Next Article in Journal
Parallel Monitoring of Glucose, Free Amino Acids, and Vitamin C in Fruits Using a High-Throughput Paper-Based Sensor Modified with Poly(carboxybetaine acrylamide)
Previous Article in Journal
Optimization of Electrolytes with Redox Reagents to Improve the Impedimetric Signal for Use with a Low-Cost Analyzer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biosensors for Odor Detection: A Review

Laboratory for Future Interdisciplinary Research of Science and Technology, Institute of Innovative Research, Tokyo Institute of Technology, 4259 Nagatsuta-cho, Midori, Yokohama 226-8503, Kanagawa, Japan
*
Author to whom correspondence should be addressed.
Biosensors 2023, 13(12), 1000; https://doi.org/10.3390/bios13121000
Submission received: 27 October 2023 / Revised: 21 November 2023 / Accepted: 23 November 2023 / Published: 27 November 2023
(This article belongs to the Section Biosensor and Bioelectronic Devices)

Abstract

:
Animals can easily detect hundreds of thousands of odors in the environment with high sensitivity and selectivity. With the progress of biological olfactory research, scientists have extracted multiple biomaterials and integrated them with different transducers thus generating numerous biosensors. Those biosensors inherit the sensing ability of living organisms and present excellent detection performance. In this paper, we mainly introduce odor biosensors based on substances from animal olfactory systems. Several instances of organ/tissue-based, cell-based, and protein-based biosensors are described and compared. Furthermore, we list some other biological materials such as peptide, nanovesicle, enzyme, and aptamer that are also utilized in odor biosensors. In addition, we illustrate the further developments of odor biosensors.

1. Introduction

Odorants are widespread in the atmosphere. The odorants emitted from grain and fruit indicate the spoiled, toxic, or ripe conditions so animals hunt for suitable food [1,2]. Pheromones secreted from one insect can trigger an alarm [3,4], aggregation [5,6], territorial [7,8], or sexual activities [9] in other insects of the same species. A diseased individual releases a distinctive odor so the population can discover or isolate it [10,11]. Besides the natural sources of odorants, industrial activities and man-made objects also produce their unique smells which transmit essential information [12,13,14]. Thus, it is highly demanded that we must detect them precisely in many forms [15,16,17].
The olfactory system of human beings is already sophisticated. However, it is far from enough to detect all important odorants, owing to the insufficient detection range of human olfaction. In order to extend the detection range, detect harmful gases, objectively evaluate the gas type and intensity, and realize automatic measurement, plenty of gas sensors have been developed. Gas sensors can be divided into different categories according to their transducers such as field-effect transistor (FET) [18,19,20], quartz crystal microbalance (QCM) [21,22], surface acoustic wave (SAW) [23,24], surface plasmon resonance (SPR) [25,26], light-addressable potentiometric sensor (LAPS) [27], microelectrode array (MEA) [28,29], and fluorescence [30,31]. The sensing materials, e.g., carbon nanotube, polymer, carbon black composite, conducting polymer, lipid, or ionic liquid are utilized for odorant measurements. These conventional gas sensors have been widely used in our daily lives. However, researchers are still looking for even better sensors.
Besides the sensing materials listed above (we call them conventional materials), scientists fuse the biological materials with the transducers to form odor biosensors. The biological materials are biological substances such as antibodies, enzymes, nucleic acids, cells, epitheliums, nanovesicles, and so on. They also include some synthetic substances similar to biological substances, such as aptamers, peptides, and molecularly imprinted polymers. The biological materials used in odor biosensors are mainly extracted from the creatures’ olfactory systems, for example, olfactory epithelium [32], olfactory sensory neuron (OSN) [33], olfactory receptor (OR) protein [34,35,36], and odorant binding protein (OBP) [37,38]. These odor biosensors have higher sensitivity and selectivity towards their ligands than conventional odor sensors, and also they are not sensitive to temperature and humidity changes. Meanwhile, odor biosensors also encounter several problems such as a short lifetime, low reproducibility, complex operation, high cost, and so on. Thus, many scientists have been working towards better odor biosensors in recent decades (Figure 1).
Before introducing the detailed biosensors, we first briefly illustrated the sensing procedures of insect and vertebrate olfaction. Although the biological materials from other creatures such as C. elegans were also exploited in odor biosensors [53,54,55], most biosensor sensing materials were extracted from insects or vertebrates. Insects detect odorants through antennae and maxillary palp. The olfactory sensillum where OSNs exist is distributed along the olfactory organ (Figure 2a) [56,57,58]. The tiny pores on the olfactory sensillum allow odorant molecules to pass through and then dissolve into the sensillum lymph. Odorant molecules are carried by OBPs to OSN dendrites and then are captured by ORs. The ion channel formed by OR and olfactory receptor co-receptor (Orco) opens, resulting in a cation influx, and then a response signal is transmitted along the axon of OSN (Figure 2b) [59,60].
The vertebrate senses volatile odorants through the olfactory epithelium (Figure 2c). When an OR is activated by its ligand, adenylate cyclase III starts to transform the adenosine triphosphate (ATP) into cyclic adenosine monophosphate (cAMP). cAMP opens the cyclic nucleotide-gated channel which leads to calcium and sodium ion influx [61] (Figure 2d). The olfactory signal from OSN is first transferred in the olfactory bulb, and then sent to the brain. The sensing procedures demonstrate that OR and OBP can combine with the odorant molecules. Therefore, the biological materials that contain OR or OBP—for example, olfactory epithelium, OSN, OR protein, and OBP—are suitable for exploitation in odor biosensors.
Many review papers related to odor biosensors have been published before [62,63,64,65,66]. In this paper, biosensors’ principle and their advantages and disadvantages were described, while the previous review papers focused on limited points. Several odor biosensors were introduced based on the sensing materials such as organ/tissue, cell, protein, peptide, nanovesicle, enzyme, and aptamer (Figure 3). We briefly explained their sensing mechanisms and listed corresponding instances. The future developments of odor biosensors were also explained in the final section.

2. Organ/Tissue-Based Odor Biosensors

2.1. Antenna-Based Odor Biosensor

The antenna is an insect sensing organ for detecting volatile organic compounds (VOCs). In a typical antenna-based odor biosensor, the antenna is usually isolated from silk moth or honeybee [67,68,69]. Two electrodes are connected to the antenna, and the electroantennography (EAG) signal is recorded as the biosensor response.
Figure 4a is an example of a male silk moth antenna combined with a drone. The antenna was fixed on a circuit with a 50 Hz sampling rate, and it was more sensitive to bombykol (a kind of sex pheromone) than other odorants such as hexane and ethanol (Figure 4b). Hence, the EAG signal intensity represented the bombykol concentration. The odorant vapor evaporated from the 20 ng/L bombykol sample could induce a visible EAG signal, and the EAG peak emerged within 1 s of odor stimulation. An air pump flew the bombykol odor in a fixed direction to generate a stable odor plume. Based on the EAG signal intensity, the drone adjusted its yaw angle and flew toward the odor source. This bio-hybrid drone is an efficient platform for odor-source localization, and is also appropriate for field VOC detection.
The EAG can also be measured from live insects [70]. In this condition, the sensor lifetime would be much longer, and also it could sense 1 ppb target odorant with a 10 ms response time. However, the immobilization procedure is more complicated compared with only using the isolated antenna, and also the insect movement would introduce additional interfering signals. Therefore, the related research is less recent.

2.2. Olfactory-Epithelium- and Olfactory-Bulb-Based Odor Biosensors

The olfactory epithelium (usually rat) can be stripped off from the nasal cavity and cultured in vitro. Its electrophysiological signal relating to cellular functions can be recorded using a MEA or LAPS [44,71,72,73]. When combining the olfactory epithelium with the MEA or LAPS, the basal membrane side contacted the MEA (Figure 4c) or LAPS (Figure 4e) surface, and the cilia side was exposed for odor stimulation. The obtained signals in MEA (Figure 4d) and LAPS (Figure 4f) were potential spikes that looked quite similar. The MEA biosensor can record the multi-channel potential signal simultaneously, thus generating a spatiotemporal pattern of applied odorants [44]. As for the LAPS biosensor, it can detect the potential change on any site of the surface instead of being limited by the position of electrodes. But it only records the potential data in the laser-illuminated place, and also the olfactory epithelium culture on LAPS is more difficult than on the MEA surface.
Besides the in vitro condition, the in vivo sensing can be realized by implanting the microelectrodes into the rat olfactory bulb [74,75,76,77]. Researchers did not choose the olfactory epithelium because the surgery in that area was much more complicated than in the olfactory bulb. For inserting the microelectrodes, the fur over the skull was shaved, the scalp was incised, the skull was removed and then the olfactory bulb, and the related brain area was exposed. According to the received electrophysiological signal in the microelectrodes, the mitral/tufted cell layer was confirmed, and the microelectrodes were chronically fixed onto the rat’s head using dental cement (Figure 4g). After recovering for 4–5 days, this in vivo biosensor can be used for odor detection. The response patterns from all microelectrodes varied among different odorant stimulations. With the proper data-processing method, such as principal component analysis, the classification of four odors could be realized (Figure 4h).
Since there were multiple types of ORs in an insect antenna or a piece of olfactory epithelium, the electrophysiological signal captured in mitral/tufted cells may come from different glomeruli. To improve the specificity of odor detection, Van Der Pers et al. recorded the single-sensillum EAG signal through an extremely narrow glass capillary (Figure 5a) [39]. Gao et al. used the transgenic technique to handle the experiment’s mice, so the glomeruli that connected with M72 OSN in the olfactory bulb were highlighted with fluorescence (Figure 5b). The electrophysiological signal from this glomeruli could detect lower to a 10−5 M liquid sample trinitrotoluene and distinguish trinitrotoluene from other similarly structured chemicals [78]. Another method is using the bioengineering technique to overexpress OR3 on the rat olfactory epithelium and recording the electrophysiological signals from the olfactory bulb [79]. The detection limit could reach around 10−5 M towards four ligand odorants.
Figure 4. Organ/tissue-based odor biosensors. (a) Antenna combined with a drone as a portable biosensor; the voltage signal in the antenna was recorded using a circuit, and then used to control the drone. (b) Male silk moth antenna was more sensitive to bombykol than air, hexane, and ethanol. (c) Recording extracellular potentials of olfactory receptor neurons in intact epithelium with microelectrodes. (d) Tissue electrophysiological signals recorded using microelectrodes after the stimulation of acetic acid and butanedione. (e) LAPS system of the olfactory mucosa tissue cells on the sensor surface. (f) Tissue electrophysiological signals recorded using LAPS after the stimulation of butanedione and acetic acid. (g) Microelectrodes were implanted into the rat olfactory bulb as an in vivo biosensor; the recording region in the olfactory bulb dorsal surface was presented in the figure. (h) PCA plot for the classification of isoamyl acetate, banana, orange, and citral. (a,b) Reprinted with permission from Ref. [49]. Copyright 2021 Elsevier; (c,d) reprinted with permission from Ref. [72]. Copyright 2010 Elsevier; (e,f) reprinted with permission from Ref. [71]. Copyright 2010 Elsevier; and (g,h) reprinted with permission from Ref. [77]. Copyright 2015 Elsevier.
Figure 4. Organ/tissue-based odor biosensors. (a) Antenna combined with a drone as a portable biosensor; the voltage signal in the antenna was recorded using a circuit, and then used to control the drone. (b) Male silk moth antenna was more sensitive to bombykol than air, hexane, and ethanol. (c) Recording extracellular potentials of olfactory receptor neurons in intact epithelium with microelectrodes. (d) Tissue electrophysiological signals recorded using microelectrodes after the stimulation of acetic acid and butanedione. (e) LAPS system of the olfactory mucosa tissue cells on the sensor surface. (f) Tissue electrophysiological signals recorded using LAPS after the stimulation of butanedione and acetic acid. (g) Microelectrodes were implanted into the rat olfactory bulb as an in vivo biosensor; the recording region in the olfactory bulb dorsal surface was presented in the figure. (h) PCA plot for the classification of isoamyl acetate, banana, orange, and citral. (a,b) Reprinted with permission from Ref. [49]. Copyright 2021 Elsevier; (c,d) reprinted with permission from Ref. [72]. Copyright 2010 Elsevier; (e,f) reprinted with permission from Ref. [71]. Copyright 2010 Elsevier; and (g,h) reprinted with permission from Ref. [77]. Copyright 2015 Elsevier.
Biosensors 13 01000 g004
Among the organ/tissue-based odor biosensors, the sensing materials are easy to obtain. Meanwhile, the responses are to the original electrophysiological signals, and thus could be applied for exploring the creatures’ olfactory systems [80,81]. However, the lifetime of in vitro organ/tissue-based biosensors is relatively short. Although the in vivo organ/tissue-based biosensors extend the lifetime from several hours to several weeks, the surgery is complicated and the experimental results vary among different animals.

