Next Article in Journal
SIR-DCGAN: An Attention-Guided Robust Watermarking Method for Remote Sensing Image Protection Using Deep Convolutional Generative Adversarial Networks
Previous Article in Journal
Swarm Intelligent Car-Following Model for Autonomous Vehicle Platoon Based on Particle Swarm Optimization Theory
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances and Applications of Odor Biosensors

1
Zhejiang Huanmao Auto-Control Technology Co., Ltd., Lianchuang Avenue No. 199, Wuchang, Yuhang, Hangzhou 311121, China
2
Zhejiang Ecological Environment Group, Lianchuang Avenue No. 199, Wuchang, Yuhang, Hangzhou 311121, China
3
Hangzhou Liqi Instrument Equipment Co., Ltd., Changsong Street No. 6, Cangqian, Yuhang, Hangzhou 311121, China
4
Institute of Integrated Research (llR), Institute of Science Tokyo, Suzukakedai Campus, Yokohama 226-0026, Japan
*
Author to whom correspondence should be addressed.
Electronics 2025, 14(9), 1852; https://doi.org/10.3390/electronics14091852
Submission received: 13 March 2025 / Revised: 27 April 2025 / Accepted: 29 April 2025 / Published: 1 May 2025
(This article belongs to the Special Issue Advanced Techniques in Biorobotics)

Abstract

:
Many odorants fall outside the sensing scope of the human olfactory system, yet they play quite important roles in our daily lives. Thus, numerous devices have been invented for qualitative or quantitative odor detection issues. Some analytical instruments, e.g., gas chromatography–mass spectrometry, are precise and reliable, but also expensive and bulky. Odor sensors with a smaller size and a lower cost play an important role in on-site rapid odor detection. The sensitivity and selectivity of these sensors are mainly determined by their sensing materials. Inspired by the powerful animal olfactory system, researchers extract diverse biological materials and combine them with transducers to form odor biosensors. In this paper, we introduce odor biosensors based on transducer types such as microelectrodes, fluorescence, surface plasmon resonance, field-effect transistor, quartz crystal micro-balance, etc. Then, we list several applications of odor biosensors, such as environmental monitoring, disease diagnosis, food quality control, and security. In addition, we analyze the future development of odor biosensors.

1. Introduction

Some researchers insist that humans can discriminate more than 1 trillion olfactory stimuli [1]. Despite having such a powerful and sophisticated olfactory system [2,3,4,5,6,7], there are still plenty of odorants that cannot be directly detected by the human sense of smell owing to sensitivity, safety, and detection range limitations [8,9,10]. To solve this problem, various devices have been proposed, including, for example, gas chromatography combined with mass spectrometry (GC-MS) [11,12,13], Raman spectroscopy [14,15,16], ion mobility spectrometry (IMS) [17,18], odor sensors [19,20,21], and so on. Among them, GC-MS, IMS, and Raman spectroscopy devices are bulky and require precise operation, while odor sensors with a small size, low power consumption, and a low cost are widely used in different scenarios [22,23]. The development of gas sensors could be traced back to the 1950s, when J. Eaton et al. measured minute air contaminants based on the surface tension change in a liquid drop in 1954 [24]. In 1961, R.W. Moncrieff designed an instrument for measuring and classifying odors and the film-coated thermistor was fixed on the Wheatstone bridge, wherein the odor carried by air induced a current signal [25]. A modern odor sensor system was invented in 1982 by Persaud and Dodd who used a metal oxide sensor array to mimic the discrimination of the mammalian olfactory system at a gross level [26]. Then, the commercially available electronic nose formed by conducting a polymer sensor array [27] and the electronic tongue for liquid-phase odor detection [28] were introduced, followed by more and more high-performance odor sensors being developed.
Odor sensors are classified based on their sensing mechanisms, i.e., transducer types, as follows: quartz crystal microbalance (QCM), surface plasmon resonance (SPR), field-effect transistor (FET), surface acoustic wave (SAW), microelectrode, and fluorescence [29,30]. These transducers themselves are not sensitive or selective towards odorants. Therefore, sensing materials are essential for capturing the target molecules, and utilizing proper sensing materials is key to enhancing sensor performance. Enlightened by the superior capability of the animal olfactory system, multiple biological materials—for instance, the aptamer, odorant-binding protein (OBP), olfactory receptor (OR) protein, olfactory sensory neuron (OSN), and olfactory epithelium—are integrated with transducers to create odor biosensors [31,32]. These biosensors inherit the characteristics of animal olfaction and can convert target odor information into a readable signal (Figure 1), thus being widely utilized in manifold aspects such as environmental monitoring, disease diagnosis, food quality control, and security [33]. On the other hand, some issues also exist in odor biosensors, e.g., low reproducibility, short lifespan, complex operation, and so on. Thus, many efforts have been made by researchers for better odor biosensors [34].
Several odor biosensors review papers have been published before. These papers mainly focused on biological sensing materials [35,36,37,38]. This review consists of three parts. First, we present some typical odor biosensors according to their sensing mechanisms and list corresponding instances. Then, we enumerate current application fields of odor biosensors. Finally, we illustrate the future perspectives of odor biosensors.

2. Odor Biosensors with Different Transducers

2.1. Microelectrode-Based Odor Biosensors

The microelectrode method measures the current [39], potential [40], or impedance variation [41] when the interaction between ligand molecules and the sensing material happens [42,43]. Thence, biological materials with spontaneous potential are suitable for current and potential measurements. Microelectrodes are fabricated via microfabrication techniques, either as planar arrays on a flat insulating substrate [44,45,46,47,48] or as needle-shaped probes [49,50] with microfabrication technology. When placed on a plane, microelectrodes always form an array, named a microelectrode array chip [51,52,53]. Multiple microelectrodes increase the data density and are capable of odor discrimination. The sensing materials, e.g., isolated OSN or olfactory epithelium, come into contact with the microelectrodes for signal recording (Figure 2a,b). As for needle-shaped microelectrodes, they could be inserted into a biological organism [54], cell complex [55], or single cell [39] (Figure 2c–e). Both invasive measurement and in vitro olfactory epithelium signal recording [56] shorten the lifetime of a biosensor. With the advancement of microfabrication technology, researchers have attempted to create an in vivo biosensor that could monitor real-time odorant-induced response for one week or even longer working periods [57,58,59]. This technique has also promoted the study of animal olfaction, including, for example, the quantitative evaluation of odor masking [60].
Measuring impedance spectroscopy variations could allow one to detect the combination of odorant molecules and sensing material. The measurement part consists of a counter electrode, a working electrode, and a reference electrode [61]. The sensing material could be an OR protein [62,63] or OBP [64,65], and recently, OBP has been preferred due to easy protein expression and purification. To obtain the desired OBP, a DNA sequence encoding the OBP is expressed in the host cell. After harvesting the host cells, the supernatant is handled for the OBP material [65]. The procedure for obtaining the OR protein is more complex because OR is embedded in the cell membrane. The OR gene is cloned into a vector, then transformed into a host cell. After culturing them for a sufficient amount of time, the cells are lysed, and the sample is centrifuged. The protein in the supernatant is purified with a His-Trap column and a Vivapsin column [63]. The immobilization of the protein on the bare electrode is not very stable; thus, N-(3-dimehylaminopropyl)-N′-ethylcarbodiimide hydrochlori (EDC) and N-hydrox-ysuccinimide (NHS) are introduced for achieving dense steady protein coverage [66,67,68] (Figure 2f). The expression of the protein is then confirmed by Western blotting [69,70,71]. The density of protein coverage on the transducer surface can be confirmed by atomic force microscopy (AFM) or X-ray photo electron spectroscopy (XPS) measurements [72]. The normalized impedance change is obtained by calculating the electron transfer resistance for evaluating the biosensor’s properties.
Figure 2. Microelectrode-based odor biosensors: (a) isolated OSN on a microelectrode chip; (b) olfactory epithelium stripped from rat nasal cavity; (c) needle-shaped microelectrode inserted into the olfactory bulb; (d) needle-shaped microelectrode inserted into a cell complex; (e) needle-shaped microelectrode inserted into a single cell; and (f) microelectrode covered with OBP for measuring the impedance spectroscopy. Regarding the above, (a) is reprinted from [73] with permission from Elsevier, (b) is reprinted from [44] with permission from Elsevier, (c) is reprinted from [54] with permission from Elsevier, (e) is reprinted from [55] with permission from John Wiley and Sons, and (f) is reprinted from [66].
Figure 2. Microelectrode-based odor biosensors: (a) isolated OSN on a microelectrode chip; (b) olfactory epithelium stripped from rat nasal cavity; (c) needle-shaped microelectrode inserted into the olfactory bulb; (d) needle-shaped microelectrode inserted into a cell complex; (e) needle-shaped microelectrode inserted into a single cell; and (f) microelectrode covered with OBP for measuring the impedance spectroscopy. Regarding the above, (a) is reprinted from [73] with permission from Elsevier, (b) is reprinted from [44] with permission from Elsevier, (c) is reprinted from [54] with permission from Elsevier, (e) is reprinted from [55] with permission from John Wiley and Sons, and (f) is reprinted from [66].
Electronics 14 01852 g002

2.2. Fluorescence-Based Odor Biosensors

In the fluorescent method [74,75,76], cells with fluorescent dye [77] or fluorescent protein [78,79] are employed as the sensing materials. Once the target ligand is captured by the OR on the cell membrane, an ion channel opens and then the ion influx emerges. The increase in intracellular ion concentration enhances the emission light intensity and is regarded as the biosensor response. To obtain the cell with fluorescent dye, OSN is preferred because it already has expressed OR protein and an ion channel on its membrane. The isolated OSNs are incubated with calcium-sensitive fluorescent dye (e.g., Fluo-4 AM or Fluo-8 AM) [80,81] for around 20 min and are then ready for subsequent experiments (Figure 3a). These OSNs usually form a sensing matrix for more robust biosensors [77,82]. The OR types in one animal vary from several hundred to several thousand [83,84,85], which means that the OSN matrix has a good latent detection capability. However, OSN also has several drawbacks: First, the proportion of OSNs in the olfactory epithelium is around 4–9% [77,82] and most cells trapped in the sensing matrix have no relationship to odor detection. Second, the OR type of OSN is unclear, and a standard odorant sample or single-cell reverse transcription polymerase chain reaction (RT-PCR) is necessary for confirming the expressed OR on OSN [82]. Third, the fluorescent dye is not always stable, and the OSN cannot be passaged for many generations.
On the other hand, researchers have focused on cells with fluorescent proteins in recent years [86,87,88,89]. In some of these studies, cells that could passage for multiple generations (for instance, Sf21, Hana3A, or HEK293) [87,90,91] were selected and expression vectors containing OR and fluorescent protein (usually sensitive to calcium ion or cyclic adenosine monophosphate) [29,87] genes were constructed [92]. Then, cells were transfected with expression vectors and cultured for around 1 week to establish a stable cell line (Figure 3b). The expression level of OR and fluorescent proteins was checked with PCR or Western blot methods. The OR type was determined by the operator. Hence, when a specific target ligand needed to be gauged, researchers could find and express its corresponding OR. After the cell line was established, the sensing material’s preparation in the following experiments was low-cost and simple. Meanwhile, the nanovesicle that secreted from the cell line also contained the same proteins. It could be exploited in fluorescent biosensors [93]. In addition, the fluorescent protein was also used to label a specific OSN for in vivo biosensors [58].

2.3. SPR-Based Odor Biosensors

Surface plasmon resonance is an optical phenomenon; its sensing mechanism comprises molecular interactions near the surface of nanoparticles shifting the SPR spectra peak [94]. When the size of the nanoparticle which forms the surface of the sensor is equal to or smaller than the wavelength of incident light, free electrons collectively oscillate, named localized surface plasmon polaritons [95,96,97,98]. This collective oscillation has its maximum absorbance at the resonant wavelength. Scientists can quantify the surface binding through monitoring the changes in resonance wavelength [99,100,101,102].
The sensing materials used in SPR-based biosensors are cells [100], membrane fragments [103], nanosomes [104], OBP [102], liposomes [105], and so on (Figure 4). The conformation changes in these sensing materials are captured by SPR-based biosensors. As cells per se can adhere to a gold surface, no additional substance is required [106]. OBP or OR proteins usually require EDC+NHS [103] or an antibody [105] to assist with the immobilization procedure. OR proteins need to be inserted into the membrane or liposome to maintain their seven transmembrane structures for normal function. On the contrary, OBP is directly secreted into the mucus for carrying hydrophobic odorant molecules. So, it is expressed, purified easily, and kept bioactive in vitro. Besides odorant detection, SPR-based odor biosensors can gauge the affinity of OBP with OR proteins [107]. This helps researchers gain a deeper understanding of the role of OBP in animal olfaction.

2.4. FET-Based Odor Biosensors

The FET sensor has gate, source, and drain electrodes. When the gate voltage is above the threshold voltage, carriers move in the channel between the source and drain electrodes, and the FET sensor can be used for odor detection when its threshold voltage changes [108,109,110,111]. Bare FETs have no selectivity towards odorants. Hence, their modification by sensing materials is necessary. The sensing mechanisms of FET-based biosensors is the modulation of drain source current.
In most FET biosensor application conditions, electrolyte or insulating layers cover the drain and source electrodes [112], and the gate electrode (usually Ag/AgCl electrode) is submerged in a buffer solution. To fabricate the desired structure, researchers first created a carbon nanotube [113,114,115,116,117] or graphene [72,118,119] pattern. In carbon nanotube FET (CNTFET), a photoresist is patterned on a silicon oxide wafer via standard photolithography [117]. Then, the patterned wafer is submerged in a CNT solution for CNT adhesion. Finally, the metal electrodes are formed by photolithography, thermal evaporation, and the lift-off process (Figure 5a). When it comes to the graphene FET, a single graphene layer is loaded on the silicon wafer using chemical vapor deposition and dry-transfer methods [120]. The graphene pattern is fabricated on a silicon oxide wafer using photolithography and reactive ion etching processes. Finally, the drain and source electrodes are constructed using the same methods in CNTFET (Figure 5b). Nanoparticles are also suitable for loading the sensing protein (Figure 5c) [121]. Proteins [116], peptides [72], and aptamers [122] are commonly used in FET-based biosensors for odor detection owing to their small size and high sensitivity. With the extended gate, cells can also be used for FET-based odor biosensors (Figure 5e) [123].

2.5. QCM-Based Odor Biosensors

QCM is a kind of resonant sensor based on the piezoelectric effect of quartz crystal [124,125,126], which almost linearly transforms the mass change absorbed on the QCM surface into its resonant frequency change [127]. There are many quartz cutting types—for example, AT cut, BT cut, and SC cut—depending on the quartz cutting angles [128,129,130,131]. Among them, the AT cut is the most used due to its good frequency–temperature characteristics. A typical QCM sensor has a sandwich structure, where the middle layer is piezoelectric quartz crystal, and its upper and lower surfaces are coated with thin (~100 nm) metal electrodes. Once an alternating current is applied on the electrodes, the quartz crystal undergoes a reciprocating motion. When the alternating electric field frequency matches the quartz crystal intrinsic frequency, resonance occurs, and the mechanical deformation and vibration intensity reach their zenith. In the resonant state, the substance attached to the electrodes results in the resonant frequency change, thereby enabling the sensing ability.
In the very beginning, QCM-based odor biosensors used the cell membrane fraction for odor detection [70,71,132]. HEK293 or E. coli was transfected with a plasmid containing the OR gene and then expressed OR protein on its plasma membrane. The membrane fraction was directly spread on the electrode or immobilized with the assistance of an aptamer (Figure 6a). Considering that the lipid in the membrane had no contribution towards odor detection, scientists extracted the OR protein or OBP as the sensing materials (Figure 6b) [133,134]. In order to further improve the effectiveness of QCM-based odor biosensors, a peptide mimicking OBP was introduced [135]. Researchers attempted different amino sequences in the peptide for more advanced sensing capabilities.

2.6. SAW-Based Odor Biosensors

QCM and SAW are both acoustic sensing methods. Their difference is the acoustic mode type: QCM uses bulk acoustic waves while SAW uses surface acoustic waves. Interdigital electrodes are fabricated on the piezoelectric substrate (e.g., LiNbO3, ZnO, AlN, SiO2) [136,137]. Once an alternating electrical signal is applied to the transmitter electrode, an acoustic wave is generated. This acoustic wave propagates along the interface of the substrate [138]. When it reaches the receiver electrode, the electrode’s characteristic size (for example, width) is an integer multiple of the acoustic wave’s half wavelength. The acoustic wave resonates in the interdigital channel and is converted into an electrical signal via the direct piezoelectric effect [139]. The gaps between the transmitter and receiver electrodes are covered with a sensing layer, where the target analyte is adsorbed. The mass loading of target molecules leads to variations in frequency, phase, or acoustic velocity, which are considered the sensor response [140,141]. The wave types of SAW devices are listed as follows: first, the Rayleigh wave (R-SAW), mainly used for gas-phase detection; second, the shear horizontal wave (SH-SAW), suitable for liquid-phase detection; and third, the Love wave (LW-SAW), similar to shear horizontal wave with slower speed and less attenuation [142]. By comparing their characteristics, it is not difficult to see that Love wave-type SAW sensors have the widest application prospects [143,144].
Similarly to QCM-based odor biosensors, the sensing materials in SAW-based odor biosensors are mainly antibodies [145], OR proteins [146], and OBP [147,148] (Figure 7). The sensing layer, which is thinner than that in QCM-based sensors, is preferable due to its attenuation. These proteins are first suspended in buffer and then added to SAW transducers for odor detection.
The comparison of different transducers in odor biosensors is shown in Table 1. The selection of transducers should consider specific measurement requirements.

2.7. Other Odor Biosensors

Besides the aforementioned biosensors, some other biosensors can also be used for odor detection.
LAPS has electrolyte, insulator, and semiconductor layers; its surface is flat and thus suitable for sensitive material loading [149]. When a laser beam irradiates the LAPS semiconductor layer, electron hole pairs are generated. Meanwhile, the bias voltage generates a depletion layer between the semiconductor and the insulation layer. When the electron hole pairs excited by light diffuse into the depletion layer, due to the electrical properties of the electrons and holes themselves, they are torn apart by the electric field between the semiconductor and insulating layer and move towards the two poles to form a current. At this point, if a modulated light source with a certain frequency irradiates the semiconductor layer, a phenomenon of alternating movement of electron hole pairs occurs, allowing for the detection of alternating photocurrent in the external circuit [150]. The interaction between the sensing layer and the target substance changes the original membrane potential balance of the sensitive layer, and with the continuous change in the concentration of the target substance to be measured, the membrane potential also changes accordingly, thus achieving the purpose of quantitative detection of the target substance [151]. In past studies, olfactory mucosa tissue was isolated from rat nasal cavity and the cilia receptor side was placed upwards on the LAPS [152,153]. Odorant-induced extracellular potential changes could be recorded by the LAPS. The OSN itself or OSN heterologously expressing OR could also be used for odor detection [154,155,156,157] (Figure 8a). To extend the sensing material lifetime and maintaining stable characteristics, HEK293 cells were utilized to replace OSN [158].
In another study, OR and Orco were inserted into a bilayer lipid membrane to imitate a cell [159]. When the OR captured its target ligand, the ion flow emerged, resulting in the current variation (Figure 8b). The Ag/AgCl electrodes at the bottom recorded the ionic current as a biosensor response [160]. The ion gate formed by OR/Orco could be substituted by a nanopore-forming membrane protein (e.g., alpha-hemolysin) [161]. In this condition, the aptamer carried the odorant molecules to trigger the current spike.

3. Odor Biosensor Application Fields

3.1. Environmental Monitoring

There are numerous volatile organic compounds (VOCs) in the ambient environment. Some are produced in nature while others are artificial odorants. These odorants are related to environmental pollution conditions and the state of survival and reproduction of animals and plants [162]. Hence, it is necessary to detect them precisely.
Actinomycetes’ and blue-green algae’s metabolism produces geosmin [163]. Geosmin has an earthy smell and is regarded as a contaminant in tap water [119,164] (Figure 9a). Termtanasombat et al. selected the Drosophila olfactory receptor Or56a, which is quite sensitive and selective to geosmin [165]. In that paper, the sensing cells that expressed diverse ORs only responded to their corresponding stimulus. In order to extend the biosensor detection range from the liquid phase into the gas phase, Deng et al. kept a thin liquid layer on the cells which allowed the target ligand to penetrate the buffer and maintained cell viability [166].
The substances produced by human activities are a major source of environmental pollution. Fujii et al. combined aptamers and nanopores together for sensing omethoate, a commonly used pesticide [161]. Agarose gel and lipid were alternately added into two similar small chambers, and then the nanopore protein and DNA aptamers were applied to the agarose gel. When the omethoate vapor was introduced, the aptamer captured the pesticide molecules and passed through the nanopore, resulting in a current spike (Figure 9b). The intensity of the current spike had a positive correlation with the vapor exposure time and showed no significant difference when other stimuli were applied, thereby proving the effectiveness of this biosensor.
Odor source localization was an important auxiliary technology for environmental monitoring [167,168]. Terutsuki et al. mounted silkmoth antennae on a portable electroantennogram and used it as the odor sniffer in a small drone [169]. The electroantennogram signal intensity was determined by the odor concentration, thereby guiding the drone to fly towards the odor source (Figure 9c). Although the odor source localized in this paper was pheromone bombykol, insect antennae can be modified through bioengineering to be sensitive to other odorants for achieving different odor source localization.
Figure 9. Odor biosensors applied in environmental monitoring: (a) OR embedded in nanovesicles for geosmin and 2-methylisoborneol detection, * p < 0.05, ** p < 0.01, *** p < 0.001 in the one-way ANOVA model; (b) DNA aptamer capturing an omethoate molecule and then passed it to the nanopore to trigger a long current spike, DNA aptamer before (i), during (ii), and after (iii) translocation through the nanopore; and (c) antenna from male silkmoth combined with a drone for odor source localization. (a) is reprinted from [164] with permission from Elsevier. (b) is reprinted from [161] with permission from the Royal Society of Chemistry. (c) is reprinted from [169].
Figure 9. Odor biosensors applied in environmental monitoring: (a) OR embedded in nanovesicles for geosmin and 2-methylisoborneol detection, * p < 0.05, ** p < 0.01, *** p < 0.001 in the one-way ANOVA model; (b) DNA aptamer capturing an omethoate molecule and then passed it to the nanopore to trigger a long current spike, DNA aptamer before (i), during (ii), and after (iii) translocation through the nanopore; and (c) antenna from male silkmoth combined with a drone for odor source localization. (a) is reprinted from [164] with permission from Elsevier. (b) is reprinted from [161] with permission from the Royal Society of Chemistry. (c) is reprinted from [169].
Electronics 14 01852 g009

3.2. Disease Diagnosis

The smells from human sweat [170], urine [23,171,172,173], and breath [174,175,176,177,178] indicate an individual’s health condition. The accurate analysis of these odors allows one to identify disease signs, reduces the pain of traditional invasive inspections, especially for chronic disease monitoring, cancer screening, or early warning of infectious diseases, provides dynamic data support for personalized diagnosis and treatment, reduces medical costs, and improves public health prevention and control efficiency.
Cho et al. detected 2-ethyl-1-hexanol from headspace gas in lung cancer and decided to use it as a cancer biomarker [93] (Figure 10a). They checked human OR libraries and found that human OR 4D11P was sensitive to this biomarker. To obtain long-term stable sensing capability, they heterologously expressed the OR on HEK293 cells and generated nanovesicles containing OR 4D11P. Finally, they confirmed the functionality of this cell line via 2-ethyl-1-hexanol stimulation and a calcium signaling assay. Other odorant molecules could also be lung cancer biomarkers.
To fully utilize the sensing ability of OSNs, Suzuki et al. fabricated a microchamber array and filled it with isolated OSNs [82]. The OSNs that responded to biomarkers were retrieved from the microchamber array and then subjected to single-cell RT-PCR (Figure 10b). The ORs from responsive OSNs were expressed on HEK293 cells and retained the same detection capabilities. Hu et al. developed a biosensor consisting of peptides and MXene [179]. This biosensor could detect fifteen odor molecules affiliated with five categories of alcohols, ketones, aldehydes, esters, and acids by pattern recognition algorithms (Figure 10c). Furthermore, the biosensor sensed the breath samples from healthy populations and patients with lung cancer, upper digestive tract cancer, and lower digestive tract cancer. With machine learning assistance, the detection accuracy for all patients was higher than 90%. This cost-effective and precise model is suitable for non-invasive tumor diagnosis.
Besides cancer-related biomarkers, Lee et al. invented a QCM-based biosensor for low concentrations of ammonia in human breath [180], as ammonia has a close relationship with the digestive system and kidney problems (Figure 10d). Acetone in the breath has a smell of rotten apples, which is closely related to diabetic ketoacidosis. In reference [175], an acetone bio-sniffer was developed to measure the exhaled breath acetone concentration and assess lipid metabolism. A fiber-optic bio-sniffer was manufactured by attaching a flow-chamber with nicotinamide adenine dinucleotide (NADH)-dependent secondary alcohol dehydrogenase (S-ADH)-immobilized film onto a fiber-optic NADH measurement system (Figure 10e). The excitation light source was a diode emitting 335 nm ultraviolet light. NADH was consumed by the enzymatic reaction of S-ADH, and the consumption was proportional to the acetone concentration. We believe that these biosensors will be widely used in the field of rapid, non-invasive, and low-cost disease detection in the near future.

