325 °C[179]
TiO2-WO3 Thin film (Nanoparticles)Xylene10 ppm410 s160 °C[180]
Ag/TiO2NanoparticlesAcetone100 ppm11 s275 °C[181]
Fe/TiO2nanotubesH2S50 ppm22 s100 °C[182]
TiO2-ZnO core shellnanorodsN-butanol100 ppm17 s200 °C[183]
F/TiO2nanosheetsAcetone400 ppm162 sRT[184]
TiO2–SnO2nanofibersFormaldehyde100 ppm13 sRT[185]
Table 4. Summary of all carbon-based VOC sensors.
Table 4. Summary of all carbon-based VOC sensors.
MaterialMorphologyVOCSensing RangeResponse timeOperating TemperatureRefs.
GO-based VOC sensors
ZnO/GONanorods/SheetAcetone, Benzene, Ethanol, Methanol10 ppm to 500 ppm8.79 s to 12.43 s200 °C to 500 °C[225]
rGo/NiOSheet/
Hierarchical Flower-like
Methanol61.51 ppm to 100 ppm16 s to 30 s (Graph Estimate)180 °C[191]
PEDOT: PSS/GONanowireAcetone, p-Xylene,
Ethanol, n-Hexane
5 ppm to 25 ppm8 s to 12 sRT[226]
CuO/rGOSheet/
Hierarchical
Ethanol5 ppm to 200 ppm23 s250 °C[192]
SnS2/rGO3D Porous
Structure
Nitrogen Dioxide0.5 ppm to 8 ppm153 sRT[194]
SnO2/GONanocomposite HeterojunctionEthanol, Acetone, Ethylbenzene1 ppm120 sRT (With UV light)
300 °C
[227]
SnO2/rGONanorods Array/3D
Hierarchical
Structure
Formaldehyde10 ppm to 50 ppm52 s50 °C to 150 °C[195]
ZnO-SnO2/rGOTernary Hybrid Nanocomposite with
p-n-n heterojunction
Acetone, Ethanol1–10 ppm10 s150 °C[198]
MWCNT-based VOC sensor
WO3/MWCNTsNanobrick thin filmAmmonia60 ppm512 sRT[228]
PDDA/MWCNTsThin filmCarbon monoxide20 ppm18 sRT[229]
ZIF-8/MWCNTsSharp facets crystals thin filmFormaldehyde5 ppm12 sRT[230]
Al2O3/MWCNTsPorous filmCarbon dioxide450 ppm53.7 sRT[231]
In2O3/MWCNTsNanobar SAED patternAcetone250 ppm9 s300 °C[232]
PPy/N-MWCNTsLoose granular structureNitrogen dioxide5 ppm65 sRT[233]
PANI/MWCNTs/MoS2Porous filmAmmonia0.25 ppm32 sRT[234]
PVA/MWCNTsEntangled porous, thin filmEthanol100 ppm24 ± 3 sRT[235]
PANI/MWCNTsEntangled porous, thin filmAmmonia2 ppm6 sRT[236]
NiO/MWCNTsAggregated flake-like structureEthanol100 ppm27 s180 °C[237]
AC-based VOC sensors
ACNon uniform sheet-like structureIsopropanol1 µL to 10 µL100 sRT[238]
Ag-AC-PEDOT: PSSNanoparticle composite conductorEthanol(a) 250 mTorr
(b) 500 mTorr
(c) 1000 mTorr
(a) 470 s
(b) 610 s
(c) 609 s
RT[239]
ACPistachio shell baseBenzene700 ppm and 1000 ppmNA30 °C to 45 °C[240]
Granular ACNAMethanol, Butanone, Benzene, Ethanol, n-Propanol, o-xylene, Toluene6000 ppmvNART[241]
TiO2/AC fiber feltPorous composite structureToluene<1150 ppmNART[242]
Table 5. Summary of all various polymers-based VOC sensors.
Table 5. Summary of all various polymers-based VOC sensors.
MaterialMorphologyVOCSensing RangeResponse TimeOperating TemperatureRefs.
PDMS-based VOC sensors
PDMSFilmToluene0.833 g.m−3 to 100 g.m−35 sRT[252]
PZT-PDMSComposite filmAcetic acid
Toluene
45 ppm to 100 ppm
0.2 ppm to 0.6 ppm
50 sRT[254]
PDMSFilmIsopropanol0 ppm to 500 ppm50 sRT[253]