3. Cell-Based Odor Biosensor

The support cell and basal cell in the olfactory epithelium cannot respond to the target odorants. Also, the recorded electrophysiological signal usually comes from diverse OSNs. Therefore, using dissociated OSN or cell expressing OR as the sensing materials seems more efficient.

3.1. Olfactory Sensory-Neuron-Based Odor Biosensors

After dissecting the olfactory epithelium from the nasal cavity, the tissue was treated with papain, trypsin, or other protein enzymes to obtain the isolated cells (Figure 6a). Measuring the potential or labeling with a fluorescent dye (i.e., Fluo-4 AM) and then measuring the fluorescent-intensity change are two methods to detect the biosensor responses [46,82,83]. Since the diameter of OSN was too small to insert the electrodes, recording the potential signal was realized using several planar electrodes (Figure 6b) [83]. Around 10% of the dissociated cells were OSNs, and all the cells were plated onto the chip. The responses from the OSNs meant that close proximity to the sensing electrodes could be recorded, and the selectivity and sensitivity were determined using the ORs expressed on the corresponding OSNs. A circuit was fabricated on the chip to amplify the raw potential signal. The biosensor responses were presented as the increase of voltage spikes (Figure 6c), which was similar to the responses of olfactory-epithelium-based biosensors. Because the OSNs were randomly distributed on the chip, users cannot select the desired OSN type for a specific target odorant.
To solve this problem, Suzuki et al. designed a microchamber array chip [46]. The OSNs were labeled with calcium indicator Fluo-4 AM, added to the flow chamber on the microchamber array, and centrifuged briefly to trap the cells. The Ringer solution was circulated into the chip for removing the floating cells (Figure 6d). When several odorants were applied, the OSNs with responses could be picked out, and their OR types were analyzed. Therefore, this functional high-throughput OSN screening system was more efficient than the planar electrode method. In addition, some scientists cultured ORNs and olfactory bulb neurons together in vitro to sense and discriminate the odor stimulations as well as to serve as a novel model for studying the physiological and pathological mechanisms of olfaction [84].

3.2. Cell’s Expressing Olfactory-Receptor-Based Odor Biosensor

In 1989, Jones et al. proposed that the odorant-induced response in the olfactory sensory neuron is mediated by a G protein [85]. To find this protein, they screened the rat olfactory complementary deoxyribonucleic acid (cDNA) library and discovered a protein that related to the olfactory signal transduction [86]. In 1991, Buck et al. cloned and characterized 18 different members of a multigene family that encoded seven transmembrane domain proteins, i.e., OR proteins whose expression was restricted to the olfactory epithelium [87]. Those members were believed to encode a diverse family of odorant receptors. To better investigate the function of OR protein, Dahmen et al. injected the messenger ribonucleic acid (mRNA) isolated from rat or catfish olfactory epithelium into the Xenopus Oocytes [88]. After two to six days, the oocyte membrane current sometimes changed with odor stimulations. For enhancing the stability of the OR heterologous expression, Raming et al. transfected the Spodoptera frugiperda (Sf9) cell with rat OR DNA (OR5, OR12) via recombinant virus; this was the first time for OR stable heterologous expression [40].
The cells expressing OR has a similar function to OSN. Meanwhile, it can be passaged for many generations with steady characteristics, and the OR types can be freely determined by the researchers [51,52,89,90]. Table 1 lists several examples of cells expressing OR-based odor biosensors. The OR can be expressed on a large cell (such as Xenopus oocytes) for easier operation [43] or a small cell (such as HEK293) for higher density [46]. According to the target odorants, researchers can select the sensitive ORs from the OR library [47]. The sensing method can be invasive or non-invasive, simple or completed based on the measurement requirement [43,91].
The development of cells expressing OR-based odor biosensors focuses on extending the detection range from the liquid phase into the gas phase and increasing the OR types. Cells require an aqueous environment for maintaining their viability, so these biosensors were developed to detect liquid-phase odorants in the beginning [92,93,94]. Considering that most odorants exist in the gas phase, developing gas-phase odor biosensors is quite important for practical application. There are several methods for gas-phase odorant detection: first was waiting for the gas to naturally dissolve into the buffer medium (Figure 7a), and the buffer volume was enough to do so during the experiment period (>45 min) [47,95]; second was covering the cell with a thin liquid layer (~150 μm), and the time required for odorant molecules to penetrate the liquid film became shorter, thus significantly shortening the response time (Figure 7b) [45,96]; third was fixing the cells into a collagen pillar (Figure 7c), and the buffer medium in the box compensated for the evaporated water from the collagen gel [97]; and fourth was designing a special experiment chamber (Figure 7d), and the cells adhered on a polycarbonate membrane with a 2 μm pore, and the side with the cells touched the culture medium, while the other side was exposed to the gas-phase odorants [98]. Among these methods, the natural dissolution was slow but easy, the collagen pillar and special experiment chamber were complicated but stable, and the response from covering with a thin liquid film was fast but the biosensor lifetime was relatively short (~11 min).
Table 1. Examples of cells expressing OR-based odor biosensors.
Table 1. Examples of cells expressing OR-based odor biosensors.
OR TypeExpressed onOdorantSensing MethodOdorant ConcentrationImportanceRef.
Human
OR 5
E. coliLilialFluorescence0.2–1 mMGlutathione S-transferase can improve the OR expression level[99]
Mouse OR-EGHEK293EugenolFluorescence0.01–3 mMReconstituted mouse OR in HEK293 cell has a similar detection function to the original one[100]
Rat
I7
YeastOctyl aldehydeFluorescence10–50 μMScreen the proper OR that is sensitive to a specific odorant[101]
Drosophila melanogaster
Or85b
Xenopus oocytes2-HeptanoeElectrode10–1000 nMBuild a highly sensitive portable odor biosensor[43]
Caenorhabditis elegans
ODR-10
HEK293DiacetylLAPS10–100 nMLabel-free functional assays of olfactory receptor[102]
Rat OR I7HEK293OctanalSPR0.1–100 mMMeasure molecular interactions in realtime without any labeling[103]
Caenorhabditis elegans
ODR-10
MCF-7DiacetylSAW10−10–10−4 mMBuild a highly sensitive odor biosensor[104]
Rat OR I7HEK293OctanalQCM10−8–100 mMFind a linear relationship between response and the odorant concentration logarithmic value[105]
Silk moth
BmOR3
Sf21BombykalFET1–10 μMExplore the suitable surface for the cell expressing OR[106]
Increasing the OR types can form a larger sensor array, thereby enhancing the sensing capability. To achieve this target, we can fix the cells in the specific positions. For example, Figueroa et al. produced a microfluidic microwell array to trap different types of OR (Figure 8a) [82]; Misawa et al. arranged multiple cells in a fluidic system (Figure 8b) [43]; and Termtanasombat et al. immobilized the same type of cells in corresponding square areas (Figure 8c) [107]. These operations required pretreatment of the measurement area, but the response data were easy to obtain. We also can mix the cells and then supply several single-component odorants to label the OR types of sensitive cells [108]. In this condition, no pretreatment was required, while the difficulty in data processing increased slightly.

4. Protein-Based Odor Biosensor

Although the OSNs or cells expressing ORs can sense the odorants with high selectivity and sensitivity, and most substances in the cells are irrelevant to odorant detection. Therefore, directly using proteins such as OR protein and OBP as the sensing elements is more efficient.

4.1. OR Protein-Based Odor Biosensors

The combination of OR protein and its ligands is fundamental to odor detection. There are vast types of ORs existing in various animals such as the pig, honeybee, fruit fly, human, and mosquito [109,110]. For a determined target odorant, we always find one or several ORs that are capable of sensing it. Therefore, OR protein-based biosensors should be the most widely used odor biosensors.
To fabricate an OR protein-based odor biosensor, the first step is obtaining the sensing material. The OR protein concentration in OSN is not so high and obtaining large amounts of OSNs with the desired OR type is difficult. There are two ways to acquire the required OR protein: first is through heterologous protein expression, and second is through cell-free synthesis. In heterologous protein expression, the expression vector containing the OR DNA is established and then introduced into the cells, e.g., E. coli and HEK293 for expressing the OR [111,112,113]. The cells are incubated in the culture medium for a period. At this moment, the cells are already suitable for using in cell expressing OR-based odor biosensors. However, for getting the OR protein, these cells are lysed using sonication, and the insoluble fractions are collected, or handled by a membrane-protein-extraction kit to extract the functional protein. As for the cell-free synthesis, researchers only need to add the DNA template and reagents into the device, then it will automatically produce the target protein [114,115,116,117,118]. This method can avoid the issues such as protein aggregation and cytotoxicity which are usually encountered in heterologous expression. However, the generated OR protein is mixed with other components, so centrifugation and purification operations are necessary to reach the final product.
The resulting OR protein can be utilized in the original format (Figure 9a), inserted into a nanodisc (Figure 9b), or embedded into a bilayer lipid membrane (Figure 9c). The original format of OR protein or cell-plasma-membrane fragment was directly immobilized on the transducer surface for capturing the target odorant molecules [111,112,113]. A nanodisc was believed to be more stable than the original OR protein; a nanodisc was composed of a receptor, a lipid bilayer, and membrane scaffold proteins [119,120,121,122]. To construct a nanodisc, the lipids were mixed and solubilized with HEPES buffer, and then the purified OR protein was added, followed by membrane scaffold protein. After incubating for several hours and removing the unbound units, the nanodisc was collected. To embed the OR protein into a bilayer lipid membrane, two kinds of lipid were mixed, and then 1% agarose gel and buffer solution were added, thus forming a bilayer lipid membrane [123,124,125]. Then, the OR protein or OR/Orco complex was added to the bilayer lipid membrane to finish the embedding operation. The basic sensing mechanism of the original OR protein and nanodisc is mass or conformation change. The structure of bilayer lipid membrane with OR is similar to the cell, so the presence of Orco can improve the performance of biosensors [126,127], and also the biosensor response presents as the current change [124].
The OR proteins can combine with diverse transducers to form odor biosensors. FET [113,119,120,121,122], interdigitated microelectrode array [111], SAW [104], QCM [55,112], SPR [128], electrochemical impedance spectroscopy (EIS) [126,127,129,130], and a planar electrode pair [123,124,125] are all suitable for measuring the biosensor responses (Figure 10). Among them, the research in the last decade is mainly focused on the combination of OR proteins with FET sensors owing to their excellent sensitivity. Unlike the cells expressing OR, which can automatically adhere to the surface of the transducer, researchers should use some immobilization methods for fixing the OR protein. The easier way is using physical absorption, and the protein or membrane fraction is suspended in the solution and then evenly spread on the sensing area of the transducer [55], but the time required is relatively long, and the stability is poor. Another way is using chemical covalent binding, scientists choose the proper material to connect the OR protein and the transducer surface [112]; this method is commonly used most recently because of its high stability [113,119,120,122].

4.2. OBP-Based Odor Biosensors

OBP is a small soluble protein that transports the odorant molecules through the aqueous mucus or sensillum lymph [131,132,133,134,135,136,137]. Although the types of OBP [136,138,139,140] are less than OR [141] in the same animal, there are still many OBP-based odor biosensors [142,143,144,145,146,147,148,149,150]. The purification of OBPs is easier than OR proteins because they are secreted into the culture medium rather than remaining in the cell. The transducers and sensing procedures in OBP-based odor biosensors are similar to the OR protein-based odor biosensors, so we will not go into detail here.
On the other hand, OBP can be utilized to enhance odorant detection in OR protein-based odor biosensors. Ko et al. inserted the rat OBP3 into a mammalian expression vector pcDNA3 and then transfected it into HEK293 cell [151]. The HEK293-expressing rat OR I7 was sensitive to octanal, and the addition of OBP3 could enhance the responses. A similar conclusion was also proposed by Fukutani et al. regarding using silkworm moth OBP to improve the mouse OR sensitivity [152]. Recently, Choi et al. employed the rat OBP3 as a transporter for insoluble odorant molecules in a buffer medium [50]. The rat OR I7 was embedded into a nanodisc and immobilized on a carbon nanotube FET (CNT-FET). When gas-phase octanal was applied, the existence of OBP can enhance the response of this odor biosensor (Figure 11).
A comparison of different types of odor biosensors is presented in Table 2. We can select appropriate odor biosensors based on the specific measurement targets.
Table 2. Comparison of different odor biosensors.
Table 2. Comparison of different odor biosensors.
TypeSensing MaterialAdvantagesDisadvantagesRef.
Organ/TissueAntennaLow cost
Good sensitivity
Low selectivity
Low reproducibility
Short lifetime
[39,69]
Olfactory epithelium/bulbLow cost
Multi-channel data
Complex operation
Short lifetime
[72,153]
CellOSNEasy to form a large sensor array
Good sensitivity and selectivity
Hard to obtain the desired OSN type
Unable to subculture
[46,83]
Cell expressing ORLow cost
Easy for use
Good sensitivity and selectivity
Stable characteristic
Hard to obtain a favorable cell line
Large individual difference
[47,154]
ProteinOR proteinHigh sensitivity and selectivity
Easy to combine with transducers
Hard to purify
High cost
[119,125]
OBPEasy to purify
Good sensitivity
Easy to combine with transducers
High cost
Low selectivity
[147,150]

5. Other Biological Materials for Odor Biosensors

Besides the aforementioned sensing materials, some other biological substances that can specifically bind to odorant molecules are also suitable for odor biosensors.
First are the peptide-based odor biosensors. Although the OR protein or OBP are proper sensing materials for high sensitivity and selectivity odor biosensors, the manufacturing and purification processes of those proteins are complicated and labor intensive, also maintaining the stable structures of these proteins is relatively difficult. Considering that the part that binds the target odorant molecule is only a tiny portion of protein, using the OR- or OBP-derived peptide is a more effective solution for odor biosensors [155,156,157]. Lim et al. manufactured a single-walled-carbon nanotube FET (SWNT-FET) functionalized with OR-derived peptides (Figure 12a). Functionalization was performed using the property of SWNT for which aromatic rings were stacked on the surface using π–π interactions [158]. This biosensor was able to sensitively and selectively detect trimethylamine at a concentration of 10 fM and discriminate TMA from other similar molecules in real-time. In addition, peptides can be freely designed by the researchers. Homma et al. [159] designed two peptides and immobilized them, as well as a molecular scaffold peptide, on the graphene surface to form three FET biosensors. These biosensors could detect 10 pM limonene and discriminate different odorants. But, the desorption was not so successful after odor stimulation.
Second is the nanovesicle-based odor biosensors. Nanovesicle is secreted from the cell for intercellular communication. The protein, DNA, and RNA in the nanovesicle are the same as the original cell. Therefore, a nanovesicle from a cell heterologously expressing OR also has the same OR protein on its membrane and could be exploited for odor detection [160,161]. Jin et al. transfected the HEK293 cell with human OR 2AG1 and then collected the nanovesicle (Figure 12b) [160]. Because the OR protein was embedded in the bilayer lipid membrane and the nanovesicle was much smaller than the original cell, the OR function was stable and the nanovesicle could be easily combined with CNT-FET. This biosensor could sense amyl butyrate at 1 fM concentration. With the handling of fluorescent dye Fura 2-AM, the biosensor-response signal is also presented as fluorescent-intensity change.
Third is the enzyme-based odor biosensors. In reference [162], a flow-cell with nicotinamide adenine dinucleotide (NADH)-dependent secondary alcohol dehydrogenase (S-ADH) immobilized membrane was attached onto a fiber-optic NADH measurement device to form the fiber-optic biochemical-gas-sensing system (Figure 12c). The enzymatic reaction of acetone and NADH was evaluated using fluorescent-intensity change, and this system could measure the acetone gas from 20 to 5300 ppb.
Fourth is the aptamer-based biosensors. Aptamer can be RNA, single-stranded DNA, or double-stranded DNA. The combination of the aptamer and target molecule is similar to that of antigen–antibody. In reference [48], the aptamer was immobilized onto the surface of an ion-selective FET (ISFET). When vanillin odorant was applied, it diffused through the pores and dissolved into the buffer medium, and then was captured by the Van74 DNA aptamer (Figure 10d). Vanillin molecules replaced the hybridized probes on the sensitive surface resulting in an increase of surface potential, and this biosensor can detect vanillin over a concentration range from 2.7 ppt to 0.3 ppm.
Some other materials such as the taste-receptor protein [163,164] and antibody [42] were also employed for odor biosensors. Due to the space limitation, we cannot enumerate all types of odor biosensors here. The selection of biological materials should be determined based on the actual application situations.
Figure 12. Other biological materials for odor biosensors. (a) Peptide-based odor biosensor; peptides derived from OBP were assembled with a PDMS-based microfluidic layer between top and bottom frames. (b) Nanovesicle-based odor biosensor; the nanovesicle was generated from the cell expressing OR, and it had the same membrane proteins and cytosolic components as the original cell, while the size was much smaller. (c) Enzyme-based odor biosensor; the enzyme was immobilized on the flow cell, and the fluorescent-intensity change indicated the concentrations of gaseous acetone. (d) Aptamer-based odor biosensor; when vanillin combined with the aptamer on the ISFET, the surface potential increased. (a) Reprinted with permission from Ref. [158]. Copyright 2015 Elsevier. (b) Reprinted with permission from Ref. [160]. Copyright 2012 Elsevier. (c) Reprinted with permission from Ref. [162]. Copyright 2015 Elsevier. (d) Reprinted with permission from Ref. [48]. Copyright 2019 Elsevier.
Figure 12. Other biological materials for odor biosensors. (a) Peptide-based odor biosensor; peptides derived from OBP were assembled with a PDMS-based microfluidic layer between top and bottom frames. (b) Nanovesicle-based odor biosensor; the nanovesicle was generated from the cell expressing OR, and it had the same membrane proteins and cytosolic components as the original cell, while the size was much smaller. (c) Enzyme-based odor biosensor; the enzyme was immobilized on the flow cell, and the fluorescent-intensity change indicated the concentrations of gaseous acetone. (d) Aptamer-based odor biosensor; when vanillin combined with the aptamer on the ISFET, the surface potential increased. (a) Reprinted with permission from Ref. [158]. Copyright 2015 Elsevier. (b) Reprinted with permission from Ref. [160]. Copyright 2012 Elsevier. (c) Reprinted with permission from Ref. [162]. Copyright 2015 Elsevier. (d) Reprinted with permission from Ref. [48]. Copyright 2019 Elsevier.
Biosensors 13 01000 g012

6. Conclusions

The trends of odor biosensors are more efficient and compact, have higher sensitivity and selectivity, and a wider detection range. Tissues/organs, cells, and proteins are all capable of being the sensing materials of odor biosensors, but the smaller size is easier for combining with transducers, and the non-sensitive substance is less. The peptide seems to be the ultimate form of sensing material owing to its simplicity, diversity, and ease of production. The design of peptides could be inspired by the working area of the OR protein or OBP, and is also freely determined by scientists. Therefore, the latent detection ability of peptide should be higher than that of the existing OR protein or OBP. The sensitivity and selectivity of biosensors are mainly determined using the sensing material, while they are also affected by the transducer. Hence, the OR and OBP libraries are established for researchers to select the optimal sensing material for the specific ligand [165]. Meanwhile, the high-sensitive transducers, e.g., CNT-FET and QCM have been commonly used recently. The nanomaterials, such as the nanowire [166,167], nanoparticle [168,169,170], and nanotube [171,172], that have a large surface area to volume ratio can improve the sensitivity of biosensors, thereby becoming well-developed in recent years. Most odor biosensors work in the liquid phase, while most VOCs exist in the atmosphere. To extend the detection range, accelerating the VOC dissolution [173,174,175] and direct gas-phase odorant detection, there are two optional methods.

7. Future Perspectives

Although many odor biosensors have been developed, odor sensing towards complex odor mixtures has rarely been discussed. With the data from OR response towards odor mixture, we can build a response model that enables us to predict the OR responses under other odor stimulations, and finally predict the response of the animal olfaction to natural odorants, especially complex mixtures of numerous molecules.
The types of OR in one odor biosensor were usually less than four [176,177], which is much less compared to the OR types in animals. Meanwhile, a biosensor usually contains only one transducer. These facts result in the detection ability of current odor biosensors being completely unable to match animal olfactory perception [178]. A sensor array which consists of a set of sensors with various sensing materials can overcome the individual differences in biological material, has a more powerful sensing capability, and could be utilized for reconstructing the animal olfactory system [179]. Thus, manufacturing a large sensor array and through fusion of diverse transducers we are expected to improve biosensor performance [180]. In Figure 8, different types of cells expressing OR are put together to enhance the biosensor detection capability. Other biological materials such as aptamer, enzyme, OR protein, or OBP could also be assembled together to generate a sensor array. Also, diverse transducers could be connected in parallel (the unknown odor to be measured is separated into different channels and each channel has one transducer) or in serial (there is only one channel and all transducers are installed in this channel). In this condition, the biosensor is capable of mimicking the natural olfactory system [179]. On the other hand, the dimensions of the gathered experiment data will be much larger than the current situation [180]. Hence, we need machine learning methods for pattern recognition to gain insight into complex data. The machine learning methodologies are mainly utilized in three types of tasks: classification, clustering, and regression [181,182,183]. Scientists can effectively detect the key parameters or hidden patterns with the assistance of machine learning. The study of odor biosensors is still in the early stages. With the deepening of research, more powerful biosensors will be developed to contribute to our daily lives.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Danchuk, A.I.; Komova, N.S.; Mobarez, S.N.; Doronin, S.Y.; Burmistrova, N.A.; Markin, A.V.; Duerkop, A. Optical sensors for determination of biogenic amines in food. Anal. Bioanal. Chem. 2020, 412, 4023–4036. [Google Scholar] [CrossRef] [PubMed]
  2. Weston, M.; Phan, M.A.T.; Arcot, J.; Chandrawati, R. Anthocyanin-based sensors derived from food waste as an active use-by date indicator for milk. Food Chem. 2020, 326, 127017. [Google Scholar] [CrossRef]
  3. Saraiva, L.R.; Kondoh, K.; Ye, X.; Yoon, K.H.; Hernandez, M.; Buck, L.B. Combinatorial effects of odorants on mouse behavior. Proc. Natl. Acad. Sci. USA 2016, 113, E3300–E3306. [Google Scholar] [CrossRef] [PubMed]
  4. Dewaele, A.; Badonnel, K.; Persuy, M.A.; Durieux, D.; Bombail, V.; Favreau-Peigné, A.; Baly, C. Effect of environmental exposure to a maternally-learned odorant on anxiety-like behaviors at weaning in mice. Anim. Cogn. 2020, 23, 881–891. [Google Scholar] [CrossRef] [PubMed]
  5. McBride, C.S.; Baier, F.; Omondi, A.B.; Spitzer, S.A.; Lutomiah, J.; Sang, R.; Ignell, R.; Vosshall, L.B. Evolution of mosquito preference for humans linked to an odorant receptor. Nature 2014, 515, 222–227. [Google Scholar] [CrossRef] [PubMed]
  6. Lahondère, C.; Vinauger, C.; Okubo, R.P.; Wolff, G.H.; Chan, J.K.; Akbari, O.S.; Riffell, J.A. The olfactory basis of orchid pollination by mosquitoes. Proc. Natl. Acad. Sci. USA 2020, 117, 708–716. [Google Scholar] [CrossRef] [PubMed]
  7. Hansson, B.S.; Stensmyr, M.C. Evolution of insect olfaction. Neuron 2011, 72, 698–711. [Google Scholar] [CrossRef]
  8. Pelletier, J.; Xu, P.; Yoon, K.S.; Clark, J.M.; Leal, W.S. Odorant receptor-based discovery of natural repellents of human lice. Insect Biochem. Mol. Biol. 2015, 66, 103–109. [Google Scholar] [CrossRef]
  9. Dweck, H.K.M.; Ebrahim, S.A.M.; Thoma, M.; Mohamed, A.A.M.; Keesey, I.W.; Trona, F.; Lavista-Llanos, S.; Svatoš, A.; Sachse, S.; Knaden, M.; et al. Pheromones mediating copulation and attraction in Drosophila. Proc. Natl. Acad. Sci. USA 2015, 112, E2829–E2835. [Google Scholar] [CrossRef]
  10. Sarolidou, G.; Tognetti, A.; Lasselin, J.; Regenbogen, C.; Lundström, J.N.; Kimball, B.A.; Garke, M.; Lekander, M.; Axelsson, J.; Olsson, M.J. Olfactory Communication of Sickness Cues in Respiratory Infection. Front. Psychol. 2020, 11, 1004. [Google Scholar] [CrossRef]
  11. Tognetti, A.; Williams, M.N.; Lybert, N.; Lekander, M.; Axelsson, J.; Olsson, M.J. Humans can detect axillary odor cues of an acute respiratory infection in others. Evol. Med. Public Health 2023, 11, 219–228. [Google Scholar] [CrossRef] [PubMed]
  12. Debeda, H.; Dulau, L.; Dondon, P.; Menil, F.; Lucat, C.; Massok, P. Development of a reliable methane detector. Sens. Actuators B Chem. 1997, 44, 248–256. [Google Scholar] [CrossRef]
  13. Darmastuti, Z.; Bur, C.; Möller, P.; Rahlin, R.; Lindqvist, N.; Andersson, M.; Schütze, A.; Spetz, A.L. SiC-FET based SO2 sensor for power plant emission applications. Sens. Actuators B Chem. 2014, 194, 511–520. [Google Scholar] [CrossRef]
  14. Liu, S.F.; Lin, S.; Swager, T.M. An Organocobalt-Carbon Nanotube Chemiresistive Carbon Monoxide Detector. ACS Sens. 2016, 1, 354–357. [Google Scholar] [CrossRef] [PubMed]
  15. Fu, X.-A.; Li, M.; Knipp, R.J.; Nantz, M.H.; Bousamra, M. Noninvasive detection of lung cancer using exhaled breath. Cancer Med. 2014, 3, 174–181. [Google Scholar] [CrossRef] [PubMed]
  16. 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] [PubMed]
  17. Dent, A.G.; Sutedja, T.G.; Zimmerman, P.V. Exhaled breath analysis for lung cancer. J. Thorac. Dis. 2013, 5 (Suppl. S5), S540. [Google Scholar]
  18. Kumar, M.A.; Jayavel, R.; Mahalingam, S.; Kim, J.; Atchudan, R. Detection of Interleukin-6 Protein Using Graphene Field-Effect Transistor. Biosensors 2023, 13, 834. [Google Scholar] [CrossRef]
  19. Kwak, H.T.; Kim, H.; Yoo, H.; Choi, M.; Oh, K.; Kim, Y.; Kong, B.D.; Baek, C.K. Hydrogen Fluoride Gas Sensor by Silicon Nanosheet Field-Effect Transistor. IEEE Sens. J. 2023, 23, 16545–16552. [Google Scholar] [CrossRef]
  20. Kim, Y.; Lee, D.; Van Nguyen, K.; Lee, J.H.; Lee, W.H. Optimization of Gas-Sensing Properties in Poly(triarylamine) Field-Effect Transistors by Device and Interface Engineering. Polymers 2023, 15, 3463. [Google Scholar] [CrossRef]
  21. Wang, L.; Song, J.; Ruan, C.; Xu, S.; Yu, C. MoS2 based dual mine gas disaster sensor that operates at room temperature. Sens. Actuators A Phys. 2023, 359, 114517. [Google Scholar] [CrossRef]
  22. Esmeryan, K.D.; Lazarov, Y.; Grakov, T.; Fedchenko, Y.I.; Vergov, L.G.; Staykov, S. Metal–Phenolic Film Coated Quartz Crystal Microbalance as a Selective Sensor for Methanol Detection in Alcoholic Beverages. Micromachines 2023, 14, 1274. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, L.; Che, J.; Cao, Q.; Shen, S.; Tang, Y. Highly Sensitive Surface Acoustic Wave H2S Gas Sensor Using Electron-beam-evaporated CuO as Sensitive Layer. Sens. Mater. 2023, 35, 2293–2304. [Google Scholar] [CrossRef]
  24. 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]
  25. Li, Y.; Chen, H.; Chen, Q.; Li, H.; Gao, Z. Surface plasmon resonance induced methane gas sensor in hollow core anti-resonant fiber. Opt. Fiber Technol. 2023, 78, 103293. [Google Scholar] [CrossRef]
  26. Sassi, I.A.; El Hadj Rhouma, M.B.; Daher, M.G. Highly sensitive refractive index gas sensor using two-dimensional silicon carbide grating based on surface plasmon resonance. Opt. Quantum Electron. 2023, 55, 402. [Google Scholar] [CrossRef]
  27. Ito, Y.; Morimoto, K.; Tsunoda, Y. Light-addressable potentiometric (LAP) gas sensor. Sens. Actuators B. Chem. 1993, 13, 348–350. [Google Scholar] [CrossRef]
  28. Kaaliveetil, S.; Lee, Y.; Li, Z.; Cheng, Y.; Menon, N.H.; Dongare, S.; Gurkan, B.; Basuray, S. Ionic Liquid-Packed Microfluidic Device with Non-Planar Microelectrode as a Miniaturized Electrochemical Gas Sensor. J. Electrochem. Soc. 2023, 170, 087508. [Google Scholar] [CrossRef]
  29. Anjos, T.G.; Hahn, C.E.W. The development of a membrane-covered microelectrode array gas sensor for oxygen and carbon dioxide measurement. Sens. Actuators B Chem. 2008, 135, 224–229. [Google Scholar] [CrossRef]
  30. Farooq, A.; Al-Jowder, R.; Narayanaswamy, R.; Azzawi, M.; Roche, P.J.R.; Whitehead, D.E. Gas detection using quenching fluorescence of dye-immobilised silica nanoparticles. Sens. Actuators B Chem. 2013, 183, 230–238. [Google Scholar] [CrossRef]
  31. Wang, B.; Deng, Z.; Fu, Y.; Kerman, S.; Xu, W.; Li, H.; Liu, H.; He, Q.; Chen, C.; Cheng, J. On-Chip Fluorescent Sensor for Chemical Vapor Detection. Adv. Mater. Technol. 2023, 8, 2300609. [Google Scholar] [CrossRef]
  32. 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]
  33. Wu, C.S.; Chen, P.H.; Yuan, Q.; Wang, P. Response enhancement of olfactory sensory neurons-based biosensors for odorant detection. J. Zhejiang Univ. Sci. B 2009, 10, 285–290. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, S.O.; Kim, S.G.; Ahn, H.; Yoo, J.; Jang, J.; Park, T.H. Ni-rGO Sensor Combined with Human Olfactory Receptor-Embedded Nanodiscs for Detecting Gas-Phase DMMP as a Simulant of Nerve Agents. ACS Sens. 2023, 8, 3095–3103. [Google Scholar] [CrossRef] [PubMed]
  35. Lan, K.; Liu, S.; Wang, Z.; Wei, J.; Qin, G. High-performance olfactory receptor-derived peptide sensor for trimethylamine detection on the pyramid substrate structure. Sens. Actuators A Phys. 2023, 358, 114452. [Google Scholar] [CrossRef]
  36. Kuroda, S.; Nakaya-Kishi, Y.; Tatematsu, K.; Hinuma, S. Human Olfactory Receptor Sensor for Odor Reconstitution. Sensors 2023, 23, 6164. [Google Scholar] [CrossRef] [PubMed]
  37. Chang, X.; Qiu, F.; Wang, C.; Qian, Y.; Zhang, Y.; Guo, Q.; Wang, Q.; Wang, S.; Yang, M.; Yu, J. Possibility for detecting 14 typical odorants occurring in drinking water by employing human odor-binding protein OBP2a. Environ. Sci. Eur. 2023, 35, 1–12. [Google Scholar] [CrossRef]
  38. Nakano-Baker, O.; Fong, H.; Shukla, S.; Lee, R.V.; Cai, L.; Godin, D.; Hennig, T.; Rath, S.; Novosselov, I.; Dogan, S.; 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]
  39. Van Der Pers, J.N.C.; Den Otter, C.J. Single cell responses from olfactory receptors of small ermine moths to sex-attractants. J. Insect Physiol. 1978, 24, 337–343. [Google Scholar] [CrossRef]
  40. Raming, K.; Krieger, J.; Strotmann, J.; Boekhoff, I.; Kubick, S.; Baumstark, C.; Breer, H. Cloning and expression of odorant receptors. Nature 1993, 361, 353–356. [Google Scholar] [CrossRef]
  41. Mitsubayashi, K.; Yokoyama, K.; Takeuchi, T.; Karube, I. Gas-Phase Biosensor for Ethanol. Anal. Chem. 1994, 66, 3297–3302. [Google Scholar] [CrossRef]
  42. 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]
  43. 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] [PubMed]
  44. 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]
  45. 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] [PubMed]
  46. 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] [PubMed]
  47. Kida, H.; Fukutani, Y.; Mainland, J.D.; de March, C.A.; Vihani, A.; Li, Y.R.; Chi, Q.; Toyama, A.; Liu, L.; Kameda, M.; et al. Vapor detection and discrimination with a panel of odorant receptors. Nat. Commun. 2018, 9, 4556. [Google Scholar] [CrossRef]
  48. 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]
  49. 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]
  50. Choi, D.; Lee, S.J.; Baek, D.; Kim, S.; Shin, J.; Choi, Y.; Cho, Y.; Bang, S.; Park, J.Y.; Lee, S.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]
  51. 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]
  52. 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]
  53. Lin, A.; Qin, S.; Casademunt, H.; Wu, M.; Hung, W.; Cain, G.; Tan, N.Z.; Valenzuela, R.; Lesanpezeshki, L.; Venkatachalam, V.; et al. Functional imaging and quantification of multineuronal olfactory responses in C. elegans. Sci. Adv. 2023, 9, eade1249. [Google Scholar] [CrossRef] [PubMed]
  54. Reilly, D.K.; Schwarz, E.M.; Muirhead, C.S.; Robidoux, A.N.; Narayan, A.; Doma, M.K.; Sternberg, P.W.; Srinivasan, J. Transcriptomic profiling of sex-specific olfactory neurons reveals subset-specific receptor expression in Caenorhabditis elegans. Genetics 2023, 223, iyad026. [Google Scholar] [CrossRef] [PubMed]
  55. Sung, J.H.; Ko, H.J.; Park, T.H. Piezoelectric biosensor using olfactory receptor protein expressed in Escherichia coli. Biosens. Bioelectron. 2006, 21, 1981–1986. [Google Scholar] [CrossRef] [PubMed]
  56. Kaissling, K.E. Chemo-electrical transduction in insect olfactory receptors. Annu. Rev. Neurosci. 1986, 9, 121–145. [Google Scholar] [CrossRef] [PubMed]
  57. Jacquin-Joly, E.; Merlin, C. Insect olfactory receptors: Contributions of molecular biology to chemical ecology. J. Chem. Ecol. 2004, 30, 2359–2397. [Google Scholar] [CrossRef] [PubMed]
  58. Ando, T.; Sekine, S.; Inagaki, S.; Misaki, K.; Badel, L.; Moriya, H.; Sami, M.M.; Itakura, Y.; Chihara, T.; Kazama, H. Nanopore Formation in the Cuticle of an Insect Olfactory Sensillum. Curr. Biol. 2019, 29, 1512–1520.e6. [Google Scholar] [CrossRef]
  59. Stengl, M.; Funk, N.W. The role of the coreceptor Orco in insect olfactory transduction. J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol. 2013, 199, 897–909. [Google Scholar] [CrossRef]
  60. Mukunda, L.; Lavista-Llanos, S.; Hansson, B.S.; Wicher, D. Dimerisation of the Drosophila odorant coreceptor Orco. Front. Cell. Neurosci. 2014, 8, 261. [Google Scholar] [CrossRef]
  61. Hayden, S.; Teeling, E.C. The Molecular Biology of Vertebrate Olfaction. Anat. Rec. 2014, 297, 2216–2226. [Google Scholar] [CrossRef] [PubMed]
  62. Mitsubayashi, K.; Toma, K.; Iitani, K.; Arakawa, T. Gas-phase biosensors: A review. Sens. Actuators B Chem. 2022, 367, 132053. [Google Scholar] [CrossRef]
  63. El Kazzy, M.; Weerakkody, J.S.; Hurot, C.; Mathey, R.; Buhot, A.; Scaramozzino, N.; 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] [PubMed]
  64. Hirata, Y.; Oda, H.; Osaki, T.; Takeuchi, S. Biohybrid sensor for odor detection. Lab Chip 2021, 21, 2643–2657. [Google Scholar] [CrossRef] [PubMed]
  65. 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]
  66. Brito, N.F.; Oliveira, D.S.; Santos, T.C.; Moreira, M.F.; Melo, A.C.A. Current and potential biotechnological applications of odorant-binding proteins. Appl. Microbiol. Biotechnol. 2020, 104, 8631–8648. [Google Scholar] [CrossRef] [PubMed]
  67. Oh, Y.; Lee, Y.; Heath, J.; Kim, M. Applications of animal biosensors: A review. IEEE Sens. J. 2014, 15, 637–645. [Google Scholar]
  68. Sakurai, T.; Mitsuno, H.; Haupt, S.S.; Uchino, K.; Yokohari, F.; Nishioka, T.; Kobayashi, I.; Sezutsu, H.; Tamura, T.; Kanzaki, R. A single sex pheromone receptor determines chemical response specificity of sexual behavior in the silkmoth Bombyx mori. PLoS Genet. 2011, 7, e1002115. [Google Scholar] [CrossRef]
  69. Pawson, S.M.; Kerr, J.L.; O’Connor, B.C.; Lucas, P.; Martinez, D.; Allison, J.D.; 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]
  70. Schütz, S.; Weißbecker, B.; Hummel, H.E. Biosensor for volatiles released by damaged plants. Biosens. Bioelectron. 1996, 11, 427–433. [Google Scholar] [CrossRef]
  71. 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. 2010, 146, 527–533. [Google Scholar] [CrossRef]
  72. 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]
  73. 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 bioelectronic nose. Biosens. Bioelectron. 2013, 43, 186–192. [Google Scholar] [CrossRef] [PubMed]
  74. Davison, I.G.; Katz, L.C. Sparse and selective odor coding by mitral/tufted neurons in the main olfactory bulb. J. Neurosci. 2007, 27, 2091–2101. [Google Scholar] [CrossRef] [PubMed]
  75. 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]
  76. Zhuang, L.; Hu, N.; Tian, F.; Dong, Q.; Hu, L.; Li, R.; Wang, P. 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]
  77. Zhuang, L.; Guo, T.; Cao, D.; Ling, L.; Su, K.; Hu, N.; Wang, P. Detection and classification of natural odors with an in vivo bioelectronic nose. Biosens. Bioelectron. 2015, 67, 694–699. [Google Scholar] [CrossRef]
  78. 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]
  79. 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]
  80. Dasgupta, D.; Warner, T.P.A.; Erskine, A.; Schaefer, A.T. Coupling of Mouse Olfactory Bulb Projection Neurons to Fluctuating Odor Pulses. J. Neurosci. 2022, 42, 4278–4296. [Google Scholar] [CrossRef]
  81. 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] [PubMed]
  82. 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 Chip 2010, 10, 1120–1127. [Google Scholar] [CrossRef]
  83. 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]
  84. 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] [PubMed]
  85. Jones, D.T.; Reed, R.R. Golf: An olfactory neuron specific-G protein involved in odorant signal transduction. Science 1989, 244, 790–795. [Google Scholar] [CrossRef] [PubMed]
  86. Pace, U.; Hanski, E.; Salomon, Y.; Lancet, D. Odorant-sensitive adenylate cyclase may mediate olfactory reception. Nature 1985, 316, 255–258. [Google Scholar] [CrossRef] [PubMed]
  87. Buck, L.; Axel, R. A novel multigene family may encode odorant receptors: A molecular basis for odor recognition. Cell 1991, 65, 175–187. [Google Scholar] [CrossRef]
  88. Dahmen, N.; Wang, H.-L.; Margolis, F.L. Expression of Olfactory Receptors in Xenopus Oocytes. J. Neurochem. 1992, 58, 1176–1179. [Google Scholar] [CrossRef]
  89. 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]
  90. 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]
  91. Lee, S.H.; Jun, S.B.; Ko, H.J.; Kim, S.J.; Park, T.H. Cell-based olfactory biosensor using microfabricated planar electrode. Biosens. Bioelectron. 2009, 24, 2659–2664. [Google Scholar] [CrossRef] [PubMed]
  92. Zhang, Y.; Chou, J.H.; Bradley, J.; Bargmann, C.I.; Zinn, K. The Caenorhabditis elegans seven-transmembrane protein ODR-10 functions as an odorant receptor in mammalian cells. Proc. Natl. Acad. Sci. USA 1997, 94, 12162–12167. [Google Scholar] [CrossRef] [PubMed]
  93. Krautwurst, D.; Yau, K.W.; Reed, R.R. Identification of ligands for olfactory receptors by functional expression of a receptor library. Cell 1998, 95, 917–926. [Google Scholar] [CrossRef]
  94. Gimelbrant, A.A.; Stoss, T.D.; Landers, T.M.; McClintock, T.S. Truncation releases olfactory receptors from the endoplasmic reticulum of heterologous cells. J. Neurochem. 1999, 72, 2301–2311. [Google Scholar] [CrossRef] [PubMed]
  95. 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]
  96. 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]
  97. Hirata, Y.; Morimoto, Y.; Takeuchi, S. Cell-laden micropillars detect gaseous odorants on a liquid-air interface. In Proceedings of the IEEE International Conference on Micro Electro Mechanical Systems (MEMS), Belfast, UK, 21–25 January 2018; IEEE: Piscataway, NJ, USA, 2018; Volume 2018, pp. 304–307. [Google Scholar]
  98. 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]
  99. Kiefer, H.; Krieger, J.; Olszewski, J.D.; Von Heijne, G.; Prestwich, G.D.; Breer, H. Expression of an olfactory receptor in Escherichia coli: Purification, reconstitution, and ligand binding. Biochemistry 1996, 35, 16077–16084. [Google Scholar] [CrossRef]
  100. Kajiya, K.; Inaki, K.; Tanaka, M.; Haga, T.; Kataoka, H.; Touhara, K. Molecular bases of odor discrimination: Reconstitution of olfactory receptors that recognize overlapping sets of odorants. J. Neurosci. 2001, 21, 6018–6025. [Google Scholar] [CrossRef]
  101. 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]
  102. 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]
  103. 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]
  104. 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] [PubMed]
  105. 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]
  106. Terutsuki, D.; Mitsuno, H.; Sakurai, T.; Okamoto, Y.; Tixier-Mita, A.; Toshiyoshi, H.; Mita, 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] [PubMed]
  107. 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. J. Chem. Ecol. 2016, 42, 716–724. [Google Scholar] [CrossRef] [PubMed]
  108. 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]
  109. Leal, W.S. Odorant Reception in Insects: Roles of Receptors, Binding Proteins, and Degrading Enzymes. Annu. Rev. Entomol. 2013, 58, 373–391. [Google Scholar] [CrossRef]
  110. Jones, P.L.; Pask, G.M.; Rinker, D.C.; Zwiebel, L.J. Functional agonism of insect odorant receptor ion channels. Proc. Natl. Acad. Sci. USA 2011, 108, 8821–8825. [Google Scholar] [CrossRef]
  111. Lee, S.H.; Kwon, O.S.; Song, H.S.; Park, S.J.; Sung, J.H.; Jang, J.; Park, T.H. Mimicking the human smell sensing mechanism with an artificial nose platform. Biomaterials 2012, 33, 1722–1729. [Google Scholar] [CrossRef]
  112. Du, L.; Wu, C.; Peng, H.; Zou, L.; Zhao, L.; Huang, L.; Wang, P. Piezoelectric olfactory receptor biosensor prepared by aptamer-assisted immobilization. Sens. Actuators B Chem. 2013, 187, 481–487. [Google Scholar] [CrossRef]
  113. 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.; et al. An Ultrasensitive, Selective, Multiplexed Superbioelectronic Nose That Mimics the Human Sense of Smell. Nano Lett. 2015, 15, 6559–6567. [Google Scholar] [CrossRef] [PubMed]
  114. Garenne, D.; Haines, M.C.; Romantseva, E.F.; Freemont, P.; Strychalski, E.A.; Noireaux, V. Cell-free gene expression. Nat. Rev. Methods Primers 2021, 1, 49. [Google Scholar] [CrossRef]
  115. Chen, F.; Wang, J.; Du, L.; Zhang, X.; Zhang, F.; Chen, W.; Cai, W.; Wu, C.; Wang, P. Functional expression of olfactory receptors using cell-free expression system for biomimetic sensors towards odorant detection. Biosens. Bioelectron. 2019, 130, 382–388. [Google Scholar] [CrossRef]
  116. Li, J.; Liu, X.; Man, Y.; Chen, Q.; Pei, D.; Wu, W. Cell-free expression, purification and characterization of Drosophila melanogaster odorant receptor OR42a and its co-receptor. Protein Expr. Purif. 2019, 159, 27–33. [Google Scholar] [CrossRef] [PubMed]
  117. Ritz, S.; Hulko, M.; Zerfaß, C.; May, S.; Hospach, I.; Krasteva, N.; Nelles, G.; Sinner, E.K. Cell-free expression of a mammalian olfactory receptor and unidirectional insertion into small unilamellar vesicles (SUVs). Biochimie 2013, 95, 1909–1916. [Google Scholar] [CrossRef] [PubMed]
  118. Kaiser, L.; Graveland-Bikker, J.; Steuerwald, D.; Vanberghem, M.; Herlihy, K.; Zhang, S. Efficient cell-free production of olfactory receptors: Detergent optimization, structure, and ligand binding analyses. Proc. Natl. Acad. Sci. USA 2008, 105, 15726–15731. [Google Scholar] [CrossRef] [PubMed]
  119. 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]
  120. 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]
  121. 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]
  122. Oh, J.; Yang, H.; Jeong, G.E.; Moon, D.; Kwon, O.S.; Phyo, S.; Lee, J.; Song, H.S.; Park, T.H.; 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]
  123. Fujii, S.; Nobukawa, A.; Osaki, T.; Morimoto, Y.; Kamiya, K.; Misawa, N.; Takeuchi, S. Pesticide vapor sensing using an aptamer, nanopore, and agarose gel on a chip. Lab Chip 2017, 17, 2421–2425. [Google Scholar] [CrossRef] [PubMed]
  124. 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] [PubMed]
  125. 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] [PubMed]
  126. 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] [PubMed]
  127. Khadka, R.; Aydemir, N.; Carraher, C.; Hamiaux, C.; Colbert, D.; Cheema, J.; Malmström, 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]
  128. 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]
  129. Hou, Y.; Jaffrezic-Renault, N.; Martelet, C.; Zhang, A.; Minic-Vidic, J.; Gorojankina, T.; Persuy, M.A.; Pajot-Augy, E.; Salesse, R.; Akimov, V. A novel detection strategy for odorant molecules based on controlled bioengineering of rat olfactory receptor I7. Biosens. Bioelectron. 2007, 22, 1550–1555. [Google Scholar] [CrossRef]
  130. 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]
  131. Matarazzo, V.; Zsurger, N.; Guillemot, J.C.; Clot-Faybesse, O.; Botto, J.M.; Dal Farra, C.; Crowe, M.; Demaille, J.; Vincent, J.P.; Mazella, J. Porcine odorant-binding protein selectively binds to a human olfactory receptor. Chem. Senses 2002, 27, 691–701. [Google Scholar] [CrossRef]
  132. Pevsner, J.; Hou, V.; Snowman, A.M.; Snyder, S.H. Odorant-binding protein. Characterization of ligand binding. J. Biol. Chem. 1990, 265, 6118–6125. [Google Scholar] [CrossRef]
  133. Steinbrecht, M.A. Odorant-Binding Proteins: Expression and Function. Ann. N. Y. Acad. Sci. 1998, 855, 323–332. [Google Scholar] [CrossRef] [PubMed]
  134. Wojtasek, H.; Leal, W.S. Conformational change in the pheromone-binding protein from Bombyx mori induced by pH and by interaction with membranes. J. Biol. Chem. 1999, 274, 30950–30956. [Google Scholar] [CrossRef] [PubMed]
  135. Vincent, F.; Spinelli, S.; Ramoni, R.; Grolli, S.; Pelosi, P.; Cambillau, C.; Tegoni, M. Complexes of porcine odorant binding protein with odorant molecules belonging to different chemical classes. J. Mol. Biol. 2000, 300, 127–139. [Google Scholar] [CrossRef] [PubMed]
  136. Tegoni, M.; Pelosi, P.; Vincent, F.; Spinelli, S.; Campanacci, V.; Grolli, S.; Ramoni, R.; Cambillau, C. Mammalian odorant binding proteins. Biochim. Et Biophys. Acta-Protein Struct. Mol. Enzymol. 2000, 1482, 229–240. [Google Scholar] [CrossRef] [PubMed]
  137. Campanacci, V.; Krieger, J.; Bette, S.; Sturgis, J.N.; Lartigue, A.; Cambillau, C.; Breer, H.; Tegoni, M. Revisiting the Specificity of Mamestra brassicae and Antheraea polyphemus Pheromone-binding Proteins with a Fluorescence Binding Assay. J. Biol. Chem. 2001, 276, 20078–20084. [Google Scholar] [CrossRef] [PubMed]
  138. Sankarganesh, D.; Kirkwood, R.N.; Meillour, P.N.-L.; Angayarkanni, J.; Achiraman, S.; Archunan, G. Pheromones, binding proteins, and olfactory systems in the pig (Sus scrofa): An updated review. Front. Vet. Sci. 2022, 9, 989409. [Google Scholar] [CrossRef]
  139. Li, M.Y.; Jiang, X.Y.; Qi, Y.Z.; Huang, Y.J.; Li, S.G.; Liu, S. Identification and expression profiles of 14 odorant-binding protein genes from Pieris rapae (lepidoptera: Pieridae). J. Insect Sci. 2020, 20, 2. [Google Scholar] [CrossRef]
  140. Capo, A.; Cozzolino, S.; Cavallari, A.; Bruno, U.; Calabrese, A.; Pennacchio, A.; Camarca, A.; Staiano, M.; D’auria, S.; Varriale, A. The Porcine Odorant-Binding Protein as a Probe for an Impedenziometric-Based Detection of Benzene in the Environment. Int. J. Mol. Sci. 2022, 23, 4039. [Google Scholar] [CrossRef]
  141. Kaupp, U.B. Olfactory signalling in vertebrates and insects: Differences and commonalities. Nat. Rev. Neurosci. 2010, 11, 188–200. [Google Scholar] [CrossRef]
  142. Forstner, M.; Breer, H.; Krieger, J. A receptor and binding protein interplay in the detection of a distinct pheromone component in the silkmoth Antheraea polyphemus. Int. J. Biol. Sci. 2009, 5, 745–757. [Google Scholar] [CrossRef] [PubMed]
  143. 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] [PubMed]
  144. 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. 2012, 41, 328–334. [Google Scholar]
  145. Lu, Y.; Li, H.; Zhuang, S.; Zhang, D.; Zhang, Q.; Zhou, J.; Dong, S.; Liu, Q.; Wang, P. Olfactory biosensor using odorant-binding proteins from honeybee: Ligands of floral odors and pheromones detection by electrochemical impedance. Sens. Actuators B Chem. 2014, 193, 420–427. [Google Scholar] [CrossRef]
  146. Zhang, Q.; Li, S.; Huang, Y.; Yao, Y.; Zhang, D.; Luo, S.; Lu, Y.; Liu, Q. Impedance spectroscopy analysis of human odorant binding proteins immobilized on nanopore arrays for biochemical detection. Biosens. Bioelectron. 2015, 79, 251–257. [Google Scholar]
  147. Larisika, M.; Kotlowski, C.; Steininger, C.; Mastrogiacomo, R.; Pelosi, P.; Schütz, S.; Peteu, S.F.; Kleber, C.; Reiner-Rozman, C.; Nowak, C.; et al. Electronic Olfactory Sensor Based on A. mellifera Odorant-Binding Protein 14 on a Reduced Graphene Oxide Field-Effect Transistor. Angew. Chem. Int. Ed. 2015, 54, 13245–13248. [Google Scholar] [CrossRef] [PubMed]
  148. Di Pietrantonio, F.; Benetti, M.; Cannatà, D.; Verona, E.; Palla-Papavlu, A.; Fernández-Pradas, J.M.; Serra, P.; Staiano, M.; Varriale, A.; D’Auria, S. A surface acoustic wave bio-electronic nose for detection of volatile odorant molecules. Biosens. Bioelectron. 2015, 67, 516–523. [Google Scholar] [CrossRef]
  149. Gao, A.; Wang, Y.; Zhang, D.; He, Y.; Zhang, L.; Liu, Y.; Wang, Y.; Song, H.; Li, T. Highly sensitive and selective detection of human-derived volatile organic compounds based on odorant binding proteins functionalized silicon nanowire array. Sens. Actuators B Chem. 2020, 309, 127762. [Google Scholar] [CrossRef]
  150. Scorsone, E.; Manai, R.; Cali, K.; Ricatti, M.J.; Farno, S.; Persaud, K.; Mucignat, C. Biosensor array based on ligand binding proteins for narcotics and explosives detection. Sens. Actuators B Chemical. 2021, 334, 129587. [Google Scholar] [CrossRef]
  151. Ko, H.J.; Park, T.H. Enhancement of odorant detection sensitivity by the expression of odorant-binding protein. Biosens. Bioelectron. 2008, 23, 1017–1023. [Google Scholar] [CrossRef]
  152. Fukutani, M.Y.Y.; Hori, A.; Tsukada, S.; Sato, R.; Ishii, J.; Kondo, A.; Matsunami, H. Improving the odorant sensitivity of olfactory receptor-expressing yeast with accessory proteins. Anal. Biochem. J. 2015, 35, 46–53. [Google Scholar] [CrossRef] [PubMed]
  153. 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]
  154. 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]
  155. 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] [PubMed]
  156. Wasilewski, T.; Szulczyński, B.; Wojciechowski, M.; Kamysz, W.; Gębicki, J. A highly selective biosensor based on peptide directly derived from the harmOBP7 aldehyde binding site. Sensors 2019, 19, 4284. [Google Scholar] [CrossRef] [PubMed]
  157. Lim, J.H.; Park, J.; Ahn, J.H.; Jin, H.J.; Hong, S.; Park, T.H. A peptide receptor-based bioelectronic nose for the real-time determination of seafood quality. Biosens. Bioelectron. 2013, 39, 244–249. [Google Scholar] [CrossRef] [PubMed]
  158. Lee, S.H.; Lim, J.H.; Park, J.; Hong, S.; Park, T.H. Bioelectronic nose combined with a microfluidic system for the detection of gaseous trimethylamine. Biosens. Bioelectron. 2015, 71, 179–185. [Google Scholar] [CrossRef]
  159. 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. 2023, 224, 115047. [Google Scholar] [CrossRef]
  160. Jin, H.J.; Lee, S.H.; Kim, T.H.; Park, J.; Song, H.S.; Park, T.H.; Hong, S. Nanovesicle-based bioelectronic nose platform mimicking human olfactory signal transduction. Biosens. Bioelectron. 2012, 35, 335–341. [Google Scholar] [CrossRef]
  161. 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]
  162. 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]
  163. Son, M.; Kim, D.; Ko, H.J.; Hong, S.; Park, T.H. A portable and multiplexed bioelectronic sensor using human olfactory and taste receptors. Biosens. Bioelectron. 2017, 87, 901–907. [Google Scholar] [CrossRef]
  164. Son, M.; Park, T.H. The bioelectronic nose and tongue using olfactory and taste receptors: Analytical tools for food quality and safety assessment. Biotechnol. Adv. 2018, 36, 371–379. [Google Scholar] [CrossRef]
  165. Münch, D.; Galizia, C.G. DoOR 2.0—Comprehensive Mapping of Drosophila melanogaster Odorant Responses. Sci. Rep. 2016, 6, 21841. [Google Scholar] [CrossRef] [PubMed]
  166. Ambhorkar, P.; Wang, Z.; Ko, H.; Lee, S.; Koo, K.I.; Kim, K.; Cho, D.I.D. Nanowire-based biosensors: From growth to applications. Micromachines 2018, 9, 679. [Google Scholar] [CrossRef] [PubMed]
  167. Smith, A.F.; Smith, A.F.; Liu, X.; Woodard, T.L.; Fu, T.; Emrick, T.; Jiménez, J.M.; Lovley, D.R.; Yao, J. Bioelectronic protein nanowire sensors for ammonia detection. Nano Res. 2020, 13, 1479–1484. [Google Scholar] [CrossRef]
  168. Selimoğlu, F.; Gür, B.; Ayhan, M.E.; Gür, F.; Kalita, G.; Tanemura, M.; Alma, M.H. Silver nanoparticle doped graphene-based impedimetric biosensor towards sensitive detection of procalcitonin. Mater. Chem. Phys. 2023, 297, 127339. [Google Scholar] [CrossRef]
  169. Tshobeni, Z.Z.; January, J.L.; Ngema, N.P.P.; Jijana, A.N.; Iwuoha, E.I.; Mulaudzi, T.; Douman, S.F.; Ajayi, R.F. Thioglycolic acid-capped gold nanoparticle/cytochrome P450-2E1 electrochemical biosensor for isoniazid. Sens. Bio-Sens. Res. 2023, 41, 100583. [Google Scholar] [CrossRef]
  170. Zhou, J.; Xiao, F.; Fu, J.; Jia, N.; Huang, X.; Sun, C.; Liu, C.; Huan, H.; Wang, Y. Rapid detection of monkeypox virus by multiple cross displacement amplification combined with nanoparticle-based biosensor platform. J. Med. Virol. 2023, 95, e28479. [Google Scholar] [CrossRef]
  171. Edwards, C.; Duanghathaipornsuk, S.; Goltz, M.; Kanel, S.; Kim, D.S. Peptide nanotube encapsulated enzyme biosensor for vapor phase detection of malathion, an organophosphorus compound. Sensors 2019, 19, 3856. [Google Scholar] [CrossRef]
  172. Zhang, Y.; Xiao, P.; Zhou, X.; Liu, D.; Garcia, B.B.; Cao, G. Carbon monoxide annealed TiO2 nanotube array electrodes for efficient biosensor applications. J. Mater. Chem. 2009, 19, 948–953. [Google Scholar] [CrossRef]
  173. Terutsuki, D.; Mitsuno, H.; Sato, K.; Sakurai, T.; Mase, N.; Kanzaki, R. Highly effective volatile organic compound dissolving strategy based on mist atomization for odorant biosensors. Anal. Chim. Acta 2020, 1139, 178–188. [Google Scholar] [CrossRef] [PubMed]
  174. Warden, A.C.; Trowell, S.C.; Gel, M. A Miniature Gas Sampling Interface with Open Microfluidic Channels: Characterization of Gas-to-Liquid Extraction Efficiency of Volatile Organic Compounds. Micromachines 2019, 10, 486. [Google Scholar] [CrossRef] [PubMed]
  175. Kida, H.; Tsukada, S.; Tagawa, Y.; Sato, R.; Kameda, M. Effective collection of volatile organic compounds in water using rimming flow with odorant-binding proteins. Mech. Eng. J. 2015, 2, 15-00244. [Google Scholar] [CrossRef]
  176. Yokoshiki, Y.; Nagayoshi, K.; Mitsuno, H.; Niki, S.; Tokuda, T.; Nakamoto, T. Improving Performance of Odor Biosensor System Using Olfactory Receptor. Sens. Mater. 2022, 34, 1617–1626. [Google Scholar] [CrossRef]
  177. 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]
  178. 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]
  179. 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]
  180. Yu, T.; Xianyu, Y. Array-Based Biosensors for Bacteria Detection: From the Perspective of Recognition. Small 2021, 17, 2006230. [Google Scholar] [CrossRef]
  181. Arano-Martinez, J.A.; Martínez-González, C.L.; Salazar, M.I.; Torres-Torres, C. A Framework for Biosensors Assisted by Multiphoton Effects and Machine Learning. Biosensors 2022, 12, 710. [Google Scholar] [CrossRef]
  182. Ma, J.; Guan, Y.; Xing, F.; Eltzov, E.; Wang, Y.; Li, X.; Tai, B. Accurate and non-destructive monitoring of mold contamination in foodstuffs based on whole-cell biosensor array coupling with machine-learning prediction models. J. Hazard. Mater. 2023, 449, 131030. [Google Scholar] [CrossRef]
  183. Cui, J.; Han, X.; Shi, G.; Li, K.; Yuan, W.; Zhou, W.; Li, Z.; Wang, M. Hierarchical structure SERS biosensor: A machine learning-driven ultra-sensitive platform for trace detection of amygdalin. Opt. Mater. 2023, 143, 114170. [Google Scholar] [CrossRef]
Figure 1. The roadmap of odor biosensors. The corresponding papers are (a) Moth antenna for pheromone detection [39]. (b) Infect Sf9 cell line with baculovirus harbouring the OR DNA [40]. (c) Measure ethanol concentration with enzyme system [41]. (d) Antibody for odorant detection [42]. (e) Connect multiple oocytes in a fluidic system [43]. (f) Record extracellular potentials of olfactory epithelium [44]. (g) Cell complex expressing OR for direct gas-phase odorant detection [45]. (h) Trap OSNs in a microchamber array [46]. (i) Select suitable ORs for target odorants [47]. (j) Immobilize aptamer on FET for gas-phase odorant detection [48]. (k) Combine moth antenna with drone for odor source localization [49]. (l) Use OBP as a molecular transporter for gas-phase odorant detection [50]. (m) Extend lifetime of gas-phase odor biosensor using liquid thickness control and liquid exchange [51]. (n) Actively track temporally changing gas-phase odor mixture [52].
Figure 1. The roadmap of odor biosensors. The corresponding papers are (a) Moth antenna for pheromone detection [39]. (b) Infect Sf9 cell line with baculovirus harbouring the OR DNA [40]. (c) Measure ethanol concentration with enzyme system [41]. (d) Antibody for odorant detection [42]. (e) Connect multiple oocytes in a fluidic system [43]. (f) Record extracellular potentials of olfactory epithelium [44]. (g) Cell complex expressing OR for direct gas-phase odorant detection [45]. (h) Trap OSNs in a microchamber array [46]. (i) Select suitable ORs for target odorants [47]. (j) Immobilize aptamer on FET for gas-phase odorant detection [48]. (k) Combine moth antenna with drone for odor source localization [49]. (l) Use OBP as a molecular transporter for gas-phase odorant detection [50]. (m) Extend lifetime of gas-phase odor biosensor using liquid thickness control and liquid exchange [51]. (n) Actively track temporally changing gas-phase odor mixture [52].