3.3. Food Safety

Food safety is directly related to public health, social stability, and economic development. Unsafe food causes foodborne illnesses, ranging from mild abdominal pain and vomiting to organ damage and even death, particularly threatening vulnerable groups such as children and pregnant women. Frequent food safety incidents weaken consumer confidence and impact the food industry chain. Therefore, ensuring food safety is essential to maintain individual life rights, promote sustainable economic operation, and maintain social stability.
Just as people judge the degree of food spoilage by the smell of food, biosensors judge the state of food by detecting specific odors produced during food spoilage. Yang et al. produced nanodisks embedded with an OR produced in E. coli and then immobilized it on the floating electrodes of a carbon nanotube transistor [115]. Cadaverine as a death-associated odor was a sign of food spoilage (Figure 11a). Trace-amine-associated receptor 13c was employed to detect the target ligand. Study result indicated that this biosensor was sensitive to cadaverine stimulation as low as 10 pM. Calabrese et al. designed a non-Faradaic impedimetric biosensor for monitoring the presence of VOCs (1-octen-3-ol, trans-2-hexen-1-ol, and hexanal), involved in food spoilage [66]. This biosensor exploited pig OBP and could be used for measurements in 0.1 µM liquid and air samples.
Salmonella is a common foodborne pathogen that causes a series of symptoms such as typhoid or paratyphoid fever symptoms, bacterial blood symptoms, local infection symptoms, etc. In reference [181], researchers detected Salmonella contamination in packaged beef by measuring 3-methyl-1-butanol and 1-hexanol. The response signals from four QCMs were processed by principal component analysis and utilized for discriminating between these two odorants. Wang et al. immobilized silk fibroin microspheres onto lotus silk to form a high-surface-area mace structure and then functionalized it with Drosophila OBP-derived peptide [182]. The detection limit of this biosensor reached 1 ppt and sensed a 105 CFU concentration of Salmonella in ham. Similar detection effects can be found in references [133,183] (Figure 11b).
Besides bacterial contamination, antibiotic abuse results in residues in animal-derived foods, which, when accumulated in the human body via the food chain, cause ototoxicity, nephrotoxicity, and neuromuscular blockade. Cui et al. proposed an aptamer probe that contained an aptamer unit for recognizing kanamycin, a trigger unit for triggering catalytic hairpin assembly and a suppression unit for ensuring the stability of the aptamer probe [184]. The kanamycin concentration was reflected as a fluorescent intensity with a detection limit of 0.26 ng/mL.

3.4. Security

Explosive detection is essential for maintaining public safety. By accurately identifying and intercepting explosive devices or hazardous chemicals, it effectively prevents casualties, property damage, and social panic. In airports, stations, and large events, rapid and sensitive detection technologies can screen suspicious items in real time and block potential threats.
Explosive detection via odor biosensors started from whole animals [185]. Animals such as dogs, pigs, mice, and honeybees could perform specific actions based on some explosive odors after training [186], thereby alerting security personnel to investigate potential hazards [187]. To save time and economic costs in training, Gao et al. implanted a microelectrode array into mouse olfactory bulb [58]. Since the mouse OSN M72 was sensitive to trinitrotoluene (Figure 12a), they used green fluorescent protein to highlight the corresponding section in the olfactory bulb. Then, the electrode probe could precisely be inserted into the correct area. OBP, OR proteins, or major urinary proteins have been shown to be more effective than in vivo strategies [102,188,189] (Figure 12b). These proteins have thus been extracted and purified for the measurement of various trace explosives or explosive residues.

4. Future Prospects

Biosensors have been developed for over four decades since the concept was first proposed in the 1980s. Their detection performance has also significantly improved in terms of sensitivity, selectivity, response time, sensing range, and so on. The number of sensing elements in biosensors has increased to enhance their detection capability [39,88]. Suitable sensing equipment and signal processing methods have been proposed to improve the devices’ signal-to-noise ratio [190,191]. Controlling a thin liquid layer or utilizing OBP as a molecular transporter has achieved both liquid- and gas-phase odorant detection [55,116,192].
Despite the fact that so many efforts have been made, odor biosensors’ performance is still far from that of animal olfactory systems. The lifetime of odor biosensors is hours or days, and, while in vivo biosensors are a promising solution, they would fail due to protein-wrapped electrodes [59,60]. The responses from biosensors usually become smaller or even invisible under continuous odor stimulation, which is the main barrier to long-term repeatable odor quantification [193]. Although some calibration or compensation strategies have been proposed [194], the problem of response attenuation has not been fundamentally solved yet. Variation in environmental factors, e.g., temperature, humidity, and light intensity, also influences biosensors’ response. Therefore, commercially available biosensors that can be deployed in industrial sites are quite rare. Moreover, the preparation of biological sensing materials is complex and expensive. Due to these limitations, biosensors are still mainly studied in the laboratory, with few successful application cases.
As for the future development of odor biosensors, higher sensitivity and selectivity and shorter response times are major objectives pursued by scientists. With the deepening of studies on animal olfaction, researchers have explored the response repertoires of different ORs and OBP. A database with more complete response information is required, as it would help users determine the optimal sensing material based on the target odorant. Moreover, peptides derived from various proteins inherit the detection performance of the original proteins and are easy to integrate into various transducers, making them ideal materials for rapid detection kits. We believe that, with the advancement of these technologies, commercial biosensors will soon be launched, thereby promoting biosensor research as a whole.

5. Conclusions

Odor biosensors, as an interdisciplinary technology which combines biometric elements with physical and chemical transducers, have shown broad application prospects in the environmental monitoring, disease diagnosis, food safety, and security fields due to their high specificity, sensitivity, and real-time response capabilities. By synergistically designing biological components such as OR proteins, OBPs, enzymes, antibodies, OSNs, or aptamers with electrochemical, optical, piezoelectric, and other transducers, these biosensors can accurately capture trace target odors and convert them into quantifiable signals, providing efficient solutions for dynamic detection in complex scenarios. In the future, with the deep integration of nanomaterials, flexible electronics, and artificial intelligence technologies, odor biosensors will develop towards miniature, intelligence, and multi-omics applications, further promoting their innovative use in precision medicine, intelligent environmental governance, and real-time risk warning, becoming the pioneers of “artificial olfaction” in order to safeguard human health and safety.

Author Contributions

Conceptualization, H.D., H.Z. and T.N.; investigation, H.D., Z.C., P.F., L.T., H.Z. and T.N.; resources, H.Z.; writing—original draft preparation, H.D., Z.C., P.F. and L.T.; writing—review and editing, H.Z. and T.N.; project administration, Z.C. and H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

Authors Hongchao Deng, Lifeng Tian and Huijuan Zong were employed by the company Zhejiang Huanmao Auto-Control Technology Co., Ltd.; Zhangyu Chen was employed by the company Zhejiang Ecological Environment Group; Pengfei Feng was employed by the company Hangzhou Liqi Instrument Equipment Co., Ltd. The remaining author declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Bushdid, C.; Magnasco, M.O.; Vosshall, L.B.; Keller, A. Humans can discriminate more than 1 trillion olfactory stimuli. Science 2014, 343, 1370–1372. [Google Scholar] [CrossRef] [PubMed]
  2. Pevsner, J.; Reed, R.R.; Feinstein, P.G.; Snyder, S.H. Molecular cloning of odorant-binding protein: Member of a ligand carrier family. Science 1988, 241, 336–339. [Google Scholar] [CrossRef]
  3. Franco, R.; Garrigós, C.; Capó, T.; Serrano-Marín, J.; Rivas-Santisteban, R.; Lillo, J. Olfactory receptors in neural regeneration in the central nervous system. Neural Regen. Res. 2025, 20, 2480–2494. [Google Scholar] [CrossRef] [PubMed]
  4. Preuschhof, C.; Heekeren, H.R.; Taskin, B.; Schubert, T.; Villringer, A. Neural correlates of olfactory working memory in the human brain. J. Neurosci. 2006, 26, 13231–13239. [Google Scholar] [CrossRef] [PubMed]
  5. Naffaa, M.M. Neurogenesis dynamics in the olfactory bulb: Deciphering circuitry organization, function, and adaptive plasticity. Neural Regen. Res. 2025, 20, 1565–1581. [Google Scholar] [CrossRef]
  6. Keller, A.; Gerkin, R.C.; Guan, Y.; Dhurandhar, A.; Turu, G.; Szalai, B.; Mainland, J.D.; Ihara, Y.; Yu, C.W.; Meyer, P.; et al. Predicting human olfactory perception from chemical features of odor molecules. Science 2017, 355, 820–826. [Google Scholar] [CrossRef]
  7. Johnson, D.C. Airway mucus function and dysfunction. N. Engl. J. Med. 2011, 364, 978. [Google Scholar] [CrossRef]
  8. Schiefner, A.; Freier, R.; Eichinger, A.; Skerra, A. Crystal structure of the human odorant binding protein, OBPIIa. Proteins Struct. Funct. Bioinforma. 2015, 83, 1180–1184. [Google Scholar] [CrossRef]
  9. Nielsen, B.L.; Rampin, O.; Meunier, N.; Bombail, V. Behavioral responses to odors from other species: Introducing a complementary model of allelochemics involving vertebrates. Front. Neurosci. 2015, 9, 226. [Google Scholar] [CrossRef]
  10. German, P.F.; van der Poel, S.; Carraher, C.; Kralicek, A.V.; Newcomb, R.D. Insights into subunit interactions within the insect olfactory receptor complex using FRET. Insect Biochem. Mol. Biol. 2013, 43, 138–145. [Google Scholar] [CrossRef]
  11. Lee, M.; Kim, Y.S. Analysis of volatile and odor-active compounds in charcoal-grilled marinated beef using gas chromatography–mass spectrometry and gas chromatography–olfactometry. Food Sci. Biotechnol. 2025, 34, 1339–1349. [Google Scholar] [CrossRef] [PubMed]
  12. Feng, G.; Li, J.; Liu, J.; Tan, R. Examining the effects of processing techniques on the quality of hawk tea through liquid chromatography–tandem mass spectrometry and two-dimensional gas chromatography–time-of-flight mass spectrometry. Food Chem. 2025, 465, 142012. [Google Scholar] [CrossRef]
  13. Corvino, A.; Khomenko, I.; Betta, E.; Brigante, F.I.; Bontempo, L.; Biasioli, F.; Capozzi, V. Rapid Profiling of Volatile Organic Compounds Associated with Plant-Based Milks Versus Bovine Milk Using an Integrated PTR-ToF-MS and GC-MS Approach. Molecules 2025, 30, 761. [Google Scholar] [CrossRef]
  14. Fernández-Lodeiro, A.; Constantinou, M.; Panteli, C.; Agapiou, A.; Andreou, C. Breath Analysis via Surface Enhanced Raman Spectroscopy. ACS Sens. 2025, 10, 602–621. [Google Scholar] [CrossRef] [PubMed]
  15. Sobotka, A.; Krueger, T.D.; Solaris, J.; Chen, C.; Feng, Y.; Mayer, S.G.; Qian, Y.P.; Fang, C.; Qian, M.C. Developing nondestructive sensitive detection for smoke-exposure assessment in wine via advanced Raman spectroscopy. Food Chem. 2025, 475, 143327. [Google Scholar] [CrossRef]
  16. Ju, J.; Cong, S.; Hu, Y.; Xu, Z.; Ji, Z.; Xu, R.; Zhang, Z.; Cao, X. Rapid detection of nitenpyram in fruits using molecularly imprinted polymers coupled with SERS technology. J. Food Compos. Anal. 2025, 139, 107170. [Google Scholar] [CrossRef]
  17. Wang, Y.; Zhou, Z.; Chang, R.; Zhang, Z.; Xu, X.; Xu, Y.; Mao, J. New Method of Direct Sampling Gas Chromatography-Ion Mobility Spectrometry to Identify the Dynamic Retronasal Volatile Compounds in the Alcoholic Beverage and Their Release Behaviors: A Case Study on Huangjiu. J. Agric. Food Chem. 2025, 73, 4331–4341. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, J.; Wang, C.; Xie, Y.; Xu, P.; Xu, X.; Pang, L.; Lv, Z.; Lu, G. Physiological characteristics and volatile organic compound dynamics in bruising sweetpotatoes during short-term storage: Insights from GC–MS and GC-IMS. Food Chem. X 2025, 27, 102418. [Google Scholar] [CrossRef]
  19. Zhu, G.; Yang, Y.; Zhang, F.; Wei, J.; Tian, X.; Liu, L.; Ma, Z.; Zhang, G. Non-Invasive Monitoring and Differentiation of Aging Mice Treated with Goat Whey Powder by an Electronic Nose Coupled with Chemometric Methods. Sensors 2025, 25, 1496. [Google Scholar] [CrossRef]
  20. Frischmon, C.; Crosslin, J.; Burks, L.; Weckesser, B.; Hannigan, M.; Duderstadt, K. Detecting air pollution episodes and exploring their impacts using low-cost sensor data and simultaneous community symptom and odor reports. Environ. Res. Lett. 2025, 20, 044043. [Google Scholar] [CrossRef]
  21. Kong, R.; Shen, W.; Gao, Y.; Lv, D.; Ai, L.; Song, W.; Tan, R. Enhancing E-Nose Performance via Metal-Oxide Based MEMS Sensor Arrays Optimization and Feature Alignment for Drug Classification. Sensors 2025, 25, 1480. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, B.; Zhou, L.; Wang, X. Surface acoustic wave sensor for formaldehyde gas detection using the multi-source spray-deposited graphene/PMMA composite film. Front. Mater. 2023, 9, 1025903. [Google Scholar] [CrossRef]
  23. Arasaradnam, R.P.; Krishnamoorthy, A.; Hull, M.A.; Wheatstone, P.; Kvasnik, F.; Persaud, K.C. The Development and Optimisation of a Urinary Volatile Organic Compound Analytical Platform Using Gas Sensor Arrays for the Detection of Colorectal Cancer. Sensors 2025, 25, 599. [Google Scholar] [CrossRef] [PubMed]
  24. Eaton, J.R.; Christian, J.E.; Campbell, J.A. Surface phenomena related to odor measurements. Ann. N. Y. Acad. Sci. 1954, 58, 239–249. [Google Scholar] [CrossRef]
  25. Moncrieff, R.W. An instrument for measuring and classifying odors. J. Appl. Physiol. 1961, 16, 742–749. [Google Scholar] [CrossRef]
  26. Persaud, K.; Dodd, G. Analysis of discrimination mechanisms in the mammalian olfactory system using a model nose. Nature 1982, 299, 352–355. [Google Scholar] [CrossRef]
  27. Hatfield, J.V.; Neaves, P.; Hicks, P.J.; Persaud, K.; Travers, P. Towards an integrated electronic nose using conducting polymer sensors. Sens. Actuators B Chem. 1994, 18, 221–228. [Google Scholar] [CrossRef]
  28. Vlasov, Y.; Legin, A. Non-selective chemical sensors in analytical chemistry: From ‘electronic nose’ to ‘electronic tongue’, Fresenius. J. Anal. Chem. 1998, 361, 255–260. [Google Scholar] [CrossRef]
  29. Deng, H.; Mitsuno, H.; Kuroda, E.; Niki, S.; Kanzaki, R.; Nakamoto, T. Gas-phase odor mixture quantification based on relative comparison method using multiple olfactory receptors. Sens. Actuators B Chem. 2024, 401, 134995. [Google Scholar] [CrossRef]
  30. Mitsubayashi, K.; Toma, K.; Iitani, K.; Arakawa, T. Gas-phase biosensors: A review. Sens. Actuators B Chem. 2022, 367, 132053. [Google Scholar] [CrossRef]
  31. Hirata, Y.; Oda, H.; Osaki, T.; Takeuchi, S. Biohybrid sensor for odor detection. Lab A Chip 2021, 21, 2643–2657. [Google Scholar] [CrossRef]
  32. Hurot, C.; Scaramozzino, N.; Buhot, A.; Hou, Y. Bio-inspired strategies for improving the selectivity and sensitivity of artificial noses: A review. Sensors 2020, 20, 1803. [Google Scholar] [CrossRef]
  33. Deng, H.; Nakamoto, T. Odor Biosensors Based on Cell Expressing Olfactory Receptor: Recent Advances. Anal. Sens. 2024, 4, e202400006. [Google Scholar] [CrossRef]
  34. Xu, S.; Tian, M.; Li, C.; Liu, G.; Wang, J.; Yan, T.; Li, Z. Using a crumpled graphene field-effect transistor for ultrasensitive SARS-CoV-2 detection via its papain-like protease. Chem. Eng. J. 2025, 505, 159386. [Google Scholar] [CrossRef]
  35. Oprea, A.; Weimar, U. Gas sensors based on mass-sensitive transducers. Part 2: Improving the sensors towards practical application. Anal. Bioanal. Chem. 2020, 412, 6707–6776. [Google Scholar] [CrossRef] [PubMed]
  36. Bohbot, J.D.; Vernick, S. The emergence of insect odorant receptor-based biosensors. Biosensors 2020, 10, 26. [Google Scholar] [CrossRef]
  37. Rashed, M.S.; Abdelkarim, E.A.; Elsamahy, T.; Sobhy, M.; El-Mesery, H.S.; Salem, A. Advances in cell-based biosensors: Transforming food flavor evaluation with novel approaches. Food Chem. X 2025, 26, 102336. [Google Scholar] [CrossRef]
  38. Qin, C.; Wang, Y.; Hu, J.; Wang, T.; Liu, D.; Dong, J.; Lu, Y. Artificial Olfactory Biohybrid System: An Evolving Sense of Smell. Adv. Sci. 2023, 10, 2204726. [Google Scholar] [CrossRef]
  39. Misawa, N.; Mitsuno, H.; Kanzaki, R.; Takeuchi, S. Highly sensitive and selective odorant sensor using living cells expressing insect olfactory receptors. Proc. Natl. Acad. Sci. USA 2010, 107, 15340–15344. [Google Scholar] [CrossRef]
  40. Pawson, S.M.; Kerr, J.L.; O’Connor, B.C.; Lucas, P.; Strand, T.M. Light-Weight Portable Electroantennography Device as a Future Field-Based Tool for Applied Chemical Ecology. J. Chem. Ecol. 2020, 46, 557–566. [Google Scholar] [CrossRef]
  41. Lagunas, A.; Belloir, C.; Lalis, M.; Briand, L.; Topin, J.; Gorostiza, P.; Samitier, J. Ligand discrimination in hOR1A1 based on the capacitive response. Biosens. Bioelectron. 2025, 271, 117000. [Google Scholar] [CrossRef]
  42. Zhang, Z.; Chen, Z.; Liu, T.; Zhang, L. An implantable ionic liquid-gel microelectrode for in vivo monitoring of K+ levels in the living rat brain. Chem. Sci. 2025. ahead of print. [Google Scholar] [CrossRef] [PubMed]
  43. Adib, M.R.; Barrett, C.; O’Sullivan, S.; Flynn, A.; McFadden, M.; Kennedy, E.; O’Riordan, A. In situ pH-Controlled electrochemical sensors for glucose and pH detection in calf saliva. Biosens. Bioelectron. 2025, 275, 117234. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, Q.; Ye, W.; Xiao, L.; Du, L.; Hu, N.; Wang, P. Extracellular potentials recording in intact olfactory epithelium by microelectrode array for a bioelectronic nose. Biosens. Bioelectron. 2010, 25, 2212–2217. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, Q.; Hu, N.; Ye, W.; Cai, H.; Zhang, F.; Wang, P. Extracellular recording of spatiotemporal patterning in response to odors in the olfactory epithelium by microelectrode arrays. Biosens. Bioelectron. 2011, 27, 12–17. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, Q.; Zhang, F.; Hu, N.; Wang, H.; Hsia, K.J.; Wang, P. Microelectrode Recording of Tissue Neural Oscillations for a Bionic Olfactory Biosensor. J. Bionic Eng. 2012, 9, 494–500. [Google Scholar] [CrossRef]
  47. Liu, Q.; Hu, N.; Zhang, F.; Zhang, D.; Hsia, K.J.; Wang, P. Olfactory epithelium biosensor: Odor discrimination of receptor neurons from a bio-hybrid sensing system. Biomed. Microdevices 2012, 14, 1055–1061. [Google Scholar] [CrossRef]
  48. Liu, Q.; Zhang, F.; Zhang, D.; Hu, N.; Hsia, K.J.; Wang, P. Extracellular potentials recording in intact taste epithelium by microelectrode array for a taste sensor. Biosens. Bioelectron. 2013, 43, 186–192. [Google Scholar] [CrossRef]
  49. Zhuang, L.; Hu, N.; Dong, Q.; Liu, Q.; Wang, P. A high sensitive in vivo biosensing detection for odors by multiunit in rat olfactory bulb. Sens. Actuators B Chem. 2013, 186, 308–314. [Google Scholar] [CrossRef]
  50. Zhuang, L.; Hu, N.; Tian, F.; Dong, Q.; Hu, L.; Li, R. A high-sensitive detection method for carvone odor by implanted electrodes in rat olfactory bulb. Chin. Sci. Bull. 2014, 59, 29–37. [Google Scholar] [CrossRef]
  51. Gao, K.; Gao, F.; Du, L.; He, C.; Wan, H.; Wang, P. Integrated olfaction, gustation and toxicity detection by a versatile bioengineered cell-based biomimetic sensor. Bioelectrochemistry 2019, 128, 1–8. [Google Scholar] [CrossRef] [PubMed]
  52. Gao, K.; Gao, F.; Li, J.; He, C.; Liu, M.; Zhu, Q.; Qian, Z.; Ma, T.; Wang, P. Biomimetic integrated olfactory sensory and olfactory bulb systems in vitro based on a chip. Biosens. Bioelectron. 2021, 171, 112739. [Google Scholar] [CrossRef]
  53. Gao, F.; Gao, K.; Zhang, P.; Fu, Y.; Liu, X.; Bai, S.; Li, W.; Qian, Z. A biomimetic sensor using neurotransmitter detection to decode odor perception by an olfactory network. Biosens. Bioelectron. 2022, 211, 114391. [Google Scholar] [CrossRef] [PubMed]
  54. Guo, T.; Zhuang, L.; Qin, Z.; Zhang, B.; Hu, N.; Wang, P. Multi-odor discrimination by a novel bio-hybrid sensing preserving rat’s intact smell perception in vivo. Sens. Actuators B Chem. 2016, 225, 34–41. [Google Scholar] [CrossRef]
  55. Sato, K.; Takeuchi, S. Chemical vapor detection using a reconstituted insect olfactory receptor complex. Angew. Chem.—Int. Ed. 2014, 53, 11798–11802. [Google Scholar] [CrossRef]
  56. Micholt, E.; Jans, D.; Callewaert, G.; Bartic, C.; Lammertyn, J.; Nicolaï, B. Extracellular recordings from rat olfactory epithelium slices using micro electrode arrays. Sens. Actuators B Chem. 2013, 184, 40–47. [Google Scholar] [CrossRef]
  57. Liu, P.; Cao, T.; Xu, J.; Mao, X.; Wang, D.; Li, A. Plasticity of Sniffing Pattern and Neural Activity in the Olfactory Bulb of Behaving Mice During Odor Sampling, Anticipation, and Reward. Neurosci. Bull. 2020, 36, 598–610. [Google Scholar] [CrossRef]
  58. Gao, K.; Li, S.; Zhuang, L.; Qin, Z.; Zhang, B.; Huang, L.; Wang, P. In vivo bioelectronic nose using transgenic mice for specific odor detection. Biosens. Bioelectron. 2018, 102, 150–156. [Google Scholar] [CrossRef]
  59. Zhu, P.; Liu, S.; Tian, Y.; Chen, Y.; Chen, W.; Wang, P.; Du, L.; Wu, C. In Vivo Bioelectronic Nose Based on a Bioengineered Rat Realizes the Detection and Classification of Multiodorants. ACS Chem. Neurosci. 2022, 13, 1727–1737. [Google Scholar] [CrossRef]
  60. Yuan, Q.; Qin, C.; Duan, Y.; Jiang, N.; Liu, M.; Wan, H.; Zhuang, L.; Wang, P. An: In vivo bioelectronic nose for possible quantitative evaluation of odor masking using M/T cell spatial response patterns. Analyst 2022, 147, 178–186. [Google Scholar] [CrossRef]
  61. Khadka, R.; Aydemir, N.; Carraher, C.; Hamiaux, C.; Colbert, D.; Cheema, J.; Malmstrm, J.; Kralicek, A.; Travas-Sejdic, J. An ultrasensitive electrochemical impedance-based biosensor using insect odorant receptors to detect odorants. Biosens. Bioelectron. 2019, 126, 207–213. [Google Scholar] [CrossRef] [PubMed]
  62. Hou, Y.; Jaffrezic-Renault, N.; Martelet, C.; Zhang, A.; Errachid, A. A novel detection strategy for odorant molecules based on controlled bioengineering of rat olfactory receptor I7. Biosens. Bioelectron. 2006, 22, 1550–1555. [Google Scholar] [CrossRef] [PubMed]
  63. Khadka, R.; Aydemir, N.; Carraher, C.; Hamiaux, C.; Baek, P.; Cheema, J.; Kralicek, A.; Travas-Sejdic, J. Investigating Electrochemical Stability and Reliability of Gold Electrode-electrolyte Systems to Develop Bioelectronic Nose Using Insect Olfactory Receptor. Electroanalysis 2019, 31, 726–738. [Google Scholar] [CrossRef]
  64. Liu, Q.; Wang, H.; Li, H.; Zhang, J.; Zhuang, S.; Zhang, F.; Hsia, K.J.; Ping, W. Impedance sensing and molecular modeling of an olfactory biosensor based on chemosensory proteins of honeybee. Biosens. Bioelectron. 2013, 40, 174–179. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, Q.; Zhang, D.; Zhang, Q.; Huang, Y.; Luo, S.; Yao, Y.; Li, S.; Liu, Q. Impedance spectroscopy analysis of human odorant binding proteins immobilized on nanopore arrays for biochemical detection. Biosens. Bioelectron. 2016, 79, 251–257. [Google Scholar] [CrossRef]
  66. Calabrese, A.; Battistoni, P.; Ceylan, S.; Zeni, L.; Capo, A.; Varriale, A.; D’Auria, S.; Staiano, M. An Impedimetric Biosensor for Detection of Volatile Organic Compounds in Food. Biosensors 2023, 13, 341. [Google Scholar] [CrossRef]
  67. Khadka, R.; Carraher, C.; Hamiaux, C.; Travas-Sejdic, J.; Kralicek, A. Synergistic improvement in the performance of insect odorant receptor based biosensors in the presence of Orco. Biosens. Bioelectron. 2020, 153, 112040. [Google Scholar] [CrossRef]
  68. Peng, C.; Sui, Y.; Fang, C.; Sun, H.; Liu, W.; Li, X.; Qu, C.; Li, W.; Liu, J.; Wu, C. Highly sensitive and selective electrochemical biosensor using odorant-binding protein to detect aldehydes. Anal. Chim. Acta 2024, 1318, 342932. [Google Scholar] [CrossRef]
  69. Minic, J.; Persuy, M.A.; Godel, E.; Aioun, J.; Connerton, I.; Salesse, R.; Pajot-Augy, E. Functional expression of olfactory receptors in yeast and development of a bioassay for odorant screening. FEBS J. 2004, 272, 524–537. [Google Scholar] [CrossRef]
  70. Sung, J.H.; Ko, H.J.; Park, T.H. Piezoelectric biosensor using olfactory receptor protein expressed in Escherichia coli. Biosens. Bioelectron. 2005, 21, 1981–1986. [Google Scholar] [CrossRef]
  71. Ko, H.J.; Park, T.H. Piezoelectric olfactory biosensor: Ligand specificity and dose-dependence of an olfactory receptor expressed in a heterologous cell system. Biosens. Bioelectron. 2005, 20, 1327–1332. [Google Scholar] [CrossRef] [PubMed]
  72. Homma, C.; Tsukiiwa, M.; Noguchi, H.; Tanaka, M.; Okochi, M.; Tomizawa, H.; Sugizaki, Y.; Isobayashi, A.; Hayamizu, Y. Designable peptides on graphene field-effect transistors for selective detection of odor molecules. Biosens. Bioelectron. 2022, 224, 115047. [Google Scholar] [CrossRef] [PubMed]
  73. Datta-Chaudhuri, T.; Araneda, R.C.; Abshire, P.; Smela, E. Olfaction on a chip. Sens. Actuators B Chem. 2016, 235, 74–78. [Google Scholar] [CrossRef]
  74. Ozdemir, A.; Algül, Y.; Tirgil, N.Y. Reusable Turn-On Fluorescent Biosensor for Cardiac Biomarker Troponin I Detection Using QDs-SPION-Aptamer. J. Fluoresc. 2025. ahead of print. [Google Scholar] [CrossRef]
  75. Huang, Q.; Han, L.; Ma, H.; Lan, W.; Tu, K.; Peng, J.; Su, J.; Pan, L. An Aptamer Sensor Based on Alendronic Acid-Modified Upconversion Nanoparticles Combined with Magnetic Separation for Rapid and Sensitive Detection of Thiamethoxam. Foods 2025, 14, 182. [Google Scholar] [CrossRef]
  76. Park, S.Y.; Tan, J.K.S.; Mo, X.; Song, Y.; Lim, J.; Liew, X.R.; Chung, H.; Kim, S. Carbon Quantum Dots with Tunable Size and Fluorescence Intensity for Development of a Nano-biosensor. Small 2025, 21, e2404524. [Google Scholar] [CrossRef]
  77. Figueroa, X.A.; Cooksey, G.A.; Votaw, S.V.; Horowitz, L.F.; Folch, A. Large-scale investigation of the olfactory receptor space using a microfluidic microwell array. Lab a Chip 2010, 10, 1120–1127. [Google Scholar] [CrossRef]
  78. Fukutani, Y.; Koshizawa, T.; Yohda, M. Application of Vapor Phase Stimulation Method for Screening of Human Odorant Receptors Responding to Cinnamaldehyde. Sens. Mater. 2021, 33, 4319–4329. [Google Scholar] [CrossRef]
  79. de March, C.A.; Fukutani, Y.; Vihani, A.; Kida, H.; Matsunami, H. Real-time in vitro monitoring of odorant receptor activation by an odorant in the vapor phase. J. Vis. Exp. 2019, 146, e59446. [Google Scholar] [CrossRef]
  80. Hirata, Y.; Morimoto, Y.; Nam, E.; Takeuchi, S. Portable biohybrid odorant sensors using cell-laden collagen micropillars. Lab a Chip 2019, 19, 1971–1976. [Google Scholar] [CrossRef]
  81. Lee, S.H.; Oh, E.H.; Park, T.H. Cell-based microfluidic platform for mimicking human olfactory system. Biosens. Bioelectron. 2015, 74, 554–561. [Google Scholar] [CrossRef] [PubMed]
  82. Suzuki, M.; Yoshimoto, N.; Shimono, K.; Kuroda, S. Deciphering the Receptor Repertoire Encoding Specific Odorants by Time-Lapse Single-Cell Array Cytometry. Sci. Rep. 2016, 6, 19934. [Google Scholar] [CrossRef]
  83. Auer, T.O.; Khallaf, M.A.; Silbering, A.F.; Zappia, G.; Benton, R. Olfactory receptor and circuit evolution promote host specialization. Nature 2020, 579, 402–408. [Google Scholar] [CrossRef]
  84. Pfister, P.; Smith, B.C.; Evans, B.J.; Brann, J.H.; Rogers, M.E. Odorant Receptor Inhibition Is Fundamental to Odor Encoding. Curr. Biol. 2020, 30, 2574–2587.e6. [Google Scholar] [CrossRef] [PubMed]
  85. Si, G.; Kanwal, J.K.; Hu, Y.; Tabone, C.J.; Baron, J.; Berck, M.; Vignoud, G.; Samuel, A.D.T. Structured Odorant Response Patterns across a Complete Olfactory Receptor Neuron Population. Neuron 2019, 101, 950–962.e7. [Google Scholar] [CrossRef] [PubMed]
  86. Mitsuno, H.; Sakurai, T.; Namiki, S.; Mitsuhashi, H.; Kanzaki, R. Novel cell-based odorant sensor elements based on insect odorant receptors. Biosens. Bioelectron. 2015, 65, 287–294. [Google Scholar] [CrossRef]
  87. Kida, H.; Fukutani, Y.; Mainland, J.D.; De March, C.A.; Vihani, A.; Li, Y.R.; Chi, Q.; Toyama, A.; Liu, L.; Kameda, M. Vapor detection and discrimination with a panel of odorant receptors. Nat. Commun. 2018, 9, 4556. [Google Scholar] [CrossRef] [PubMed]
  88. Sukekawa, Y.; Mitsuno, H.; Kanzaki, R.; Nakamoto, T. Odor discrimination using cell-based odor biosensor system with fluorescent image processing. IEEE Sens. J. 2019, 19, 7192–7200. [Google Scholar] [CrossRef]
  89. Deng, H.; Mitsuno, H.; Kanzaki, R.; Nakamoto, T. Gas Phase Odorant Detection by Insect Olfactory Receptor. IEEE Sens. J. 2021, 21, 21184–21191. [Google Scholar] [CrossRef]
  90. Park, T.H.; Lee, S.H.; Oh, E.H.; Lee, S.H.; Ko, H.J. Cell-based high-throughput odorant screening system through visualization on a microwell array. Biosens. Bioelectron. 2013, 53, 18–25. [Google Scholar] [CrossRef]
  91. Deng, H.; Mitsuno, H.; Kanzaki, R.; Nakamoto, T. Study of Liquid Film Thickness for Gas Phase Odor Biosensor. IEEE Sens. J. 2022, 22, 16785–16793. [Google Scholar] [CrossRef]
  92. Chen, T.W.; Wardill, T.J.; Sun, Y.; Pulver, S.R.; Renninger, S.L.; Baohan, A.; Schreiter, E.R.; Kerr, R.A.; Orger, M.B.; Jayaraman, V.; et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 2013, 499, 295–300. [Google Scholar] [CrossRef] [PubMed]
  93. Cho, S.W.; Ko, H.J.; Park, T.H. Identification of a Lung Cancer Biomarker Using a Cancer Cell Line and Screening of Olfactory Receptors for Biomarker Detection. Biotechnol. Bioprocess Eng. 2021, 26, 55–62. [Google Scholar] [CrossRef]
  94. El Kazzy, M.; Weerakkody, J.S.; Hurot, C.; Mathey, R.; Hou, Y. An overview of artificial olfaction systems with a focus on surface plasmon resonance for the analysis of volatile organic compounds. Biosensors 2021, 11, 244. [Google Scholar] [CrossRef]
  95. Woyessa, G.; Bang, O. Sensitivity Enhancement of Polymer Optical Fiber Surface Plasmon Resonance Sensor Utilizing ITO Overlayer. Sensors 2025, 25, 1863. [Google Scholar] [CrossRef] [PubMed]
  96. Ashrafian, M.; Olyaee, S.; Seifouri, M. Highly sensitive cancer detection using an open D-channel PCF-based SPR biosensor. Sci. Rep. 2025, 15, 10168. [Google Scholar] [CrossRef]
  97. Xu, Q.; Yin, H.; Cui, M.; Huang, R.; Su, R. An Enhanced Bimetallic Optical Fiber SPR Biosensor Using Graphene Oxide for the Label-Free and Sensitive Detection of Human IgG. Sensors 2025, 25, 1630. [Google Scholar] [CrossRef]
  98. Quintanilla-Villanueva, G.E.; Rodríguez-Quiroz, O.; Sánchez-Álvarez, A.; Rodríguez-Delgado, J.M.; Villarreal-Chiu, J.F.; Luna-Moreno, D.; Rodríguez-Delgado, M.M. An Innovative Enzymatic Surface Plasmon Resonance-Based Biosensor Designed for Precise Detection of Glycine Amino Acid. Biosensors 2025, 15, 81. [Google Scholar] [CrossRef]
  99. Rossi, C.; Homand, J.; Bauche, C.; Hamdi, H.; Ladant, D.; Chopineau, J. Differential Mechanisms for Calcium-Dependent Protein/Membrane Association as Evidenced from SPR-Binding Studies on Supported Biomimetic Membranes. Biochemistry 2003, 42, 15273–15283. [Google Scholar] [CrossRef]
  100. Lee, J.Y.; Ko, H.J.; Lee, S.H.; Park, T.H. Cell-based measurement of odorant molecules using surface plasmon resonance. Enzym. Microb. Technol. 2006, 39, 375–380. [Google Scholar] [CrossRef]
  101. Cennamo, N.; Stefano, D.G.; Antonio, V.; Maria, S.; Fabio, D.P.; Andrea, N.; Luigi, Z.; Sabato, D.; Karl-Wilhelm, K. Easy to use plastic optical fiber-based biosensor for detection of butanal. PLoS ONE 2015, 10, e0116770. [Google Scholar] [CrossRef]
  102. Zhang, D.; Lu, Y.; Zhang, Q.; Yao, Y.; Li, S.; Li, H.; Zhuang, S.; Jing, J.; Gang, L.L.; Liu, Q. Nanoplasmonic monitoring of odorants binding to olfactory proteins from honeybee as biosensor for chemical detection. Sens. Actuators B Chem. 2015, 221, 341–349. [Google Scholar] [CrossRef]
  103. Choi, Y.R.; Shim, J.; Park, J.H.; Kim, Y.S.; Kim, M.J. Discovery of orphan olfactory receptor 6m1 as a new anticancer target in mcf-7 cells by a combination of surface plasmon resonance-based and cell-based systems. Sensors 2021, 21, 3468. [Google Scholar] [CrossRef]
  104. Benilova, I.; Chegel, V.I.; Ushenin, Y.V.; Vidic, J.; Soldatkin, A.P.; Martelet, C.; Pajot, E.; Jaffrezic-Renault, N. Stimulation of human olfactory receptor 17–40 with odorants probed by surface plasmon resonance. Eur. Biophys. J. 2008, 37, 807–814. [Google Scholar] [CrossRef] [PubMed]
  105. Oh, E.H.; Lee, S.H.; Ko, H.J.; Park, T.H. Odorant detection using liposome containing olfactory receptor in the SPR system. Sens. Actuators B Chem. 2014, 198, 188–193. [Google Scholar] [CrossRef]
  106. Lee, S.H.; Ko, H.J.; Park, T.H. Real-time monitoring of odorant-induced cellular reactions using surface plasmon resonance. Biosens. Bioelectron. 2009, 25, 55–60. [Google Scholar] [CrossRef] [PubMed]
  107. Vidic, J.; Grosclaude, J.; Monnerie, R.; Persuy, M.A.; Badonnel, K.; Baly, C.; Caillol, M.; Briand, L.; Salesse, R.; Pajot-Augy, E.; et al. On a chip demonstration of a functional role for odorant binding protein in the preservation of olfactory receptor activity at high odorant concentration. Lab a Chip 2008, 8, 678–688. [Google Scholar] [CrossRef]
  108. Lee, S.; Park, J.; Song, J.; Lee, J.J.; Kim, J. The Effect of GO Flake Size on Field-Effect Transistor (FET)-Based Biosensor Performance for Detection of Ions and PACAP 38. Biosensors 2025, 15, 86. [Google Scholar] [CrossRef]
  109. Hadded, A.; Ayed, M.B.; Alshaya, S.A. An FPGA-Based SiNW-FET Biosensing System for Real-Time Viral Detection: Hardware Amplification and 1D CNN for Adaptive Noise Reduction. Sensors 2025, 25, 236. [Google Scholar] [CrossRef]
  110. Nandanwar, S.; Lee, S.; Park, M.; Kim, H.J. Label-Free Extended Gate Field-Effect Transistor for Sensing Microcystin-LR in Freshwater Samples. Sensors 2025, 25, 1587. [Google Scholar] [CrossRef]
  111. Das, D.; Dhar, R.S.; Ghosh, P.K.; Sharma, Y.; Banerjee, A. Sensitivity Analysis of Biosensor based SiGe Source Dual Gate Tunnel FET Having Negative Capacitance. IEEE Access 2025, 13, 7426–7436. [Google Scholar] [CrossRef]
  112. Hao, R.; Liu, L.; Yuan, J.; Wu, L.; Lei, S. Recent Advances in Field Effect Transistor Biosensors: Designing Strategies and Applications for Sensitive Assay. Biosensors 2023, 13, 426. [Google Scholar] [CrossRef] [PubMed]
  113. Murugathas, T.; Zheng, H.Y.; Colbert, D.; Kralicek, A.V.; Carraher, C.; Plank, N.O.V. Biosensing with Insect Odorant Receptor Nanodiscs and Carbon Nanotube Field-Effect Transistors. ACS Appl. Mater. Interfaces 2019, 11, 9530–9538. [Google Scholar] [CrossRef] [PubMed]
  114. Lee, M.; Yang, H.; Kim, D.; Yang, M.; Park, T.H.; Hong, S. Human-like smelling of a rose scent using an olfactory receptor nanodisc-based bioelectronic nose. Sci. Rep. 2018, 8, 13945. [Google Scholar] [CrossRef]
  115. Yang, H.; Kim, D.; Kim, J.; Moon, D.; Song, H.S.; Lee, M.; Hong, S.; Park, T.H. Nanodisc-Based Bioelectronic Nose Using Olfactory Receptor Produced in Escherichia coli for the Assessment of the Death-Associated Odor Cadaverine. ACS Nano 2017, 11, 11847–11855. [Google Scholar] [CrossRef]
  116. Choi, D.; Lee, S.J.H.; Baek, D.; Kim, S.O.; Shin, J.; Choi, Y.; Cho, Y.; Bang, S.; Park, J.Y.; Lee, S.J.H.; et al. Bioelectrical Nose Platform Using Odorant-Binding Protein as a Molecular Transporter Mimicking Human Mucosa for Direct Gas Sensing. ACS Sens. 2022, 7, 3399–3408. [Google Scholar] [CrossRef]
  117. Kim, B.; Lee, J.; Namgung, S.; Kim, J.; Park, J.Y.; Lee, M.S.; Hong, S. DNA sensors based on CNT-FET with floating electrodes. Sens. Actuators B Chem. 2012, 169, 182–187. [Google Scholar] [CrossRef]
  118. Lin, Z.; Wu, G.; Zhao, L.; Lai, K.W.C. Detection of bacterial metabolic volatile indole using a graphene-based field-effect transistor biosensor. Nanomaterials 2021, 11, 1155. [Google Scholar] [CrossRef]
  119. Park, S.J.; Seo, S.E.; Kim, K.H.; Lee, S.H.; Kim, J.; Ha, S.; Song, H.S.; Lee, S.H.; Kwon, O.S. Real-time monitoring of geosmin based on an aptamer-conjugated graphene field-effect transistor. Biosens. Bioelectron. 2021, 174, 112804. [Google Scholar] [CrossRef]
  120. Kwon, O.S.; Song, H.S.; Park, S.J.; Lee, S.H.; An, J.H.; Park, J.W.; Yang, H.; Yoon, H.; Bae, J.; Park, T.H. An Ultrasensitive, Selective, Multiplexed Superbioelectronic Nose That Mimics the Human Sense of Smell. Nano Lett. 2015, 15, 6559–6567. [Google Scholar] [CrossRef]
  121. Oh, J.; Yang, H.; Jeong, G.E.; Moon, D.; Jang, J. Ultrasensitive, Selective, and Highly Stable Bioelectronic Nose That Detects the Liquid and Gaseous Cadaverine. Anal. Chem. 2019, 91, 12181–12190. [Google Scholar] [CrossRef] [PubMed]
  122. Kuznetsov, A.E.; Komarova, N.V.; Kuznetsov, E.V.; Andrianova, M.S.; Grudtsov, V.P.; Rybachek, E.N.; Puchnin, K.V.; Ryazantsev, D.V.; Saurov, A.N. Integration of a field effect transistor-based aptasensor under a hydrophobic membrane for bioelectronic nose applications. Biosens. Bioelectron. 2019, 129, 29–35. [Google Scholar] [CrossRef]
  123. Terutsuki, D.; Mitsuno, H.; Sakurai, T.; Okamoto, Y.; Kanzaki, R. Increasing cell–device adherence using cultured insect cells for receptor-based biosensors. R. Soc. Open Sci. 2018, 5, 172366. [Google Scholar] [CrossRef]
  124. Obořilová, R.; Kučerová, E.; Botka, T.; Vaisocherová-Lísalová, H.; Skládal, P.; Farka, Z. Piezoelectric biosensor with dissipation monitoring enables the analysis of bacterial lytic agent activity. Sci. Rep. 2025, 15, 3419. [Google Scholar] [CrossRef]
  125. Siddiq, M.A. Novel plant extract-chitosan nanocomposite (SCE/ChNPs) biosensor for trace level arsenic sensing in water samples using quartz crystal microbalance with dissipation monitoring. Nano Express 2025, 6, 015004. [Google Scholar] [CrossRef]
  126. Park, S.; Gerber, A.; Santa, C.; Aktug, G.; Hengerer, B.; Clark, H.A.; Jonas, U.; Dostalek, J.; Sergelen, K. Molecularly Responsive Aptamer-Functionalized Hydrogel for Continuous Plasmonic Biomonitoring. J. Am. Chem. Soc. 2025, 147, 11485–11500. [Google Scholar] [CrossRef] [PubMed]
  127. Ogi, H. Ultrasensitive wireless quartz crystal microbalance bio/gas sensors. Jpn. J. Appl. Phys. 2024, 63, 040802. [Google Scholar] [CrossRef]
  128. Su, X.; Chen, D.; Li, N.; Stevenson, A.C.; Li, G.; Hu, R. A wireless electrode-free QCM-D in a multi-resonance mode for volatile organic compounds discrimination. Sens. Actuators A Phys. 2020, 305, 111938. [Google Scholar] [CrossRef]
  129. Chen, D.; Li, H.; Su, X.; Li, N.; Wang, Y.; Stevenson, A.C.; Hu, R.; Li, G. A wireless-electrodeless quartz crystal microbalance method for non-enzymatic glucose monitoring. Sens. Actuators B Chem. 2019, 287, 35–41. [Google Scholar] [CrossRef]
  130. Chen, D.; Zhang, K.; Zhou, H.; Fan, G.; Wang, Y.; Li, G.; Hu, R. A wireless-electrodeless quartz crystal microbalance with dissipation DMMP sensor. Sens. Actuators B Chem. 2018, 261, 408–417. [Google Scholar] [CrossRef]
  131. Chen, T.; Sha, P.; Xu, Y.; Su, X.; Chen, D.; Li, N.; Li, G.; Hu, R. Online pH monitoring based on a wireless electrodeless quartz crystal microbalance with dissipation. Sens. Actuators A Phys. 2021, 331, 112952. [Google Scholar] [CrossRef]
  132. Du, L.; Wu, C.; Peng, H.; Zou, L.; Zhao, L.; Huang, L.; Ping, W. Piezoelectric olfactory receptor biosensor prepared by aptamer-assisted immobilization. Sens. Actuators B Chem. 2013, 187, 481–487. [Google Scholar] [CrossRef]
  133. Sankaran, S.; Panigrahi, S.; Mallik, S. Olfactory receptor based piezoelectric biosensors for detection of alcohols related to food safety applications. Sens. Actuators B Chem. 2011, 155, 8–18. [Google Scholar] [CrossRef]
  134. Nardiello, M.; Scieuzo, C.; Salvia, R.; Farina, D.; Franco, A.; Cammack, J.A.; Tomberlin, J.K.; Falabella, P.; Persaud, K.C. Odorant binding proteins from Hermetia illucens: Potential sensing elements for detecting volatile aldehydes involved in early stages of organic decomposition. Nanotechnology 2022, 33, 205501. [Google Scholar] [CrossRef] [PubMed]
  135. Wasilewski, T.; Szulczyński, B.; Wojciechowski, M.; Kamysz, W.; Gębicki, J. Determination of long-chain aldehydes using a novel quartz crystal microbalance sensor based on a biomimetic peptide. Microchem. J. 2020, 154, 104509. [Google Scholar] [CrossRef]
  136. Wang, Y.; Yan, C.; Liang, C.; Liu, Y.; Li, H.; Zhang, C.; Duan, X.; Pan, Y. Sensitive Materials Used in Surface Acoustic Wave Gas Sensors for Detecting Sulfur-Containing Compounds. Polymers 2024, 16, 457. [Google Scholar] [CrossRef]
  137. Liang, C.; Yan, C.; Zhai, S.; Wang, Y.; Hu, A.; Wang, W.; Pan, Y. Recent Progress in Flexible Surface Acoustic Wave Sensing Technologies. Micromachines 2024, 15, 357. [Google Scholar] [CrossRef]
  138. Zhao, H.; Jiao, C.; Wang, Q.; Gao, C.; Sun, J. Surface Acoustic Wave Sensor for Selective Multi-Parameter Measurements in Cardiac Magnetic Field Detection. Appl. Sci. 2025, 15, 3583. [Google Scholar] [CrossRef]
  139. Zeng, Y.; Yuan, R.; Fu, H.; Xu, Z.; Wei, S. Foodborne pathogen detection using surface acoustic wave biosensors: A review. RSC Adv. 2024, 14, 37087–37103. [Google Scholar] [CrossRef]
  140. Aleixandre, M.; Horrillo, M.C. Recent Advances in SAW Sensors for Detection of Cancer Biomarkers. Biosensors 2025, 15, 88. [Google Scholar] [CrossRef]
  141. Kukaev, A.; Shalymov, E.; Shevchenko, S.; Sorvina, M.; Venediktov, V. Advancements in Surface Acoustic Wave Gyroscope Technology in 2015–2024. Sensors 2025, 25, 877. [Google Scholar] [CrossRef] [PubMed]
  142. Mandal, D.; Banerjee, S. Surface Acoustic Wave (SAW) Sensors: Physics, Materials, and Applications. Sensors 2022, 22, 820. [Google Scholar] [CrossRef] [PubMed]
  143. Cui, B.; Wang, W.; Cheng, L.; Jin, J.; Hu, A.; Ren, Z.; Xue, X.; Liang, Y. Acoustic impedance-based surface acoustic wave chip for gas leak detection and respiratory monitoring. Commun. Eng. 2025, 4, 15. [Google Scholar] [CrossRef] [PubMed]
  144. Caliendo, C.; Benetti, M.; Cannatà, D.; Laidoudi, F.; Petrone, G. Interface Acoustic Waves in 128° YX-LiNbO3/SU-8/Overcoat Structures. Micromachines 2025, 16, 99. [Google Scholar] [CrossRef]
  145. Stubbs, D.D.; Hunt, W.D.; Lee, S.H.; Doyle, D.F. Gas phase activity of anti-FITC antibodies immobilized on a surface acoustic wave resonator device. Biosens. Bioelectron. 2002, 17, 471–477. [Google Scholar] [CrossRef]
  146. Wu, C.; Du, L.; Wang, D.; Wang, L.; Zhao, L.; Wang, P. A novel surface acoustic wave-based biosensor for highly sensitive functional assays of olfactory receptors. Biochem. Biophys. Res. Commun. 2011, 407, 18–22. [Google Scholar] [CrossRef]
  147. Cannatà, D.; Di Pietrantonio, F.; Varriale, A.; D’Auria, S.; Staiano, M.; Verona, E.; Benetti, M. Detection of odorant molecules via surface acoustic wave biosensor array based on odorant-binding proteins. Biosens. Bioelectron. 2013, 41, 328–334. [Google Scholar] [CrossRef]
  148. Palla-Papavlu, A.; Patrascioiu, A.; Di Pietrantonio, F.; Fernández-Pradas, J.M.; Cannatà, D.; Benetti, M.; D’Auria, S.; Verona, E.; Serra, P. Preparation of surface acoustic wave odor sensors by laser-induced forward transfer. Sens. Actuators B Chem. 2014, 192, 369–377. [Google Scholar] [CrossRef]
  149. Wang, J.; Du, L.; Krause, S.; Wu, C.; Wang, P. Surface modification and construction of LAPS towards biosensing applications. Sens. Actuators B Chem. 2018, 265, 161–173. [Google Scholar] [CrossRef]
  150. Hou, C.; Dong, J.; Qin, H.; Huo, D.; Shen, C. Development of Cell-Based Olfactory Biosensors. Anal. Lett. 2012, 45, 202–218. [Google Scholar] [CrossRef]
  151. Li, X.; Liu, S.; Tan, J.; Wu, C. Light-Addressable Potentiometric Sensors in Microfluidics. Front. Bioeng. Biotechnol. 2022, 10, 833481. [Google Scholar] [CrossRef] [PubMed]
  152. Liu, Q.; Ye, W.; Hu, N.; Cai, H.; Yu, H.; Wang, P. Olfactory receptor cells respond to odors in a tissue and semiconductor hybrid neuron chip. Biosens. Bioelectron. 2010, 26, 1672–1678. [Google Scholar] [CrossRef] [PubMed]
  153. Liu, Q.; Ye, W.; Yu, H.; Hu, N.; Du, L.; Wang, P.; Yang, M. Olfactory mucosa tissue-based biosensor: A bioelectronic nose with receptor cells in intact olfactory epithelium. Sens. Actuators B Chem. 2009, 146, 527–533. [Google Scholar] [CrossRef]
  154. Wu, C.; Chen, P.; Yu, H.; Liu, Q.; Zong, X.; Hua, C.; Ping, W. A novel biomimetic olfactory-based biosensor for single olfactory sensory neuron monitoring. Biosens. Bioelectron. 2008, 24, 1498–1502. [Google Scholar] [CrossRef]
  155. Wu, C.S.; Chen, P.H.; Yuan, Q.; Wang, P. Response enhancement of olfactory sensory neurons-based biosensors for odorant detection. J. Zhejiang Univ. B 2009, 10, 285–290. [Google Scholar] [CrossRef]
  156. Du, L.; Wu, C.; Peng, H.; Zhao, L.; Huang, L.; Wang, P. Bioengineered olfactory sensory neuron-based biosensor for specific odorant detection. Biosens. Bioelectron. 2013, 40, 401–406. [Google Scholar] [CrossRef]
  157. Du, L.; Chen, W.; Tian, Y.; Zhu, P.; Wang, J.; Cai, W.; Wu, C. A light-addressable microfluidic device for label-free functional assays of bioengineered taste receptor cells via extracellular recording. Biophys. Rep. 2019, 5, 73–79. [Google Scholar] [CrossRef]
  158. Du, L.; Zou, L.; Zhao, L.; Huang, L.; Wang, P.; Wu, C. Label-free functional assays of chemical receptors using a bioengineered cell-based biosensor with localized extracellular acidification measurement. Biosens. Bioelectron. 2014, 54, 623–627. [Google Scholar] [CrossRef] [PubMed]
  159. Misawa, N.; Fujii, S.; Kamiya, K.; Osaki, T.; Takaku, T.; Takahashi, Y.; Takeuchi, S. Construction of a Biohybrid Odorant Sensor Using Biological Olfactory Receptors Embedded into Bilayer Lipid Membrane on a Chip. ACS Sens. 2019, 4, 711–716. [Google Scholar] [CrossRef]
  160. Yamada, T.; Sugiura, H.; Mimura, H.; Kamiya, K.; Osaki, T.; Takeuchi, S. Highly sensitive VOC detectors using insect olfactory receptors reconstituted into lipid bilayers. Sci. Adv. 2021, 7, eabd2013. [Google Scholar] [CrossRef]
  161. Fujii, S.; Nobukawa, A.; Osaki, T.; Morimoto, Y.; Takeuchi, S. Pesticide vapor sensing using an aptamer, nanopore, and agarose gel on a chip. Lab a Chip 2017, 17, 2421–2425. [Google Scholar] [CrossRef] [PubMed]
  162. Pérez, R.L.; Ayala, C.E.; Park, J.Y.; Choi, J.W.; Warner, I.M. Coating-based quartz crystal microbalance detection methods of environmentally relevant volatile organic compounds. Chemosensors 2021, 9, 153. [Google Scholar] [CrossRef]
  163. Stensmyr, M.C.; Dweck, H.K.M.; Farhan, A.; Ibba, I.; Strutz, A.; Mukunda, L.; Linz, J.; Grabe, V.; Steck, K.; Lavista-Llanos, S.; et al. A conserved dedicated olfactory circuit for detecting harmful microbes in drosophila. Cell 2012, 151, 1345–1357. [Google Scholar] [CrossRef] [PubMed]
  164. Son, M.; Cho, D.G.; Lim, J.H.; Park, J.; Hong, S.; Ko, H.J.; Park, T.H. Real-time monitoring of geosmin and 2-methylisoborneol, representative odor compounds in water pollution using bioelectronic nose with human-like performance. Biosens. Bioelectron. 2015, 74, 199–206. [Google Scholar] [CrossRef] [PubMed]
  165. Termtanasombat, M.; Mitsuno, H.; Misawa, N.; Yamahira, S.; Sakurai, T.; Yamaguchi, S.; Nagamune, T.; Kanzaki, R. Cell-Based Odorant Sensor Array for Odor Discrimination Based on Insect Odorant Receptors. Chem. Ecol. 2016, 42, 716–724. [Google Scholar] [CrossRef]
  166. Deng, H.; Mitsuno, H.; Kanzaki, R.; Nomoto, S.; Nakamoto, T. Gas-phase Odorant Fast Quantification by Odor Biosensor Based on Reference Response Model. IEEE Sens. J. 2023, 23, 24169–24178. [Google Scholar] [CrossRef]
  167. Anderson, M.J.; Sullivan, J.G.; Horiuchi, T.K.; Fuller, S.B.; Daniel, T.L. A bio-hybrid odor-guided autonomous palm-sized air vehicle. Bioinspiration Biomim. 2020, 16, 026002. [Google Scholar] [CrossRef]
  168. Lochmatter, T.; Raemy, X.; Matthey, L.; Indra, S.; Martinoli, A. A comparison of casting and spiraling algorithms for odor source localization in laminar flow. In Proceedings of the 2008 IEEE International Conference on Robotics and Automation, Pasadena, CA, USA, 19–23 May 2008. [Google Scholar] [CrossRef]
  169. Terutsuki, D.; Uchida, T.; Fukui, C.; Sukekawa, Y.; Okamoto, Y.; Kanzaki, R. Real-time odor concentration and direction recognition for efficient odor source localization using a small bio-hybrid drone. Sens. Actuators B Chem. 2021, 339, 129770. [Google Scholar] [CrossRef]
  170. Wu, Y.; Wang, Q.; Li, X.; Li, K.; Huang, D.; Zou, K.; Zhang, Z.; Qian, Y.; Zhu, C.; Kong, X.Y. Olfactory-Inspired Separation-Sensing Nanochannel-Based Electronics for Wireless Sweat Monitoring. ACS Nano 2025, 3, 3781–3790. [Google Scholar] [CrossRef]
  171. Pelosi, P.; Calvello, M.; Ban, L. Diversity of odorant-binding proteins and chemosensory proteins in insects. Chem. Senses 2005, 30, i291–i292. [Google Scholar] [CrossRef]
  172. Oncu, S.; Yeniterzi, D.; Karakurt, O.; Cakmaktepe, S.; Cirpan, A.; Soylemez, S.; Toppare, L. Synthesis and nanoparticle formation of pyrazine, benzodithiophene (BDT) and fluorene containing conjugated polymer (P-PBF) and its biosensing performance toward catechol. Microchem. J. 2024, 207, 112158. [Google Scholar] [CrossRef]
  173. Bordbar, M.M.; Tashkhourian, J.; Tavassoli, A.; Bahramali, E.; Hemmateenejad, B. Ultrafast detection of infectious bacteria using optoelectronic nose based on metallic nanoparticles. Sens. Actuators B Chem. 2020, 319, 128262. [Google Scholar] [CrossRef]
  174. Dent, A.G.; Sutedja, T.G.; Zimmerman, P.V. Exhaled breath analysis for lung cancer. J. Thorac. Dis. 2013, 5 (Suppl. S5), S540–S550. [Google Scholar] [CrossRef] [PubMed]
  175. Ye, M.; Chien, P.J.; Toma, K.; Arakawa, T.; Mitsubayashi, K. An acetone bio-sniffer (gas phase biosensor) enabling assessment of lipid metabolism from exhaled breath. Biosens. Bioelectron. 2015, 73, 208–213. [Google Scholar] [CrossRef]
  176. Toda, K.; Koga, T.; Kosuge, J.; Kashiwagi, M.; Oguchi, H.; Arimoto, T. Micro gas analyzer measurement of nitric oxide in breath by direct wet scrubbing and fluorescence detection. Anal. Chem. 2009, 81, 7031–7037. [Google Scholar] [CrossRef]
  177. Chen, Y.; Zhang, Y.; Pan, F.; Liu, J.; Wang, K.; Zhang, C.; Cheng, S.; Lu, L.; Zhang, W.; Zhang, Z.; et al. Breath Analysis Based on Surface-Enhanced Raman Scattering Sensors Distinguishes Early and Advanced Gastric Cancer Patients from Healthy Persons. ACS Nano 2016, 10, 8169–8179. [Google Scholar] [CrossRef]
  178. Nakano-Baker, O.; Fong, H.; Shukla, S.; Lee, R.V.; Cai, L.; Godin, D.; Hennig, T.; Rath, S.; Novosselov, I.; Dogan, S.D.; et al. Data-driven design of a multiplexed, peptide-sensitized transistor to detect breath VOC markers of COVID-19. Biosens. Bioelectron. 2023, 229, 115237. [Google Scholar] [CrossRef]
  179. Hu, J.; Hu, N.; Pan, D.; Zhu, Y.; Jin, X.; Wu, S.; Lu, Y. Smell cancer by machine learning-assisted peptide/MXene bioelectronic array. Biosens. Bioelectron. 2024, 262, 116562. [Google Scholar] [CrossRef] [PubMed]
  180. Lee, S.W.; Selyanchyn, R.; Wakamatsu, S.; Hayashi, K. Nanoassembled thin-film-coated quartz crystal microbalance odor sensors for environmental and human breath ammonia assessments. Sens. Mater. 2018, 30, 1133–1144. [Google Scholar] [CrossRef]
  181. Sankaran, S.; Panigrahi, S.; Mallik, S. Odorant binding protein based biomimetic sensors for detection of alcohols associated with Salmonella contamination in packaged beef. Biosens. Bioelectron. 2011, 26, 3103–3109. [Google Scholar] [CrossRef]
  182. Wang, Z.; Ma, W.; Yan, S.; Chen, R.; Qin, G. High-Performance Odorant Binding Protein-Derived Peptide Biosensor for Salmonella Contamination in Food Detection Based on Mace Structure. ACS Sens. 2025, 10, 897–906. [Google Scholar] [CrossRef]
  183. Son, M.; Kim, D.; Kang, J.; Lim, J.H.; Lee, S.H.; Ko, H.J.; Hong, S.; Park, T.H. Bioelectronic nose using odorant binding protein-derived peptide and carbon nanotube field-effect transistor for the assessment of salmonella contamination in food. Anal. Chem. 2016, 88, 11283–11287. [Google Scholar] [CrossRef]
  184. Cui, W.; Hu, G.; Lv, E.; Li, C.; Wang, Z.; Li, Q.; Qian, Z.; Wang, J.; Xu, S.; Wang, R. A label-free and enzyme-free fluorescent aptasensor for amplified detection of kanamycin in milk sample based on target-triggered catalytic hairpin assembly. Food Control 2021, 133, 108654. [Google Scholar] [CrossRef]
  185. Wasilewski, T.; Gębicki, J.; Kamysz, W. Bio-inspired approaches for explosives detection. TrAC Trends Anal. Chem. 2021, 142, 116330. [Google Scholar] [CrossRef]
  186. Otto, J.; Brown, M.F.; Long, W. Training rats to search and alert on contraband odors. Appl. Anim. Behav. Sci. 2002, 77, 217–232. [Google Scholar] [CrossRef]
  187. Persaud, K.C. Biomimetic olfactory sensors. IEEE Sens. J. 2012, 12, 3108–3112. [Google Scholar] [CrossRef]
  188. Scorsone, E.; Manai, R.; Cali, K.; Ricatti, M.J.; Mucignat, C. Biosensor array based on ligand binding proteins for narcotics and explosives detection. Sens. Actuators B Chem. 2021, 334, 129587. [Google Scholar] [CrossRef]
  189. Radhika, V.; Proikas-Cezanne, T.; Jayaraman, M.; Onesime, D.; Ha, J.H.; Dhanasekaran, D.N. Chemical sensing of DNT by engineered olfactory yeast strain. Nat. Chem. Biol. 2007, 3, 325–330. [Google Scholar] [CrossRef]
  190. Yokoshiki, Y.; Nagayoshi, K.; Nakamoto, T. Improving Performance of Odor Biosensor System Using Olfactory Receptor. Sens. Mater. 2022, 34, 1617–1626. [Google Scholar] [CrossRef]
  191. Mujiono, T.; Sukekawa, Y.; Nakamoto, T.; Mitsuno, H.; Termtanasombat, M.; Kanzaki, R.; Misawa, N. Sensitivity improvement by applying lock-in technique to fluorescent instrumentation for cell-based odor sensor. Sens. Mater. 2017, 29, 65–76. [Google Scholar] [CrossRef]
  192. Deng, H.; Mitsuno, H.; Kanzaki, R.; Nakamoto, T. Extending lifetime of gas-phase odor biosensor using liquid thickness control and liquid exchange. Biosens. Bioelectron. 2022, 199, 113887. [Google Scholar] [CrossRef] [PubMed]
  193. Deng, H.; Mitsuno, H.; Kanzaki, R.; Nakamoto, T. Discrete and continuous odor quantification in gas-phase odor biosensor. Meas. Sci. Technol. 2024, 35, 075105. [Google Scholar] [CrossRef]
  194. Deng, H.; Sukekawa, Y.; Mitsuno, H.; Kanzaki, R.; Nakamoto, T. Active Tracking of Temporally Changing Gas-Phase Odor Mixture Using an Array of Cells Expressing Olfactory Receptors. Anal. Chem. 2023, 95, 11558–11565. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The schematic diagram of odor biosensors. Trace amount of the target odor are captured by the sensing material. In odor biosensors, the sensing materials include but are not limited to OSN, cells expressing OR, olfactory epithelium, olfactory bulb, antenna, OR protein, OBP, peptide, nanovesicle, enzyme, and aptamer. Transducers—e.g., microelectrode, fluorescence, QCM, light-addressable potentiometric sensor (LAPS), FET, SPR, and SAW—transform the combination of target ligand and sensing material into a readable signal and thereby detect trace amounts of the target odorant.