PDMS and PDMS-PDPSFilmm-xylene and cyclohexane0 ppm to 1000 ppm
0 ppm to 3500 ppm
150 sRT[287]
PDMS and SU-8FilmToluene
acetone
0 ng to 20 ng
0 ng to 150 ng
0.9 s
0.5 s
RT[288]
PDMSFilmAlcohol, ethers, alkanes0 ppm to 24,000 ppm240 sRT[289]
Polypyrrole-based VOC sensors
PPy loaded Sn1−xSbxO2NanocubesAmmonia0 ppm to 20 ppm~4 sRT[290]
PPySingle nanowireHeptanal, Acetophenone, Isopropyl Myristate and 2-Propanol0.3% to 5%˂200 sNA[291]
PPy/MoO3FilmFormaldehyde and acetaldehyde100 ppm140 sRT[292]
PPy doped dodecylbenzene sulfonic acidFilmMethane20 ppmNART[293]
Au/PPyNanorodsAcetic acid, Benzene, toluene10 ppm to 100 ppm20 sNA[294]
PPyNanoparticlesMethanol, acetic acid, AcetonitrileNA˂1 sRT[295]
Polythiophene-based VOC sensor
PTh/MoO3Sheet/Rectangular particleAmmonia, methanol, acetone1 M concentrationNART[296]
PTh/Al2O3NanocompositeAmmonia25 ppm to 650 ppm25 s to 72 sRT[265]
rGO-PThHybrid FilmNitrogen oxide10 ppm26.36 sRT[266]
PTh/SnO2NanocompositeNitrogen oxide10 ppm to 200 ppmFew minRT[297]
PTh/ZrO2NanocompositeEtheneNANART[267]
PTh/WO3Hybrid filmNitrogen oxide100 ppmNA70 °C[298]
PTh/ZnONanocompositeLiquefied petroleum600 ppm to 2400 ppmNART[268]
PTh/GONanocompositeEthanol400 ppm to 2000 ppmVery shortRT[270]
MIP-based VOC sensors
MIP-Ag2SNanoparticles1-butanol100 ppm to 400 ppmNANA[279]
Au NPs/MIPsNanoparticlesAcetone50 ppm to 300 ppm3.7 ± 1 s180 °C[299]
Ag–LaFeO3NanoparticlesMethanol1 ppm to 5 ppm40 s to 47 s175 °C[300]
Polyaniline-fluoral-bilayer MIPFilmAmmonia, formaldehyde0 ppm to 1 ppmNANA[301]
MIPBulkHydroquinone, phenol, toluene, benzene, heptane50 ppm to 400 ppm5 s to 12 sNA[302]
MIPBulkHexanal, hexanoic acid, octanoic acid5 µL + 5 µL + 5 µL5 sNA[303]
MIPBulkAcetaldehydeNA˂30 sNA[304]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pathak, A.K.; Swargiary, K.; Kongsawang, N.; Jitpratak, P.; Ajchareeyasoontorn, N.; Udomkittivorakul, J.; Viphavakit, C. Recent Advances in Sensing Materials Targeting Clinical Volatile Organic Compound (VOC) Biomarkers: A Review. Biosensors 2023, 13, 114. https://doi.org/10.3390/bios13010114

AMA Style

Pathak AK, Swargiary K, Kongsawang N, Jitpratak P, Ajchareeyasoontorn N, Udomkittivorakul J, Viphavakit C. Recent Advances in Sensing Materials Targeting Clinical Volatile Organic Compound (VOC) Biomarkers: A Review. Biosensors. 2023; 13(1):114. https://doi.org/10.3390/bios13010114

Chicago/Turabian Style

Pathak, Akhilesh Kumar, Kankan Swargiary, Nuntaporn Kongsawang, Pannathorn Jitpratak, Noppasin Ajchareeyasoontorn, Jade Udomkittivorakul, and Charusluk Viphavakit. 2023. "Recent Advances in Sensing Materials Targeting Clinical Volatile Organic Compound (VOC) Biomarkers: A Review" Biosensors 13, no. 1: 114. https://doi.org/10.3390/bios13010114

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop