Next Article in Journal
Early Detection of Cardiovascular Disease Using Laser-Induced Breakdown Spectroscopy Combined with Machine Learning
Previous Article in Journal
A Checkerboard Corner Detection Method for Infrared Thermal Camera Calibration Based on Physics-Informed Neural Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exhaled Breath Analysis (EBA): A Comprehensive Review of Non-Invasive Diagnostic Techniques for Disease Detection

1
Faculty of Electrical and Computer Engineering, University of Tabriz, Tabriz 51664, Iran
2
College of Engineering, Design and Physical Sciences, Brunel University London, Uxbridge UB8 3PH, UK
*
Authors to whom correspondence should be addressed.
Photonics 2025, 12(9), 848; https://doi.org/10.3390/photonics12090848
Submission received: 10 July 2025 / Revised: 21 August 2025 / Accepted: 22 August 2025 / Published: 25 August 2025

Abstract

Exhaled breath analysis (EBA) is an advanced, non-invasive diagnostic technique that utilizes volatile organic compounds (VOCs) to detect and monitor various diseases. This review examines EBA’s historical development and current status as a promising diagnostic tool. It highlights the significant contributions of modern methods such as gas chromatography–mass spectrometry (GC-MS), ion mobility spectrometry (IMS), and electronic noses in enhancing the sensitivity and specificity of EBA. Furthermore, it emphasizes the transformative role of nanotechnology and machine learning in improving the diagnostic accuracy of EBA. Despite challenges such as standardization and environmental factors, which must be addressed for the widespread adoption of this technique, EBA shows excellent potential for early disease detection and personalized medicine. The review also highlights the potential of photonic crystal fiber (PCF) sensors, known for their superior sensitivity, in the field of EBA.

1. Introduction

Exhaled breath analysis (EBA) has emerged as a groundbreaking non-invasive diagnostic technique, offering notable advantages over conventional methods such as blood tests and biopsies [1,2]. This technique leverages the analysis of volatile organic compounds (VOCs) present in exhaled breath, which allows the identification of biomarkers associated with various diseases, thereby facilitating early diagnosis and monitoring of disease progression [3]. Although volatile organic compounds (VOCs) are central to modern breath analysis, the diagnostic potential of exhaled breath extends beyond this single class of molecules. The intricate biomolecular composition of exhaled breath includes a wide range of biomarkers, such as inorganic gases and non-volatile substances, which provide a comprehensive overview of an individual’s metabolic state. For instance, endogenous inorganic gases serve as critical indicators: nitric oxide (NO) is an established biomarker for airway inflammation in asthma; ammonia (NH3) has been linked to renal and liver disorders; hydrogen (H2) and methane (CH4) are routinely measured to diagnose gastrointestinal conditions; and carbon dioxide (CO2) dynamics can also be indicative of pulmonary diseases. Additionally, analyzing non-volatile compounds—such as proteins, cytokines, and lipids—captured in exhaled breath condensate (EBC) offers a distinct yet complementary diagnostic window for pulmonary and systemic inflammation [4].
The historical origins of breath analysis can be traced back to ancient Egypt, where physicians diagnosed diseases based on breath odors. They associated specific odors with certain conditions, such as a sweet scent with diabetes mellitus, a fishy odor with liver disease, and a urine-like smell with kidney disease [5]. While using breath as a diagnostic tool is not a novel concept, its evolution into a sophisticated, scientifically driven practice is more recent. In contemporary times, the adoption of advanced analytical techniques, such as gas chromatography–mass spectrometry (GC-MS), ion mobility spectrometry (IMS), and electronic noses, has greatly enhanced the accuracy and reliability of EBA [6]. These technologies enable the detection of VOCs at extremely low concentrations, which is critical for detecting subtle metabolic changes that may signal disease.
The journey of exhaled breath research began with pioneers such as Linus Pauling, who employed gas chromatography to demonstrate the presence of numerous volatile compounds in breath samples [7]. This seminal work paved the way for contemporary research linking specific VOCs to various health conditions. Today, nanotechnology and sensor design advancements have significantly improved the sensitivity and specificity of breath analysis, making it a powerful tool in modern medicine [8].
One of the main advantages of EBA is its non-invasive nature. In contrast to blood tests and biopsies, which require the extraction of bodily fluids or tissues, breath analysis can be performed quickly and painlessly, making it more acceptable to patients [9]. The non-invasive nature of EBA is also advantageous for monitoring chronic diseases and managing critically ill patients, as it reduces infection risks and enables frequent testing [10,11,12]. In recent years, EBA has demonstrated promise in diagnosing a wide range of diseases, including respiratory conditions such as asthma and chronic obstructive pulmonary disease (COPD), metabolic disorders like diabetes, and even cancer [13]. EBA has shown potential for early and non-invasive cancer detection by analyzing VOC patterns in breath, particularly in lung cancer [14] and head and neck cancers [15]. Similarly, breath tests have been developed to monitor glucose levels in patients with diabetes, offering a convenient alternative to traditional blood glucose testing [16].
The underlying principle of EBA is that metabolic processes produce specific VOCs exhaled in breath. These VOCs can serve as biomarkers for various physiological and pathological states. For example, isoprene, acetone, and ethane are VOCs associated with oxidative stress and lipid peroxidation, which are common in many diseases [17]. The systemic pathway by which VOCs originate from tissues and are transported via the bloodstream before being expelled in breath is illustrated in Figure 1. By analyzing the concentrations and patterns of these compounds, insights can be gained into a person’s health status.
Technological innovations have been crucial in advancing exhaled breath analysis (EBA), improving its sensitivity, accuracy, and potential for clinical applications. Integrating machine learning algorithms with analytical techniques has enhanced the ability to interpret complex breath data. Machine learning models can analyze large datasets to uncover hidden patterns and correlations that traditional statistical techniques might overlook, enhancing diagnostic accuracy. This has led to the development of predictive models for disease detection and progression, increasing the clinical utility of EBA. Machine learning algorithms can identify subtle changes in VOC patterns that may indicate disease, further improving the accuracy and reliability of EBA. Despite its promise, EBA still faces several challenges that must be addressed. One of the main issues is the standardization of breath sampling and analysis procedures. Variability in sampling methods, environmental factors, and individual differences can affect the reproducibility and accuracy of results [18].
Figure 1. This figure illustrates the systemic dispersal of volatile organic compounds (VOCs) originating from gastrointestinal tumor tissues to various regions within the body [19].
Figure 1. This figure illustrates the systemic dispersal of volatile organic compounds (VOCs) originating from gastrointestinal tumor tissues to various regions within the body [19].
Photonics 12 00848 g001
Numerous reviews have been published on EBA, covering various aspects from biomarker discovery to specific sensor types. However, this review provides a unique added value by bridging the gap between established analytical techniques (like GC-MS and electronic noses) and the cutting-edge developments in photonic crystal fiber (PCF) sensor technology. While other reviews may mention PCF sensors in passing, this work focuses on them, providing a detailed overview of their operational principles and recent applications. By systematically comparing the advantages and disadvantages of PCF sensors with other prominent technologies, this review offers a forward-looking perspective on how photonics is poised to overcome key challenges in the field, paving the way for the next generation of non-invasive diagnostic tools.
In conclusion, exhaled breath analysis represents a promising non-invasive diagnostic tool with significant potential to revolutionize healthcare. Its ability to detect VOCs associated with various diseases offers a valuable means for early diagnosis and monitoring of disease progression. However, further research and technological advancements are necessary to overcome the existing challenges and fully realize the potential of this innovative diagnostic approach. EBA techniques’ continued development and validation promise to improve patient outcomes and advance personalized medicine. The potential of EBA in customized medicine is vast. By analyzing an individual’s unique VOC profile, EBA can provide tailored diagnostic and treatment strategies, leading to more effective and efficient healthcare. Additionally, developing portable and user-friendly devices for breath analysis is crucial for the widespread adoption of EBA in clinical practice.