Figure 1. The schematic diagram of odor biosensors. Trace amount of the target odor are captured by the sensing material. In odor biosensors, the sensing materials include but are not limited to OSN, cells expressing OR, olfactory epithelium, olfactory bulb, antenna, OR protein, OBP, peptide, nanovesicle, enzyme, and aptamer. Transducers—e.g., microelectrode, fluorescence, QCM, light-addressable potentiometric sensor (LAPS), FET, SPR, and SAW—transform the combination of target ligand and sensing material into a readable signal and thereby detect trace amounts of the target odorant.
Electronics 14 01852 g001
Figure 3. Fluorescence-based odor biosensors: (a) mouse olfactory epithelium treated with papain, followed by isolated OSN incubated with Fluo-4 AM and trapped in a microchamber array; (b) OR, olfactory receptor co-receptor (Orco), and fluorescent protein genes constructed on a plasmid and then transfected into Sf21 cells. (a) is reprinted from [82]. (b) is reprinted from [86] with permission from Elsevier.
Figure 3. Fluorescence-based odor biosensors: (a) mouse olfactory epithelium treated with papain, followed by isolated OSN incubated with Fluo-4 AM and trapped in a microchamber array; (b) OR, olfactory receptor co-receptor (Orco), and fluorescent protein genes constructed on a plasmid and then transfected into Sf21 cells. (a) is reprinted from [82]. (b) is reprinted from [86] with permission from Elsevier.
Electronics 14 01852 g003
Figure 4. SPR-based odor biosensors: (a) cell directly adhered to the gold surface of an SPR chip, with the surface treated with poly-D-lysine for stable attachment; (b) OR protein carried by nanosomes immobilized via interactions of cmyc sequence with anti-cmyc monoclonal Ab attached to the gold surface; (c) OBP from honeybee immobilized on nanocup array surface with EDC and NHS; (d) OR protein embedded into a liposome and then fixed on the SPR surface; and (e) cell or membrane fragment immobilized on an SPR biosensor with EDC+NHS. (a) is reprinted from [106] with permission from Elsevier. (c) is reprinted from [102] with permission from Elsevier. (d) is reprinted from [105] with permission from Elsevier. (e) is reprinted from [103].
Figure 4. SPR-based odor biosensors: (a) cell directly adhered to the gold surface of an SPR chip, with the surface treated with poly-D-lysine for stable attachment; (b) OR protein carried by nanosomes immobilized via interactions of cmyc sequence with anti-cmyc monoclonal Ab attached to the gold surface; (c) OBP from honeybee immobilized on nanocup array surface with EDC and NHS; (d) OR protein embedded into a liposome and then fixed on the SPR surface; and (e) cell or membrane fragment immobilized on an SPR biosensor with EDC+NHS. (a) is reprinted from [106] with permission from Elsevier. (c) is reprinted from [102] with permission from Elsevier. (d) is reprinted from [105] with permission from Elsevier. (e) is reprinted from [103].
Electronics 14 01852 g004
Figure 5. FET-based odor biosensors: (a) CNTFET biosensor with OR protein immobilized on floating electrode above carbon nanotube; (b) fabrication procedure of graphene FET biosensor; (c) carboxylated polypyrrole nanoparticle FET; (d) hybridized DNA probe immobilized on FET for vanillin detection; and (e) cells loaded on extended-gate electrodes. (a) is reprinted from [114]. (b) is reprinted with permission from [120] (Copyright 2015 American Chemical Society). (c) is reprinted with permission from [121] (Copyright 2019 American Chemical Society). (d) is reprinted from [122] with permission from Elsevier. (e) is reprinted from [123].
Figure 5. FET-based odor biosensors: (a) CNTFET biosensor with OR protein immobilized on floating electrode above carbon nanotube; (b) fabrication procedure of graphene FET biosensor; (c) carboxylated polypyrrole nanoparticle FET; (d) hybridized DNA probe immobilized on FET for vanillin detection; and (e) cells loaded on extended-gate electrodes. (a) is reprinted from [114]. (b) is reprinted with permission from [120] (Copyright 2015 American Chemical Society). (c) is reprinted with permission from [121] (Copyright 2019 American Chemical Society). (d) is reprinted from [122] with permission from Elsevier. (e) is reprinted from [123].
Electronics 14 01852 g005
Figure 6. QCM-based odor biosensors: (a) plasma membrane fragment from HEK-293 cell expressing ODR-10 immobilized on QCM; (b) OBP expressed by E. coli immobilized on QCM with EDC+NHS; and (c) peptide mimicking the HarmOBP7 region as a receptor element immobilized on gold substrate of QCM. (a) is reprinted from [132] with permission from Elsevier. (b) is reprinted from [134]. (c) is reprinted from [135] with permission from Elsevier.
Figure 6. QCM-based odor biosensors: (a) plasma membrane fragment from HEK-293 cell expressing ODR-10 immobilized on QCM; (b) OBP expressed by E. coli immobilized on QCM with EDC+NHS; and (c) peptide mimicking the HarmOBP7 region as a receptor element immobilized on gold substrate of QCM. (a) is reprinted from [132] with permission from Elsevier. (b) is reprinted from [134]. (c) is reprinted from [135] with permission from Elsevier.
Electronics 14 01852 g006
Figure 7. SAW-based odor biosensors: (a) antibody as a sensing material in SAW biosensors and (b) OBP as a sensing material for SAW biosensors. (a) is reprinted from [145] with permission from Elsevier. (b) is reprinted from [147] with permission from Elsevier.
Figure 7. SAW-based odor biosensors: (a) antibody as a sensing material in SAW biosensors and (b) OBP as a sensing material for SAW biosensors. (a) is reprinted from [145] with permission from Elsevier. (b) is reprinted from [147] with permission from Elsevier.
Electronics 14 01852 g007
Figure 8. Other odor biosensors: (a) primary OSN transfected with plasmid containing the ODR-10 gene loaded on an LAPS chamber for diacetyl detection; (b) OR and Orco protein inserted into a bilayer lipid membrane for imitating the cell membrane, with a pair of electrodes at the bottom recording the current change when ions flow through the ion gate composed of OR and Orco. (a) is reprinted from [156] with permission from Elsevier. (b) is reprinted with permission from [159] (Copyright 2019 American Chemical Society).
Figure 8. Other odor biosensors: (a) primary OSN transfected with plasmid containing the ODR-10 gene loaded on an LAPS chamber for diacetyl detection; (b) OR and Orco protein inserted into a bilayer lipid membrane for imitating the cell membrane, with a pair of electrodes at the bottom recording the current change when ions flow through the ion gate composed of OR and Orco. (a) is reprinted from [156] with permission from Elsevier. (b) is reprinted with permission from [159] (Copyright 2019 American Chemical Society).
Electronics 14 01852 g008
Figure 10. Odor biosensors applied in disease diagnosis: (a) headspace gas from lung cancer cell line confirmed by GC-MS; (b) biomarker of lung cancer that could induce the response of some OSNs; (c) breath from volunteers measured for disease diagnosis, “√” means healthy and “×” means unhealthy; (d) ammonia in human breath indicating latent kidney problems; and (e) measuring the acetone concentration in exhaled breath for diabetic ketoacidosis diagnosis. (a) is reprinted from [93] with permission from Springer Nature. (b) is reprinted from [82]. (c) is reprinted from [179] with permission from Elsevier. (d) is reprinted from [180]. (e) is reprinted from [175] with permission from Elsevier.
Figure 10. Odor biosensors applied in disease diagnosis: (a) headspace gas from lung cancer cell line confirmed by GC-MS; (b) biomarker of lung cancer that could induce the response of some OSNs; (c) breath from volunteers measured for disease diagnosis, “√” means healthy and “×” means unhealthy; (d) ammonia in human breath indicating latent kidney problems; and (e) measuring the acetone concentration in exhaled breath for diabetic ketoacidosis diagnosis. (a) is reprinted from [93] with permission from Springer Nature. (b) is reprinted from [82]. (c) is reprinted from [179] with permission from Elsevier. (d) is reprinted from [180]. (e) is reprinted from [175] with permission from Elsevier.
Electronics 14 01852 g010
Figure 11. Odor biosensors applied for food safety: (a) nanodisks immobilized on the floating electrodes of a carbon nanotube FET for cadaverine detection; (b) experimental setup for Salmonella-related indicator alcohol detection. (a) is reprinted with permission from [115] (Copyright 2019 American Chemical Society). (b) is reprinted from [133] with permission from Elsevier.
Figure 11. Odor biosensors applied for food safety: (a) nanodisks immobilized on the floating electrodes of a carbon nanotube FET for cadaverine detection; (b) experimental setup for Salmonella-related indicator alcohol detection. (a) is reprinted with permission from [115] (Copyright 2019 American Chemical Society). (b) is reprinted from [133] with permission from Elsevier.
Electronics 14 01852 g011
Figure 12. Odor biosensors applied in security: (a) M72 OSN responding to trinitrotoluene, presented as a higher firing rate; (b) standard explosive solution dropped onto a filter, and, once the solvent is fully evaporated, the dry residue is thermally desorbed and sent to the detector. (a) is reprinted from [58] with permission from Elsevier. (b) is reprinted from [188] with permission from Elsevier.
Figure 12. Odor biosensors applied in security: (a) M72 OSN responding to trinitrotoluene, presented as a higher firing rate; (b) standard explosive solution dropped onto a filter, and, once the solvent is fully evaporated, the dry residue is thermally desorbed and sent to the detector. (a) is reprinted from [58] with permission from Elsevier. (b) is reprinted from [188] with permission from Elsevier.
Electronics 14 01852 g012
Table 1. Comparison of different transducers in odor biosensors.
Table 1. Comparison of different transducers in odor biosensors.
Transducer TypeDetection PrincipleSensing MaterialResponse TimeFeatureReferences
MicroelectrodeCurrent, potential, and impedance changeOSN, OB, and olfactory epitheliumAround ten secondsMultiple cells or sites; invasive condition with high sensitivity but short lifetime; non-invasive condition with easy operation but low sensitivity [39,73]
FluorescenceFluorescent intensity changeOSN and cell expressing ORAround half minuteMultiple cells; non-invasive with long lifetime; easily affected by ambient light[82,87,88]
SPRResonance wavelength changeCell, OR protein, and OBPSeveral secondsRapid and label-free detection; good stability and reproducibility; complex operation[103]
FETConductivity changeOR protein, OBP, and aptamerSeveral secondsVery high sensitivity and selectivity; complex fabrication[72,116,119]
QCMResonant frequency or dissipation changeOR protein and OBPSeveral secondsLow cost; high sensitivity; complex experimental setup[127,134]
SAWResonant frequency changeOR protein and OBPSeveral secondsHigh sensitivity; moderate reproducibility[140,141,143,144]
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

Deng, H.; Chen, Z.; Feng, P.; Tian, L.; Zong, H.; Nakamoto, T. Recent Advances and Applications of Odor Biosensors. Electronics 2025, 14, 1852. https://doi.org/10.3390/electronics14091852

AMA Style

Deng H, Chen Z, Feng P, Tian L, Zong H, Nakamoto T. Recent Advances and Applications of Odor Biosensors. Electronics. 2025; 14(9):1852. https://doi.org/10.3390/electronics14091852

Chicago/Turabian Style

Deng, Hongchao, Zhangyu Chen, Pengfei Feng, Lifeng Tian, Huijuan Zong, and Takamichi Nakamoto. 2025. "Recent Advances and Applications of Odor Biosensors" Electronics 14, no. 9: 1852. https://doi.org/10.3390/electronics14091852

APA Style

Deng, H., Chen, Z., Feng, P., Tian, L., Zong, H., & Nakamoto, T. (2025). Recent Advances and Applications of Odor Biosensors. Electronics, 14(9), 1852. https://doi.org/10.3390/electronics14091852

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