Biosensors 13 01000 g001
Figure 2. Illustration of insect and vertebrate olfaction. (a) Schematic diagram of insect olfactory sensillum. (b) Signal transduction procedure of insect olfaction. (c) Schematic diagram of vertebrate olfactory system. (d) Signal transduction procedure of vertebrate olfaction.
Figure 2. Illustration of insect and vertebrate olfaction. (a) Schematic diagram of insect olfactory sensillum. (b) Signal transduction procedure of insect olfaction. (c) Schematic diagram of vertebrate olfactory system. (d) Signal transduction procedure of vertebrate olfaction.
Biosensors 13 01000 g002
Figure 3. Graphical abstract of introduced odor biosensors in this manuscript.
Figure 3. Graphical abstract of introduced odor biosensors in this manuscript.
Biosensors 13 01000 g003
Figure 5. Methods to improve the specificity of organ/tissue-based odor biosensors. (a) Schematic drawing of single antennal sensillum measurement. (b) M72 OSN-related axons and glomeruli were marked using the green fluorescence. The electrophysiological signal was recorded using an implantable MEA probe. (a) Adapted with permission from Ref. [39]. Copyright 1978 Elsevier; (b) adapted with permission from Ref. [78]. Copyright 2018 Elsevier.
Figure 5. Methods to improve the specificity of organ/tissue-based odor biosensors. (a) Schematic drawing of single antennal sensillum measurement. (b) M72 OSN-related axons and glomeruli were marked using the green fluorescence. The electrophysiological signal was recorded using an implantable MEA probe. (a) Adapted with permission from Ref. [39]. Copyright 1978 Elsevier; (b) adapted with permission from Ref. [78]. Copyright 2018 Elsevier.
Biosensors 13 01000 g005
Figure 6. An OSN-based odor biosensor. (a) A dissociated OSN from a salamander showing the cell body and several cilia. (b) Several OSNs or other cells dissociated from salamander olfactory epithelium plated onto the surface of the chip. (c) Response of one electrode channel to the odors. (d) The procedure of trapping the OSNs to the microchamber array; the diameter of each well was 10 μm. (e) The fluorescent intensity of two Fluo-4 AM labeled OSNs (ID: HI 28-03, HI 25-18). (f) The fluorescent images of OSN (ID HI28-03) at each step (I, II, III, IV, and V) of (e). (ac) Reprinted with permission from Ref. [83]. Copyright 2016 Elsevier; (df) reprinted from Ref. [46].
Figure 6. An OSN-based odor biosensor. (a) A dissociated OSN from a salamander showing the cell body and several cilia. (b) Several OSNs or other cells dissociated from salamander olfactory epithelium plated onto the surface of the chip. (c) Response of one electrode channel to the odors. (d) The procedure of trapping the OSNs to the microchamber array; the diameter of each well was 10 μm. (e) The fluorescent intensity of two Fluo-4 AM labeled OSNs (ID: HI 28-03, HI 25-18). (f) The fluorescent images of OSN (ID HI28-03) at each step (I, II, III, IV, and V) of (e). (ac) Reprinted with permission from Ref. [83]. Copyright 2016 Elsevier; (df) reprinted from Ref. [46].
Biosensors 13 01000 g006
Figure 7. Methods to detect gas-phase odorants. (a) The odorant solution was added into the space between the wells of the 96-well plate; the evaporated odorant slowly dissolved into the buffer medium. (b) Several spheroids formed by cells were loaded onto the surface of a hydrogel microchamber, and a thin liquid layer covered the spheroids. (c) Cells fixed in the collagen pillars, and the buffer medium from the collagen pedestal can prevent the dry-out problem. (d) Cells were cultured on the polycarbonate membrane and then assembled into the device; the cell side of the membrane was in contact with the culture medium. PDMS: polydimethylsiloxane. (a) Reprinted with permission from Ref. [95]. Copyright 2019 Journal of Visualized Experiments; (b) adapted with permission from Ref. [45]. Copyright 2014 John Wiley and Sons; (d) reprinted with permission from Ref. [98]. Copyright 2015 Elsevier.
Figure 7. Methods to detect gas-phase odorants. (a) The odorant solution was added into the space between the wells of the 96-well plate; the evaporated odorant slowly dissolved into the buffer medium. (b) Several spheroids formed by cells were loaded onto the surface of a hydrogel microchamber, and a thin liquid layer covered the spheroids. (c) Cells fixed in the collagen pillars, and the buffer medium from the collagen pedestal can prevent the dry-out problem. (d) Cells were cultured on the polycarbonate membrane and then assembled into the device; the cell side of the membrane was in contact with the culture medium. PDMS: polydimethylsiloxane. (a) Reprinted with permission from Ref. [95]. Copyright 2019 Journal of Visualized Experiments; (b) adapted with permission from Ref. [45]. Copyright 2014 John Wiley and Sons; (d) reprinted with permission from Ref. [98]. Copyright 2015 Elsevier.
Biosensors 13 01000 g007
Figure 8. Methods to increase the OR types in odor biosensors. (a) A microfluidic microwell array for trapping the cells; the enlarged figure was an OSN in a microwell. (b) Multiple Xenopus oocyte cells were arranged in a fluidic system. (c) The same type of cells was immobilized in the same area. (d) Cells were randomly placed, and the cell type was labeled after two single-component odorant stimulations. (a) Reprinted with permission from Ref. [82]. Copyright 2010 The Royal Society of Chemistry. (b) Reprinted from Ref. [43]. (d) Reprinted from Ref. [108].
Figure 8. Methods to increase the OR types in odor biosensors. (a) A microfluidic microwell array for trapping the cells; the enlarged figure was an OSN in a microwell. (b) Multiple Xenopus oocyte cells were arranged in a fluidic system. (c) The same type of cells was immobilized in the same area. (d) Cells were randomly placed, and the cell type was labeled after two single-component odorant stimulations. (a) Reprinted with permission from Ref. [82]. Copyright 2010 The Royal Society of Chemistry. (b) Reprinted from Ref. [43]. (d) Reprinted from Ref. [108].
Biosensors 13 01000 g008
Figure 9. The format of OR protein used in odor biosensors. (a) The original format of human OR protein was immobilized on graphene micro patterns (GMs). (b) OR protein was inserted into a nanodisc and then immobilized on the carbon nanotube. (c) OR and Orco were embedded into a bilayer lipid membrane. (a) Reprinted with permission from Ref. [113]. Copyright 2015 ACS Publications. (b) Reprinted with permission from Ref. [119]. Copyright 2017 ACS Publications. (c) Reprinted with permission from Ref. [124]. Copyright 2019 ACS Publications.
Figure 9. The format of OR protein used in odor biosensors. (a) The original format of human OR protein was immobilized on graphene micro patterns (GMs). (b) OR protein was inserted into a nanodisc and then immobilized on the carbon nanotube. (c) OR and Orco were embedded into a bilayer lipid membrane. (a) Reprinted with permission from Ref. [113]. Copyright 2015 ACS Publications. (b) Reprinted with permission from Ref. [119]. Copyright 2017 ACS Publications. (c) Reprinted with permission from Ref. [124]. Copyright 2019 ACS Publications.
Biosensors 13 01000 g009
Figure 10. The examples of FET (a), interdigitated microelectrode array (b), SAW (c), QCM (d), SPR (e), EIS (f), and planar electrode pair (g) used in OR protein-based odor biosensors. (a) Adapted from Ref. [120]. (b) Adapted with permission from Ref. [111]. Copyright 2012 Elsevier. (c) Reprinted with permission from Ref. [104]. Copyright 2011 Elsevier. (d) Adapted with permission from Ref. [112]. Copyright 2013 Elsevier. (e) Reprinted from Ref. [128]. (f) Reprinted with permission from Ref. [127]. Copyright 2019 Elsevier. (g) Adapted from Ref. [125].
Figure 10. The examples of FET (a), interdigitated microelectrode array (b), SAW (c), QCM (d), SPR (e), EIS (f), and planar electrode pair (g) used in OR protein-based odor biosensors. (a) Adapted from Ref. [120]. (b) Adapted with permission from Ref. [111]. Copyright 2012 Elsevier. (c) Reprinted with permission from Ref. [104]. Copyright 2011 Elsevier. (d) Adapted with permission from Ref. [112]. Copyright 2013 Elsevier. (e) Reprinted from Ref. [128]. (f) Reprinted with permission from Ref. [127]. Copyright 2019 Elsevier. (g) Adapted from Ref. [125].
Biosensors 13 01000 g010
Figure 11. OBP works as a molecular transporter to enhance gas-phase odorant detection ability. Adapted with permission from Ref. [50]. Copyright 2022 ACS Publications.
Figure 11. OBP works as a molecular transporter to enhance gas-phase odorant detection ability. Adapted with permission from Ref. [50]. Copyright 2022 ACS Publications.
Biosensors 13 01000 g011
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.; Nakamoto, T. Biosensors for Odor Detection: A Review. Biosensors 2023, 13, 1000. https://doi.org/10.3390/bios13121000

AMA Style

Deng H, Nakamoto T. Biosensors for Odor Detection: A Review. Biosensors. 2023; 13(12):1000. https://doi.org/10.3390/bios13121000

Chicago/Turabian Style

Deng, Hongchao, and Takamichi Nakamoto. 2023. "Biosensors for Odor Detection: A Review" Biosensors 13, no. 12: 1000. https://doi.org/10.3390/bios13121000

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