2. Methodology: A Systematic Approach to Literature Synthesis

The foundation of this comprehensive review rests upon a rigorous and systematic interrogation of the scientific literature, designed to capture the breadth and depth of research in exhaled breath analysis (EBA). Our search strategy was executed across a curated selection of premier academic databases, including the interdisciplinary archives of Web of Science and Scopus, the biomedical-centric repository of PubMed, and the engineering and technology-focused collections within IEEE Xplore and SpringerLink. Google Scholar was used as a supplementary tool to identify additional relevant publications and gray literature. The search, encompassing literature published up to early 2025, was not static but an iterative process. It employed a multi-tiered query strategy combining MeSH (Medical Subject Headings) terms with a spectrum of keywords, evolving from broad concepts to specific technologies: (“exhaled breath analysis” OR “breathomics”) AND (“volatile organic compounds” OR “VOCs” OR “breath biomarkers”) AND (“gas sensor” OR “electronic nose” OR “GC-MS” OR “non-invasive diagnosis”) AND (“photonic crystal fiber” OR “PCF”).
A stringent set of inclusion and exclusion criteria was applied to sculpt a focused and relevant narrative from the vast body of resulting literature. We prioritized peer-reviewed original research articles and authoritative reviews that form the intellectual bedrock of the field. The scope was intentionally confined to studies published in English, centering on human breath analysis for disease diagnosis or biomarker discovery and detailing the principles or applications of sensor technologies in EBA. Conversely, we systematically excluded non-human (animal) studies, ephemeral communications such as editorials or conference abstracts lacking a full paper, and articles where breath analysis was not the primary diagnostic modality, ensuring the integrity and focus of our review.
This multi-stage selection process was a meticulous exercise in intellectual distillation. The initial pool of several thousand articles was screened by title and abstract to filter for relevance. The remaining publications underwent a full-text assessment against the established criteria. This was not merely a filtering task but a synthesis process, where we sought to weave disparate research threads into a coherent and insightful tapestry. The ultimate goal was to construct a review that not only summarizes the state of the art but also illuminates the trajectory of the field, highlighting the critical transition from established methods to the next frontier of photonic-based diagnostics.

3. Common Diseases Diagnosed by Breath Analysis

Breath analysis is emerging as a critical, non-invasive diagnostic method with significant potential for the early identification and monitoring of various illnesses. This approach involves analyzing exhaled breath for volatile organic compounds (VOCs) and other markers indicative of specific diseases. The non-invasive nature of breath analysis enhances its appeal, as it reduces patient discomfort and the need for more invasive procedures, such as biopsies or blood tests. Furthermore, its ability to provide immediate results makes it a highly effective tool in clinical settings.
This section explores the application of breath analysis in diagnosing several prevalent diseases by examining their specific biomarkers in exhaled breath. A comprehensive discussion of these diseases follows. Table 1 details the biomarkers identified for each disease, explains their diagnostic use, cites relevant studies, and lists any FDA-approved markers. To complement this, Figure 2 offers a visual summary of prominent diseases and their associated breath biomarkers.
Table 1. Diseases diagnosed by breath analysis.
Table 1. Diseases diagnosed by breath analysis.
Name of the DiseaseExhaled BiomarkerDescriptionReferences of StudiesFDA-Approved Biomarkers for Disease Diagnosis
Lung CancerVOCs such as alkanes, benzene derivatives, aldehydesBreath analysis for lung cancer involves detecting specific VOCs produced by cancerous cells. Elevated levels of n-pentane and isoprene are commonly observed.[20,21]None
AsthmaNitric oxide (NO)Elevated levels of NO in exhaled breath indicate airway inflammation, which is a hallmark of asthma.[22,23]Fractional exhaled nitric oxide (FeNO)
Chronic Obstructive Pulmonary Disease (COPD)Carbon dioxide (CO2), methane (CH4), ethaneCOPD diagnosis through breath analysis involves detecting elevated levels of specific gases such as ethane and pentane.[24,25]None
Diabetes MellitusAcetoneElevated levels of acetone in breath correlate with blood glucose levels, indicating diabetes.[1,26]None
Helicobacter pylori InfectionCarbon-13 (13C) ureaThe urea breath test (UBT) detects H. pylori infection by measuring labeled CO2 after ingestion of 13C-urea.[24,27]13C-urea
Liver DiseaseAmmonia, acetoneElevated levels of ammonia and acetone indicate impaired liver function.[28,29]None
Neonatal JaundiceCarbon monoxide (CO)CO breath test detects elevated levels of CO, indicating jaundice in newborns.[30,31]Carbon monoxide test
Gastrointestinal DisordersHydrogen (H2), methane (CH4)Breath tests for hydrogen and methane are used to diagnose conditions like fructose and lactose malabsorption and bacterial overgrowth.[32,33]Hydrogen and methane breath tests
Alcohol ConsumptionEthanolEthanol breath test measures blood alcohol levels.[34]Ethanol test

4. Sensors Used in Exhaled Breath Analysis

Exhaled breath analysis is a promising non-invasive diagnostic tool that detects volatile organic compounds (VOCs) to identify various diseases. The development and application of sensors in this field have significantly advanced our ability to diagnose and monitor health conditions through breath analysis. Various types of sensors, including metal oxide gas, chemiresistive, and mid-infrared tunable laser spectroscopic, are crucial in detecting biomarkers in exhaled breath. The following section will introduce the sensors in this field for detecting gases. An overview of the various sensor technologies employed in EBA is presented in Figure 3. The general principle involves the interaction of exhaled biomarkers with a sensor, which generates a detectable signal that is subsequently amplified and analyzed, as shown schematically in Figure 4.

4.1. Role of Nanomaterials in Sensor Development

Nanomaterials, such as carbon nanotubes, graphene, and metal oxides, provide unique properties that enhance sensor performance. Their high surface area/volume ratio increases the active sites available for gas interaction, improving sensitivity. For example, single-walled carbon nanotubes (SWNTs) functionalized with specific chemical groups have shown high sensitivity to nitric oxide (NO), a significant marker in breath analysis for respiratory diseases [35]. This high sensitivity to NO is particularly crucial for clinical applications, as it allows for the non-invasive monitoring of asthma control and patient response to anti-inflammatory therapy.
Graphene and its derivatives are also notable for their exceptional electrical properties, which enhance the response of gas sensors. Graphene-based sensors can detect minute changes in electrical conductivity when VOCs are adsorbed on their surface. This property is beneficial for detecting acetone, a biomarker for diabetes [36]. The ability to detect acetone with such precision is a significant step towards developing a convenient, pain-free daily monitoring tool for diabetic patients, replacing the need for frequent blood tests.
The incorporation of nanomaterials not only improves sensitivity but also reduces sensors’ response and recovery times. Metal oxide nanomaterials, such as ZnO and SnO2, exhibit rapid response times when exposed to VOCs due to their excellent electron mobility and surface reaction kinetics. These materials have shown significant improvements in detecting low concentrations of gases, such as isopropanol, indicative of lung cancer [37].
Additionally, the stability of nanomaterial-based sensors is a critical factor for their practical application. ZnO-Bi2O3 nanosheets grown on hollow-core fibers have demonstrated exceptional stability and repeatability in detecting acetone at room temperature, making them suitable for long-term breath analysis [38]. Such exceptional stability at room temperature addresses a key challenge for practical sensors—long-term reliability—making these devices more suitable for real-world clinical and point-of-care settings.
Nanomaterials offer tunable chemical and physical properties, allowing for the customization of sensors to target specific VOCs. This versatility is crucial for developing selective sensors that distinguish between different gas analytes. For instance, ZnO nanostructures can be doped with various metals to enhance their selectivity for particular VOCs. Fe-doped ZnO nanoneedles have been shown to selectively detect isopropanol, a potential biomarker for lung diseases, at very low concentrations [37]. The chemical versatility of carbon-based nanomaterials also makes them ideal for constructing flexible and wearable sensors. This property is particularly advantageous for developing portable diagnostic devices that can provide real-time monitoring of VOCs in exhaled breath [39]. The development of advanced fabrication techniques has enabled the precise integration of nanomaterials into sensor platforms, enhancing their performance. For example, using horizontal vapor-phase crystal growth to synthesize ZnO and SnO2 nanomaterials has resulted in sensors with rapid response times and high sensitivity to VOCs [40]. Moreover, microfabrication and microfluidic technologies have facilitated the miniaturization of gas sensors, making them more compact and portable. This advancement is essential for creating point-of-care devices that can be used for early disease detection in clinical settings. Several studies have demonstrated the practical application of nanomaterial-based sensors in detecting disease biomarkers. For instance, a smartphone-based resistive gas sensor employing ZnO nanosheets in exhaled breath has shown high sensitivity to lung cancer-related VOCs, such as diethyl ketone and acetone. This innovative approach combines the advantages of nanomaterials with the accessibility of smartphone technology, providing a cost-effective solution for early disease diagnosis [41]. Another notable application is the use of polymer-modified quartz tuning forks embedded with nanomaterials to detect low concentrations of VOCs. This method has proven effective in distinguishing between healthy and VOC-spiked breath samples, showcasing its potential as a non-invasive diagnostic tool [42]. While nanomaterial-based sensors have shown significant promise, challenges remain to be overcome. One major issue is the selectivity of sensors in the presence of multiple interfering gases. Researchers are exploring strategies to improve selectivity, such as functionalizing nanomaterials with specific chemical groups or combining different nanomaterials to create composite sensors [43].
Additionally, incorporating advanced signal processing methods and machine learning algorithms can improve the precision and dependability of breath analysis results. These technologies can help in the real-time analysis of complex VOC profiles, providing more precise diagnostic information [44]. Incorporating nanomaterials into sensor technology has revolutionized the field of exhaled breath analysis. Their unique properties enhance sensors’ sensitivity, selectivity, and overall performance, making them invaluable tools for non-invasive disease diagnosis. Recent advancements in nanomaterial-based sensors, summarizing the applications and target gases discussed in this section, are presented in Table 2.

4.2. Chemiresistive Gas Sensors

Chemiresistive gas sensors are crucial for medical diagnostics and environmental monitoring to detect exhaled gases. The principle of chemiresistive gas sensors involves a change in electrical resistance when exposed to target gas molecules. The sensor, typically composed of a sensitive material like metal oxides, interacts with gas molecules in the breath. These molecules adsorb onto the surface, altering charge carrier density and leading to a measurable change in resistance correlated to gas concentration, enabling the detection and quantification of various biomarkers in exhaled breath [53,54]. This technology is widely used for gas detection due to its simplicity, low cost, and high sensitivity, playing a significant role in monitoring exhaled gases for health condition biomarkers [55]. Gas molecules either donate or withdraw electrons from the sensor material, affecting its resistance. A typical sensor has a sensing layer on an insulating substrate with electrodes for measuring resistance changes [56]. This layer can be optimized for sensitivity and selectivity to specific gases. Exhaled breath contains VOC biomarkers detectable by these sensors, enabling non-invasive diagnostics [57]. Humidity and other gases in exhaled breath can interfere with sensor readings, leading to false positives or inaccurate measurements. Research focuses on integrating humidity-resistant materials and advanced sensor designs to mitigate this issue [58]. Over time, chemiresistive sensors may experience drift in their baseline resistance, affecting their long-term stability and reliability. Efforts are being made to improve the material properties and sensor designs to enhance stability and reduce drift [59].
Accurately quantifying biomarker concentrations in exhaled breath is challenging due to the low concentrations and the presence of interfering substances. Developing highly sensitive and selective sensors is crucial for accurate quantification [60]. Current sensors may not be capable of simultaneously detecting a broad spectrum of biomarkers. This limitation is being addressed by developing sensor arrays and multifunctional sensors that can detect multiple biomarkers simultaneously [61]. Sensors based on nanoparticle-structured interfaces have been designed to detect lung cancer biomarkers in human breath with high sensitivity and a limit of detection as low as six ppb [62]. An ultrasensitive chemiresistive sensor based on γ-Bi2MoO6-CuO heterostructure can detect H2S, a biomarker for asthma, with a detection limit of 5 ppb. This sensor helps distinguish between asthmatic patients and healthy individuals and can monitor the severity of asthma [63].

4.3. Electronic Nose Technology

Electronic nose (eNose) technology, modeled after the mammalian sense of smell, has become a promising non-invasive tool for disease diagnosis by detecting volatile organic compounds (VOCs) in breath [64]. VOCs are organic chemicals with high vapor pressure at room temperature, and their analysis in exhaled breath can indicate the presence of different diseases such as cancer, lung conditions, and infections [65,66]. An eNose system typically comprises three main components: a sensor array, a signal processing circuit, and a pattern recognition system [64]. The sensor array is the heart of the eNose, consisting of multiple chemical sensors with varying sensitivity and selectivity to different VOCs [67]. These sensors can be categorized based on their transduction mechanisms, including:
  • Metal Oxide Sensors (MOSs): These sensors rely on the change in electrical conductivity of a metal oxide semiconductor upon interaction with VOCs. When VOC molecules adsorb onto the sensor surface, they change the electron density, resulting in a detectable variation in resistance [67]. MOSs are among the most widely used sensors in eNose technology due to their high sensitivity, fast response times, and relatively low cost [68]. These sensors are based on a semiconducting metal oxide layer (e.g., tin dioxide and zinc oxide) deposited on a substrate. The gas molecules adsorb onto the sensor surface upon exposure to VOCs, changing the metal oxide’s electrical conductivity. This change in conductivity is directly proportional to the concentration of the VOCs and can be measured as a change in resistance. The selectivity of MOSs can be tuned by adjusting the operating temperature, the type of metal oxide, and the addition of dopants or catalysts [69].
  • Conducting Polymer (CP) Sensors: Similar to MOSs, CP sensors also exhibit changes in electrical conductivity upon exposure to VOCs. The interaction of VOCs with the polymer matrix causes swelling or contraction, resulting in a change in resistance. CP sensors offer an alternative approach to VOC detection, leveraging changes in the electrical conductivity of a polymer film upon interaction with VOCs. These sensors often comprise a polymer matrix (e.g., polypyrrole and polyaniline) embedded with conductive particles (e.g., carbon black). The adsorption of VOCs onto the polymer matrix can cause it to swell or contract, resulting in a change in the distance between the conductive particles and, consequently, a change in resistance. CP sensors are known for their flexibility in design, ease of fabrication, and potential for miniaturization [70].
  • Quartz Crystal Microbalance (QCM) Sensors: These sensors utilize the piezoelectric effect of a quartz crystal resonator. When VOCs adsorb onto the crystal surface, the mass of the crystal changes, leading to a shift in its resonant frequency. This frequency shift is proportional to the mass of the adsorbed VOCs [71]. QCM sensors exploit the piezoelectric properties of quartz crystals to detect VOCs. These sensors consist of a quartz crystal resonator coated with a selective material that adsorbs specific VOCs. The adsorption of VOCs onto the crystal surface increases its mass, causing a decrease in the crystal’s resonant frequency. This frequency shift is proportional to the mass of the adsorbed VOCs and can be used to quantify their concentration. QCM sensors are highly sensitive and can detect even trace amounts of VOCs [72].
  • Mass Spectrometry (MS) Sensors: MS sensors ionize VOC molecules and separate them according to their mass-to-charge ratio, producing a mass spectrum that serves as a unique fingerprint for identifying the VOCs in the sample [73]. While not as common as MOS, CP sensors, or QCM sensors, MS sensors offer unparalleled selectivity and sensitivity in VOC detection. In MS sensors, VOCs are ionized and separated based on their mass-to-charge ratio. The resulting mass spectrum provides a unique fingerprint of the VOCs present in the sample. However, MS sensors are typically bulky, expensive, and require complex operation, limiting their widespread adoption in eNose technology [74]. A comparative summary of these sensor technologies, highlighting their principles of operation, advantages, and disadvantages, is provided in Table 3.
The signal processing circuit collects the signals generated by the sensor array, amplifies them, filters out noise, and converts them into digital data. This data is then fed into the pattern recognition system, which employs machine learning algorithms to identify patterns and correlations within the sensor responses. By comparing these patterns to a database of known VOC profiles, the eNose can identify and quantify the VOCs in the exhaled breath, thus aiding in disease diagnosis. Despite its potential, eNose technology faces several challenges, including sensor limitations in sensitivity and selectivity, interference from environmental factors such as humidity and temperature, and the complexity of data analysis [75,76]. However, ongoing research and development efforts are focused on overcoming these challenges by developing more sensitive and selective sensors, robust algorithms for data analysis, and portable and affordable eNose devices [77]. With continued advancements, eNose technology is promising to revolutionize disease diagnosis and personalized medicine.

4.4. Chromatography

Chromatography operates on the principle of differential partitioning, where analytes distribute themselves between a stationary phase and a mobile phase based on their affinity for each phase. The stationary phase can be solid, liquid, or gel, while the mobile phase is typically a gas (in GC) or liquid (in LC). The analytes’ unique interactions with both phases result in their separation as they travel through the chromatographic system. This principle is fundamental to all chromatographic techniques used in breath gas analysis [78].
  • Gas Chromatography (GC): GC is the workhorse of breath gas analysis, separating volatile organic compounds (VOCs) based on their boiling points and polarity. The sample is vaporized and carried through a column by an inert carrier gas. Different VOCs interact differently with the column’s stationary phase, leading to separation. This method is frequently combined with mass spectrometry (GC-MS) and offers excellent sensitivity and specificity for detecting and measuring VOCs [79].
  • Liquid Chromatography (LC): LC is used for less volatile or polar analytes, such as metabolites. In this process, the sample is dissolved in a liquid mobile phase and then injected into a column containing a solid stationary phase. Separation occurs due to differences in the analytes’ affinity for the stationary and mobile phases. LC-MS is a powerful combination for analyzing complex biological samples [80].
  • Ion Mobility Spectrometry (IMS): IMS is a rapid technique where ionized gas-phase molecules are separated based on their size, shape, and charge in an electric field. IMS is becoming increasingly popular for its speed, portability, and sensitivity in detecting VOCs, including those in exhaled breath [81].

Challenges and Limitations

  • Gas Chromatography (GC)
Despite its advantages, GC faces several challenges. The requirement for sample pre-treatment and derivatization can be time-consuming and introduce variability. Additionally, the high cost of equipment and maintenance limits its widespread clinical application [81].
  • Liquid Chromatography (LC)
LC also has its limitations, including the complexity of the system and the need for extensive method development to optimize separation conditions. The lower sensitivity compared to GC-MS can be a drawback for detecting low-abundance metabolites [81].
  • Ion Mobility Spectrometry (IMS)
While advantageous for its rapid analysis and portability, IMS suffers from lower resolution compared to traditional chromatographic methods. This may hinder its capacity to differentiate between compounds with comparable structures. Furthermore, environmental factors like humidity and temperature can impact IMS measurements, potentially reducing accuracy [81].

4.5. Infrared Spectroscopy

Mid-infrared spectroscopy, including photoacoustic spectroscopy, is highly sensitive to detecting specific biomarkers in breath samples. It is also non-invasive and provides real-time results [82]. The technique uses mid-infrared laser absorption to identify various gas molecules in exhaled breath, making it highly specific and sensitive [83].

Measurement Process

  • Sample Collection:
Exhaled breath is collected non-invasively using sampling bags or direct breath capture systems. These systems ensure that the breath sample is uncontaminated and accurately represents the gases present in the respiratory tract [84].
2.
Mid-Infrared Laser Source:
A mid-infrared laser, typically a quantum cascade laser (QCL), emits light at specific wavelengths that correspond to the absorption peaks of the target gas molecules. This light is directed through the breath sample [85].
3.
Photoacoustic Effect:
The light absorbed by the gas molecules in the sample causes them to heat up and expand, creating pressure or sound waves. This is known as the photoacoustic effect. The intensity of these sound waves is proportional to the concentration of the gas molecules [86].
4.
Detection:
The generated sound waves are detected using sensitive microphones or acoustic sensors. The signals are then processed and analyzed to determine the concentration of specific gases in the breath sample [87]. A key advantage of mid-infrared spectroscopy is its capability to deliver real-time results, which is essential for applications such as medical diagnostics requiring immediate analysis. Additionally, the non-invasive nature of the method makes it suitable for continuous monitoring without patient discomfort [85]. Despite its advantages, mid-infrared spectroscopy has certain limitations. One major drawback is the requirement for high precision in instrument calibration and maintenance to ensure accurate measurements. Moreover, the initial cost and complexity of the equipment can hinder its widespread adoption in clinical settings [84]. In conclusion, while mid-infrared spectroscopy, including photoacoustic spectroscopy, offers high sensitivity and non-invasive, real-time analysis of breath biomarkers, its high costs and stringent calibration requirements may limit its broader application [82].

4.6. Photonic Crystal Fiber Sensors

Photonic crystal fiber (PCF) sensors direct light through a microstructured optical fiber featuring a periodic pattern of air holes along its length. The key mechanism that enables high sensitivity in PCF sensors is the interaction between the light and the target gas within the fiber’s hollow core. This interaction is enhanced by the large surface area and the specific geometrical configuration of the PCF, which maximizes the overlap between the light field and the gas molecules [88]. The hollow-core design allows the light to be confined in the core while interacting with the gas molecules. This interaction leads to a change in the light’s properties, such as its intensity or wavelength, which can be measured to determine the concentration of the gas [89]. The evanescent field, which extends into the hollow core, enhances the sensitivity by increasing the light-gas interaction length [90]. A typical experimental configuration for performing such gas sensing measurements is illustrated in Figure 5.
Differential Optical Absorption Spectroscopy (DOAS): This technique involves measuring the absorption of specific wavelengths of light that correspond to the absorption lines of the target gas. The hollow-core PCF acts as a gas cell where the light travels through the gas, allowing precise gas concentration measurement based on absorption characteristics [92]. Geometrical Configuration: By optimizing the size, shape, and arrangement of the air holes in the PCF, the sensitivity and detection limits can be significantly enhanced. Different designs, such as spiral porous cores and triangular structures, have been shown to provide high sensitivity [93]. Material Selection: The choice of materials for the core and cladding can also impact the performance. For example, using materials with high refractive indices can improve the confinement of light within the core, enhancing the sensor’s sensitivity.

Outstanding Features of PCF Sensors

  • High Sensitivity: PCF sensors can detect very low concentrations of gases due to their enhanced light-gas interaction. Sensitivities as high as 75% have been reported, making them suitable for detecting trace amounts of gases such as ammonia, hydrogen, and methane [94].
  • Fast Response Time: The design of PCF sensors ensures rapid response to changes in gas concentration, which is critical for real-time monitoring and breath analysis applications. The fast response is due to the efficient gas diffusion within the hollow core and the swift interaction with the guided light [95].
  • Compact and Lightweight: PCF sensors are inherently compact and lightweight, which makes them easy to integrate into portable devices for on-site and real-time gas monitoring. This compactness is advantageous for applications in medical diagnostics, environmental monitoring, and industrial safety [96].
  • Wide Wavelength Range: These sensors can operate over a wide range of wavelengths, from ultraviolet to infrared, allowing the detection of various gases with different absorption characteristics. This versatility makes PCF sensors suitable for multiple applications [97].
  • Robustness and Durability: PCF sensors are engineered for durability and robustness, allowing them to perform reliably even under harsh environmental conditions. Their structural integrity ensures long-term reliability and consistent performance, essential for continuous monitoring applications [98]. A comprehensive comparison of these sensor technologies, outlining their respective advantages, disadvantages, and operational features, is presented in Table 4.
Photonic crystal fiber (PCF) sensors leverage advanced optical and structural properties to provide high sensitivity, fast response times, and robust performance for gas detection. Despite their complex manufacturing requirements and sensitivity to environmental factors, these features make them highly effective for applications such as breath analysis, environmental monitoring, and industrial safety. Photonic crystal fiber sensors are highly effective for detecting gases in exhaled breath because of their exceptional sensitivity, selectivity, stability, and resistance to humidity interference. These advanced qualities make them particularly suitable for non-invasive disease diagnosis via breath analysis.

5. PCF Sensors in EBA

Photonic crystal fiber (PCF) sensors have emerged as a significant advancement in the field of exhaled breath analysis (EBA). These sensors utilize distinctive optical characteristics to identify and measure volatile organic compounds (VOCs) as well as additional biomarkers present in exhaled breath. The core principle of PCF sensors is based on guiding light through a microstructured optical fiber with a periodic arrangement of air holes running along its length. This design allows a high degree of interaction between the light and the target gas molecules, thereby enhancing the sensitivity and selectivity of the sensors [104].
PCF sensors operate either by confining light within a hollow core or by utilizing the evanescent field effect, in which the light interacts with the gas molecules in the air holes [105]. The alteration of the light’s properties—such as intensity or wavelength—when interacting with gas molecules is measured to quantify the concentration of the target analytes [106]. The hollow-core PCF design, in particular, maximizes the overlap between the light field and the gas molecules, thereby significantly enhancing the sensor’s performance [107].

Types of Photonic Crystal Fiber Sensors

  • Hollow-Core PCF Sensors: These sensors confine the light within a hollow core surrounded by a microstructured cladding. The interaction between light and gas molecules takes place within the hollow core, resulting in enhanced sensitivity and swift response times [108].
  • Solid-Core PCF Sensors: These sensors use the evanescent field that extends into the microstructured cladding filled with gas. Although they generally have lower sensitivity than hollow-core PCFs, they are more straightforward to fabricate and highly effective for specific applications [109].
PCF sensors have shown great promise in detecting various biomarkers in exhaled breath, such as nitric oxide (NO), ammonia (NH3), acetone, and other VOCs associated with diseases like asthma, diabetes, and lung cancer [104]. PCF sensors’ high sensitivity and specificity make them suitable for early diagnosis and monitoring of these conditions [110]. For instance, ammonia detection using PCFs is crucial for early diagnosis of renal diseases, as elevated ammonia levels in breath indicate kidney dysfunction [104]. Additionally, detecting biomarkers like acetone is essential for monitoring diabetes, as acetone levels correlate with blood glucose levels [111]. PCF sensors have also been employed to detect hydrogen and methane, biomarkers for gastrointestinal diseases, enhancing non-invasive diagnostic capabilities [107]. A summary of various PCF sensors developed for detecting specific biomarkers associated with different diseases is presented in Table 5.

6. Challenges, Interdisciplinary Roles, and Future Perspectives

Despite the immense potential of EBA, several significant challenges must be addressed to facilitate its transition from research laboratories to routine clinical practice. Addressing these hurdles requires a concerted, interdisciplinary effort.

6.1. Standardization and Confounding Factors

A major challenge is the lack of standardization in breath sample collection and analysis protocols. Factors such as breathing rate, exhaled air volume, and the portion of breath sampled can significantly affect VOC concentrations. Furthermore, results can be influenced by numerous confounding factors, including the patient’s diet, smoking habits, medications, and environmental exposures. Establishing universal protocols is essential for ensuring the reproducibility and comparability of data across different studies.

6.2. Sensor Performance and Clinical Validation

While sensor technology has advanced rapidly, challenges related to selectivity, stability, and sensitivity in a complex and humid breath matrix persist. Sensors must be able to reliably detect ppb-level concentrations of specific biomarkers without interference from other compounds or sensor drift over time. Beyond technical performance, the most critical step involves large-scale clinical validation. Many promising biomarkers identified in small-scale studies have yet to be validated in large, diverse patient cohorts, a costly but necessary process for regulatory approval.

6.3. The Essential Role of Interdisciplinary Collaboration

The success of EBA is fundamentally dependent on collaboration between different disciplines. Physicians and clinicians are crucial for defining clinical needs, designing robust studies, and interpreting data within a pathological context. Simultaneously, engineers, chemists, and material scientists are responsible for designing novel sensor materials (such as the nanomaterials and PCFs discussed), improving sensor characteristics, and developing hardware. This synergy ensures that the technology addresses real-world health problems effectively.

6.4. Future Perspectives

The future of EBA is promising, with a clear trajectory toward portable, low-cost, point-of-care devices. Integrating artificial intelligence will enhance diagnostic accuracy and may predict disease progression or treatment response. The ultimate goal is to develop devices that can be used in a general practitioner’s office or even at home, revolutionizing personalized medicine and enabling disease screening on an unprecedented scale.

7. Conclusions

Exhaled breath analysis has evolved from ancient diagnostic practices to a sophisticated, science-driven approach, offering significant advantages over traditional methods. The ability to detect VOCs at very low concentrations enables the identification of biomarkers associated with various diseases, facilitating early diagnosis and monitoring. Modern analytical techniques such as GC-MS, IMS, and electronic noses have significantly improved the accuracy and reliability of EBA. Integrating machine learning algorithms with these techniques has further enhanced their diagnostic capabilities. Among the various sensor technologies, photonic crystal fiber (PCF) sensors stand out due to their exceptional sensitivity and specificity. PCF sensors, which leverage advanced optical properties and the interaction between light and gas molecules, offer unparalleled performance in detecting VOCs in exhaled breath. These sensors have demonstrated high sensitivity in detecting biomarkers for asthma, diabetes, and lung cancer. The hollow-core PCF design, in particular, maximizes the overlap between the light field and the gas molecules, significantly enhancing the sensor’s performance. The superior sensitivity of PCF sensors compared to other methods makes them particularly suitable for early disease detection. For instance, PCF sensors have shown high sensitivity in detecting acetone, a biomarker for diabetes, and ammonia, which is associated with renal diseases. The ability to detect such biomarkers at very low concentrations makes PCF sensors a powerful tool for non-invasive diagnostics. Despite the promising potential of EBA and PCF sensors, some challenges must be addressed. Standardizing breath sampling and analysis procedures is crucial to ensure reproducibility and accuracy. Additionally, environmental factors and individual differences can affect the results, necessitating further research and technological advancements to overcome these challenges.
In summary, exhaled breath analysis, especially when utilizing photonic crystal fiber sensors, shows great promise as a non-invasive diagnostic method with the potential to significantly advance healthcare practices. The continued development and validation of EBA techniques, coupled with advancements in sensor technology, hold the promise of improving patient outcomes and advancing personalized medicine. The development of portable and user-friendly devices for breath analysis will be crucial for the widespread adoption of EBA in clinical practice.

Author Contributions

S.M. (Sajjad Mortazavi) conducted the literature search and drafted the initial manuscript. S.M. (Somayeh Makouei) supervised, reviewed, and revised the manuscript critically for intellectual content. K.A. and S.D. wrote, reviewed, and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No data was used in this research.

Conflicts of Interest

The authors declare that there are no conflicts of interest in the present study.

References

  1. Chen, T.; Liu, T.; Li, T.; Zhao, H.; Chen, Q. Exhaled breath analysis in disease detection. Clin. Chim. Acta 2021, 515, 61–72. [Google Scholar] [CrossRef]
  2. Bajtarevic, A.; Ager, C.; Pienz, M.; Klieber, M.; Schwarz, K.; Ligor, M.; Ligor, T.; Filipiak, W.; Denz, H.; Fiegl, M.; et al. Noninvasive detection of lung cancer by analysis of exhaled breath. BMC Cancer 2009, 9, 348. [Google Scholar] [CrossRef] [PubMed]
  3. Probert, C.S.; Ahmed, I.; Khalid, T.; Johnson, E.; Smith, S.; Ratcliffe, N. Volatile organic compounds as diagnostic biomarkers in gastrointestinal and liver diseases. J. Gastrointest. Liver Dis. 2009, 18, 337–343. [Google Scholar]
  4. Mule, N.M.; Patil, D.D.; Kaur, M. A comprehensive survey on investigation techniques of exhaled breath (EB) for diagnosis of diseases in human body. Inform. Med. Unlocked 2021, 26, 100715. [Google Scholar] [CrossRef]
  5. Phillips, M. Breath tests in medicine. Sci. Am. 1992, 267, 74–79. [Google Scholar] [CrossRef] [PubMed]
  6. Biagini, D.; Lomonaco, T.; Ghimenti, S.; Bellagambi, F.G.; Onor, M.; Scali, M.C.; Barletta, V.; Marzilli, M.; Salvo, P.; Trivella, M.G.; et al. Determination of volatile organic compounds in exhaled breath of heart failure patients by needle trap micro-extraction coupled with gas chromatography-tandem mass spectrometry. J. Breath Res. 2017, 11, 047110. [Google Scholar] [CrossRef] [PubMed]
  7. Issitt, T.; Wiggins, L.; Veysey, M.; Sweeney, S.T.; Brackenbury, W.J.; Redeker, K. Volatile compounds in human breath: Critical review and meta-analysis. J. Breath Res. 2022, 16, 024001. [Google Scholar] [CrossRef]
  8. Issitt, T.; Sweeney, S.T.; Brackenbury, W.J.; Redeker, K.R. Sampling and analysis of low-molecular-weight volatile metabolites in cellular headspace and mouse breath. Metabolites 2022, 12, 599. [Google Scholar] [CrossRef]
  9. Mazzone, P.J. Exhaled breath volatile organic compound biomarkers in lung cancer. J. Breath Res. 2012, 6, 027106. [Google Scholar] [CrossRef]
  10. Buszewski, B.; Kęsy, M.; Ligor, T.; Amann, A. Human exhaled air analytics: Biomarkers of diseases. Biomed. Chromatogr. 2007, 21, 553–566. [Google Scholar] [CrossRef]
  11. Aresta, A.M.; De Vietro, N.; Picciariello, A.; Rotelli, M.T.; Altomare, D.F.; Dezi, A.; Martines, G.; Di Gilio, A.; Palmisani, J.; De Gennaro, G.; et al. Volatile organic compounds determination from intestinal polyps and in exhaled breath by gas chromatography–mass spectrometry. Appl. Sci. 2023, 13, 6083. [Google Scholar] [CrossRef]
  12. Rahimpour, E.; Khoubnasabjafari, M.; Jouyban-Gharamaleki, V.; Jouyban, A. Non-volatile compounds in exhaled breath condensate: Review of methodological aspects. Anal. Bioanal. Chem. 2018, 410, 6411–6440. [Google Scholar] [CrossRef] [PubMed]
  13. Horvath, I.; Lazar, Z.; Gyulai, N.; Kollai, M.; Losonczy, G. Exhaled biomarkers in lung cancer. Eur. Respir. J. 2009, 34, 261–275. [Google Scholar] [CrossRef] [PubMed]
  14. Gashimova, E.; Osipova, A.; Temerdashev, A.; Porkhanov, V.; Polyakov, I.; Perunov, D.; Dmitrieva, E. Exhaled breath analysis using GC-MS and an electronic nose for lung cancer diagnostics. Anal. Methods 2021, 13, 4793–4804. [Google Scholar] [CrossRef] [PubMed]
  15. Mäkitie, A.A.; Almangush, A.; Youssef, O.; Metsälä, M.; Silen, S.; Nixon, I.J.; Haigentz, M., Jr.; Rodrigo, J.P.; Saba, N.F.; Vander Poorten, V.; et al. Exhaled breath analysis in the diagnosis of head and neck cancer. Head Neck 2020, 42, 787–793. [Google Scholar] [CrossRef]
  16. Guo, D.; Zhang, D.; Zhang, L.; Lu, G. Non-invasive blood glucose monitoring for diabetics by means of breath signal analysis. Sens. Actuators B Chem. 2012, 173, 106–113. [Google Scholar] [CrossRef]
  17. Schnabel, R.; Fijten, R.; Smolinska, A.; Dallinga, J.; Boumans, M.L.; Stobberingh, E.; Boots, A.; Roekaerts, P.; Bergmans, D.; van Schooten, F.J. Analysis of volatile organic compounds in exhaled breath to diagnose ventilator-associated pneumonia. Sci. Rep. 2015, 5, 17179. [Google Scholar] [CrossRef]
  18. Amann, A.; Smith, D. Breath analysis for clinical diagnosis and therapeutic monitoring. Siriraj Med. Thurs. Dec 20 2005, 64, 201218. [Google Scholar]
  19. Zheng, W.; Min, Y.; Pang, K.; Wu, D. Sample collection and processing in volatile organic compound analysis for gastrointestinal cancers. Diagnostics 2024, 14, 1563. [Google Scholar] [CrossRef]
  20. Amann, A.; Miekisch, W.; Schubert, J.; Buszewski, B.; Ligor, T.; Jezierski, T.; Pleil, J.; Risby, T. Analysis of exhaled breath for disease detection. Annu. Rev. Anal. Chem. 2014, 7, 455–482. [Google Scholar]
  21. Amor, R.E.; Nakhleh, M.K.; Barash, O.; Haick, H. Breath analysis of cancer in the present and the future. Eur. Respir. Rev. 2019, 28, 190002. [Google Scholar] [CrossRef]
  22. Smith, A.D.; Cowan, J.O.; Filsell, S.; McLachlan, C.; Monti-Sheehan, G.; Jackson, P.; Taylor, D.R. Diagnosing asthma: Comparisons between exhaled nitric oxide measurements and conventional tests. Am. J. Respir. Crit. Care Med. 2004, 169, 473–478. [Google Scholar] [CrossRef]
  23. Silkoff, P.E.; Carlson, M.; Bourke, T.; Katial, R.; Ögren, E.; Szefler, S.J. The Aerocrine exhaled nitric oxide monitoring system NIOX is cleared by the US Food and Drug Administration for monitoring therapy in asthma. J. Allergy Clin. Immunol. 2004, 114, 1241–1256. [Google Scholar] [CrossRef]
  24. Kim, K.H.; Jahan, S.A.; Kabir, E. A review of breath analysis for diagnosis of human health. TrAC Trends Anal. Chem. 2012, 33, 1–8. [Google Scholar] [CrossRef]
  25. Westhoff, M.; Litterst, P.; Maddula, S.; Bödeker, B.; Baumbach, J.I. Statistical and bioinformatical methods to differentiate chronic obstructive pulmonary disease (COPD) including lung cancer from healthy control by breath analysis using ion mobility spectrometry. Int. J. Ion Mobil. Spectrom. 2011, 14, 139–149. [Google Scholar] [CrossRef]
  26. Pradhan, U.U.; Bhat, P. Breathe analysis for medical diagnostics—A review. Int. J. Innov. Res. Dev. 2015, 4, 240–246. [Google Scholar]
  27. Cao, W.; Duan, Y. Breath analysis: Potential for clinical diagnosis and exposure assessment. Clin. Chem. 2006, 52, 800–811. [Google Scholar] [CrossRef] [PubMed]
  28. Miekisch, W.; Schubert, J.K.; Noeldge-Schomburg, G.F. Diagnostic potential of breath analysis—Focus on volatile organic compounds. Clin. Chim. Acta 2004, 347, 25–39. [Google Scholar] [CrossRef]
  29. Schubert, J.K.; Miekisch, W.; Geiger, K.; Nöldge–Schomburg, G.F. Breath analysis in critically ill patients: Potential and limitations. Expert Rev. Mol. Diagn. 2004, 4, 619–629. [Google Scholar] [CrossRef]
  30. Osborne, J.; Sobh, M.; Trudel, G. Carbon monoxide as a clinical marker of hemolysis. Am. J. Hematol. 2023, 98, 1127–1159. [Google Scholar] [CrossRef] [PubMed]
  31. Morimatsu, H.; Takahashi, T.; Maeshima, K.; Inoue, K.; Kawakami, T.; Shimizu, H.; Takeuchi, M.; Yokoyama, M.; Katayama, H.; Morita, K. Increased heme catabolism in critically ill patients: Correlation among exhaled carbon monoxide, arterial carboxyhemoglobin, and serum bilirubin IXα concentrations. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2006, 290, L114–L119. [Google Scholar] [CrossRef]
  32. Di Stefano, M.; Corazza, G.R. Role of hydrogen and methane breath testing in gastrointestinal diseases. Dig. Liver Dis. Suppl. 2009, 3, 40–43. [Google Scholar] [CrossRef]
  33. Rezaie, A.; Buresi, M.; Lembo, A.; Lin, H.; McCallum, R.; Rao, S.; Schmulson, M.; Valdovinos, M.; Zakko, S.; Pimentel, M. Hydrogen and methane-based breath testing in gastrointestinal disorders: The North American consensus. Off. J. Am. Coll. Gastroenterol.|ACG 2017, 112, 775–784. [Google Scholar] [CrossRef]
  34. Lindberg, L.; Brauer, S.; Wollmer, P.; Goldberg, L.; Jones, A.W.; Olsson, S.G. Breath alcohol concentration determined with a new analyzer using free exhalation predicts almost precisely the arterial blood alcohol concentration. Forensic Sci. Int. 2007, 168, 200–207. [Google Scholar] [CrossRef]
  35. Jeong, D.W.; Kim, K.H.; Kim, B.S.; Byun, Y.T. Characteristics of highly sensitive and selective nitric oxide gas sensors using defect-functionalized single-walled carbon nanotubes at room temperature. Appl. Surf. Sci. 2021, 550, 149250. [Google Scholar] [CrossRef]
  36. Yin, F.; Yue, W.; Li, Y.; Gao, S.; Zhang, C.; Kan, H.; Niu, H.; Wang, W.; Guo, Y. Carbon-based nanomaterials for the detection of volatile organic compounds: A review. Carbon 2021, 180, 274–297. [Google Scholar] [CrossRef]
  37. Luo, Y.; Ly, A.; Lahem, D.; Zhang, C.; Debliquy, M. A novel low-concentration isopropanol gas sensor based on Fe-doped ZnO nanoneedles and its gas sensing mechanism. J. Mater. Sci. 2021, 56, 3230–3245. [Google Scholar] [CrossRef]
  38. Liu, W.; Zheng, Y.; Wang, Z.; Wang, Z.; Yang, J.; Chen, M.; Qi, M.; Ur Rehman, S.; Shum, P.P.; Zhu, L.; et al. Ultrasensitive exhaled breath sensors based on anti-resonant hollow core fiber with in situ grown ZnO-Bi2O3 nanosheets. Adv. Mater. Interfaces 2021, 8, 2001978. [Google Scholar] [CrossRef]
  39. Nath, N.; Kumar, A.; Chakroborty, S.; Soren, S.; Barik, A.; Pal, K.; de Souza, F.G., Jr. Carbon nanostructure embedded novel sensor implementation for detection of aromatic volatile organic compounds: An organized review. ACS Omega 2023, 8, 4436–4452. [Google Scholar] [CrossRef]
  40. Sajor, N.J.; Foronda, J.R.; Olarve, R.S.; Torre, H.D.; Santos, M.G.; Lopez, T.B.; Haygood, K.J.; Santos, G.N.; Koledov, V.; Gratowski, S.V. Synthesis of metal oxide nanomaterials for early lung disease detection. J. Phys. Conf. Ser. 2020, 1461, 012149. [Google Scholar] [CrossRef]
  41. Salimi, M.; Hosseini, S.M.R.M. Smartphone-based detection of lung cancer-related volatile organic compounds (VOCs) using rapid synthesized ZnO nanosheet. Sens. Actuators B Chem. 2021, 344, 130127. [Google Scholar] [CrossRef]
  42. Ray, B.; Desai, S.M.; Parmar, S.; Datar, S. Polymer-Modified Quartz Tuning Forks for Breath Biomarker Sensing. Eng. Proc. 2021, 6, 62. [Google Scholar]
  43. Zhou, X.; Xue, Z.; Chen, X.; Huang, C.; Bai, W.; Lu, Z.; Wang, T. Nanomaterial-based gas sensors used for breath diagnosis. J. Mater. Chem. B 2020, 8, 3231–3248. [Google Scholar] [CrossRef]
  44. Maciel, M.; Sankari, S.; Woollam, M.; Agarwal, M. Optimization of metal oxide nanosensors and development of a feature extraction algorithm to analyze VOC profiles in exhaled breath. IEEE Sens. J. 2023, 23, 16571–16578. [Google Scholar] [CrossRef]
  45. Sun, J.Y.; Salahuddin, U.; Zhu, C.; Gao, P.X. Medical Diagnosis Using Volatile Organic Compounds Sensors. Int. J. High Speed Electron. Syst. 2022, 31, 2240004. [Google Scholar] [CrossRef]
  46. Velumani, M.; Prasanth, A.; Narasimman, S.; Chandrasekhar, A.; Sampson, A.; Meher, S.R.; Rajalingam, S.; Rufus, E.; Alex, Z.C. Nanomaterial-Based Sensors for Exhaled Breath Analysis: A Review. Coatings 2022, 12, 1989. [Google Scholar] [CrossRef]
  47. Saidi, T.; Palmowski, D.; Babicz-Kiewlicz, S.; Welearegay, T.G.; El Bari, N.; Ionescu, R.; Smulko, J.; Bouchikhi, B. Exhaled breath gas sensing using pristine and functionalized WO3 nanowire sensors enhanced by UV-light irradiation. Sens. Actuators B Chem. 2018, 273, 1719–1729. [Google Scholar] [CrossRef]
  48. Postica, V.; Vahl, A.; Santos-Carballal, D.; Dankwort, T.; Kienle, L.; Hoppe, M.; Cadi-Essadek, A.; De Leeuw, N.H.; Terasa, M.I.; Adelung, R.; et al. Tuning ZnO sensors reactivity toward volatile organic compounds via Ag doping and nanoparticle functionalization. ACS Appl. Mater. Interfaces 2019, 11, 31452–31466. [Google Scholar] [CrossRef]
  49. Hanh, N.H.; Ngoc, T.M.; Van Duy, L.; Hung, C.M.; Van Duy, N.; Hoa, N.D. A comparative study on the VOCs gas sensing properties of Zn2SnO4 nanoparticles, hollow cubes, and hollow octahedra towards exhaled breath analysis. Sens. Actuators B Chem. 2021, 343, 130147. [Google Scholar] [CrossRef]
  50. Lagopati, N.; Valamvanos, T.F.; Proutsou, V.; Karachalios, K.; Pippa, N.; Gatou, M.A.; Vagena, I.A.; Cela, S.; Pavlatou, E.A.; Gazouli, M.; et al. The role of nano-sensors in breath analysis for early and non-invasive disease diagnosis. Chemosensors 2023, 11, 317. [Google Scholar] [CrossRef]
  51. Kalidoss, R.; Umapathy, S.; Anandan, R.; Ganesh, V.; Sivalingam, Y. Comparative study on the preparation and gas sensing properties of reduced graphene oxide/SnO2 binary nanocomposite for detection of acetone in exhaled breath. Anal. Chem. 2019, 91, 5116–5124. [Google Scholar] [CrossRef]
  52. Andre, R.S.; Sanfelice, R.C.; Pavinatto, A.; Mattoso, L.H.; Correa, D.S. Hybrid nanomaterials designed for volatile organic compounds sensors: A review. Mater. Des. 2018, 156, 154–166. [Google Scholar] [CrossRef]
  53. Yang, D.; Gopal, R.A.; Lkhagvaa, T.; Choi, D. Metal-oxide gas sensors for exhaled-breath analysis: A review. Meas. Sci. Technol. 2021, 32, 102004. [Google Scholar] [CrossRef]
  54. Zonta, G.; Rispoli, G.; Malagù, C.; Astolfi, M. Overview of gas sensors focusing on chemoresistive ones for cancer detection. Chemosensors 2023, 11, 519. [Google Scholar] [CrossRef]
  55. Kim, J.N.; Kim, H.J. A Chemoresistive Gas Sensor Readout Integrated Circuit with Sensor Offset Cancellation Technique. IEEE Access 2023, 11, 85405–85413. [Google Scholar] [CrossRef]
  56. Vajhadin, F.; Mazloum-Ardakani, M.; Amini, A. Metal oxide-based gas sensors for the detection of exhaled breath markers. Med. Devices Sens. 2021, 4, e10161. [Google Scholar] [CrossRef] [PubMed]
  57. Aroutiounian, V.M. Microelectronic Gas sensors for Non-invasive Analysis of Exhaled Gases. Med. Sci. Armen. 2020, 60, 3–15. [Google Scholar]
  58. Wu, T.C.; De Luca, A.; Zhong, Q.; Zhu, X.; Ogbeide, O.; Um, D.S.; Hu, G.; Albrow-Owen, T.; Udrea, F.; Hasan, T. Inkjet-printed CMOS-integrated graphene–metal oxide sensors for breath analysis. npj 2D Mater. Appl. 2019, 3, 42. [Google Scholar] [CrossRef]
  59. Barreca, D.; Maccato, C.; Gasparotto, A. Metal oxide nanosystems as chemoresistive gas sensors for chemical warfare agents: A focused review. Adv. Mater. Interfaces 2022, 9, 2102525. [Google Scholar] [CrossRef]
  60. Janfaza, S.; Banan Nojavani, M.; Nikkhah, M.; Alizadeh, T.; Esfandiar, A.; Ganjali, M.R. A selective chemiresistive sensor for the cancer-related volatile organic compound hexanal by using molecularly imprinted polymers and multiwalled carbon nanotubes. Microchim. Acta 2019, 186, 137. [Google Scholar] [CrossRef]
  61. Rodríguez-Aguilar, M.; de León-Martínez, L.D.; Gorocica-Rosete, P.; Pérez-Padilla, R.; Domínguez-Reyes, C.A.; Tenorio-Torres, J.A.; Ornelas-Rebolledo, O.; Mehta, G.; Zamora-Mendoza, B.N.; Flores-Ramírez, R. Application of chemoresistive gas sensors and chemometric analysis to differentiate the fingerprints of global volatile organic compounds from diseases. Preliminary results of COPD, lung cancer and breast cancer. Clin. Chim. Acta 2021, 518, 83–92. [Google Scholar] [CrossRef] [PubMed]
  62. Jing, Q.; Gong, C.; Bian, W.; Tian, Q.; Zhang, Y.; Chen, N.; Xu, C.; Sun, N.; Wang, X.; Li, C.; et al. Ultrasensitive chemiresistive gas sensor can diagnose asthma and monitor its severity by analyzing its biomarker H2S: An experimental, clinical, and theoretical study. ACS Sens. 2022, 7, 2243–2252. [Google Scholar] [CrossRef]
  63. Shang, G.; Dinh, D.; Mercer, T.; Yan, S.; Wang, S.; Malaei, B.; Luo, J.; Lu, S.; Zhong, C.J. Chemiresistive sensor array with nanostructured interfaces for detection of human breaths with simulated lung cancer breath VOCs. ACS Sens. 2023, 8, 1328–1338. [Google Scholar] [CrossRef]
  64. Wilson, A.D.; Baietto, M. Applications and advances in electronic-nose technologies. Sensors 2009, 9, 5099–5148. [Google Scholar] [CrossRef] [PubMed]
  65. Hakim, M.; Broza, Y.Y.; Barash, O.; Peled, N.; Phillips, M.; Amann, A.; Haick, H. Volatile organic compounds of lung cancer and possible biochemical pathways. Chem. Rev. 2012, 112, 5949–5966. [Google Scholar] [CrossRef]
  66. Dragonieri, S.; Schot, R.; Mertens, B.J.; Le Cessie, S.; Gauw, S.A.; Spanevello, A.; Resta, O.; Willard, N.P.; Vink, T.J.; Rabe, K.F.; et al. An electronic nose in the discrimination of patients with asthma and controls. J. Allergy Clin. Immunol. 2007, 120, 856–862. [Google Scholar] [CrossRef] [PubMed]
  67. Korotcenkov, G. Metal oxides for solid-state gas sensors: What determines our choice? Mater. Sci. Eng. B 2007, 139, 1–23. [Google Scholar] [CrossRef]
  68. Righettoni, M.; Amann, A.; Pratsinis, S.E. Breath analysis by nanostructured metal oxides as chemo-resistive gas sensors. Mater. Today 2015, 18, 163–171. [Google Scholar] [CrossRef]
  69. van Geffen, W.H.; Lamote, K.; Costantini, A.; Hendriks, L.E.; Rahman, N.M.; Blum, T.G.; Van Meerbeeck, J. The electronic nose: Emerging biomarkers in lung cancer diagnostics. Breathe 2020, 15, e135–e141. [Google Scholar] [CrossRef]
  70. Tripathy, P.; Biswas, S. Mechanical and thermal properties of mineral fiber based polymeric nanocomposites: A review. Polym.-Plast. Technol. Mater. 2022, 61, 1385–1410. [Google Scholar] [CrossRef]
  71. Arshak, K.; Moore, E.; Lyons, G.M.; Harris, J.; Clifford, S. A review of gas sensors employed in electronic nose applications. Sens. Rev. 2004, 24, 181–198. [Google Scholar] [CrossRef]
  72. Xie, J.; Zhang, L.; Xing, H.; Bai, P.; Liu, B.; Wang, C.; Lei, K.; Wang, H.; Peng, S.; Yang, S. Gas sensing of ordered and disordered structure SiO2 and their adsorption behavior based on quartz crystal microbalance. Sens. Actuators B Chem. 2020, 305, 127479. [Google Scholar] [CrossRef]
  73. Smith, D.; Španěl, P. Selected ion flow tube mass spectrometry (SIFT-MS) for on-line trace gas analysis. Mass Spectrom. Rev. 2005, 24, 661–700. [Google Scholar] [CrossRef]
  74. Su, R.; Yang, T.; Zhang, X.; Li, N.; Zhai, X.; Chen, H. Mass spectrometry for breath analysis. TrAC Trends Anal. Chem. 2023, 158, 116823. [Google Scholar] [CrossRef]
  75. Röck, F.; Barsan, N.; Weimar, U. Electronic nose: Current status and future trends. Chem. Rev. 2008, 108, 705–725. [Google Scholar] [CrossRef]
  76. Kalidoss, R.; Surya, V.J.; Sivalingam, Y. Recent progress in graphene derivatives/metal oxides binary nanocomposites based chemi-resistive sensors for disease diagnosis by breath analysis. Curr. Anal. Chem. 2022, 18, 563–576. [Google Scholar] [CrossRef]
  77. Nakhleh, M.K.; Amal, H.; Jeries, R.; Broza, Y.Y.; Aboud, M.; Gharra, A.; Ivgi, H.; Khatib, S.; Badarneh, S.; Har-Shai, L.; et al. Diagnosis and classification of 17 diseases from 1404 subjects via pattern analysis of exhaled molecules. ACS Nano 2017, 11, 112–125. [Google Scholar] [CrossRef] [PubMed]
  78. Harris, D.C. Quantitative Chemical Analysis; Macmillan: New York, NY, USA, 2010. [Google Scholar]
  79. de Lacy Costello, B.; Amann, A.; Al-Kateb, H.; Flynn, C.; Filipiak, W.; Khalid, T.; Osborne, D.; Ratcliffe, N.M. A review of the volatiles from the healthy human body. J. Breath Res. 2014, 8, 014001. [Google Scholar] [CrossRef] [PubMed]
  80. Ligor, M.; Ligor, T.; Bajtarevic, A.; Ager, C.; Pienz, M.; Klieber, M.; Denz, H.; Fiegl, M.; Hilbe, W.; Weiss, W.; et al. Determination of volatile organic compounds in exhaled breath of patients with lung cancer using solid phase microextraction and gas chromatography mass spectrometry. Clin. Chem. Lab. Med. 2009, 47, 550–560. [Google Scholar] [CrossRef]
  81. Eiceman, G.A.; Karpas, Z. Ion Mobility Spectrometry; CRC Press: Boca Raton, FL, USA, 2005. [Google Scholar]
  82. Selvaraj, R.; Vasa, N.J.; Nagendra, S.S.; Mizaikoff, B. Advances in mid-infrared spectroscopy-based sensing techniques for exhaled breath diagnostics. Molecules 2020, 25, 2227. [Google Scholar] [CrossRef]
  83. Hannemann, M.; Antufjew, A.; Borgmann, K.; Hempel, F.; Ittermann, T.; Welzel, S.; Weltmann, K.D.; Völzke, H.; Röpcke, J. Influence of age and sex in exhaled breath samples investigated by means of infrared laser absorption spectroscopy. J. Breath Res. 2011, 5, 027101. [Google Scholar] [CrossRef]
  84. Wojtas, J.; Tittel, F.K.; Stacewicz, T.; Bielecki, Z.; Lewicki, R.; Mikolajczyk, J.; Nowakowski, M.; Szabra, D.; Stefanski, P.; Tarka, J. Cavity-enhanced absorption spectroscopy and photoacoustic spectroscopy for human breath analysis. Int. J. Thermophys. 2014, 35, 2215–2225. [Google Scholar] [CrossRef]
  85. Petersen, J.C.; Lamard, L.; Feng, Y.; Focant, J.F.; Lassen, M. Quartz-enhanced photoacoustic spectroscopy as a platform for non-invasive trace gas analyser targeting breath analysis. arXiv 2017, arXiv:1704.07442. [Google Scholar]
  86. Yi, H.; Laurent, O.; Schilt, S.; Ramonet, M.; Gao, X.; Dong, L.; Chen, W. Simultaneous Monitoring of Atmospheric CH4, N2O, and H2O Using a Single Gas Sensor Based on Mid-IR Quartz-Enhanced Photoacoustic Spectroscopy. Anal. Chem. 2022, 94, 17522–17532. [Google Scholar] [CrossRef]
  87. Wojtas, J.; Gluszek, A.; Hudzikowski, A.; Tittel, F.K. Mid-infrared trace gas sensor technology based on intracavity quartz-enhanced photoacoustic spectroscopy. Sensors 2017, 17, 513. [Google Scholar] [CrossRef] [PubMed]
  88. Rabee, A.S.H.; Hameed, M.F.O.; Heikal, A.M.; Obayya, S.S.A. Highly sensitive photonic crystal fiber gas sensor. Optik 2019, 188, 78–86. [Google Scholar] [CrossRef]
  89. Yang, J.; Che, X.; Shen, R.; Wang, C.; Li, X.; Chen, W. High-sensitivity photonic crystal fiber long-period grating methane sensor with cryptophane-A-6Me absorbed on a PAA-CNTs/PAH nanofilm. Opt. Express 2017, 25, 20258–20267. [Google Scholar] [CrossRef] [PubMed]
  90. Saber, A.; Hameed, M.F.O.; Heikal, A.M.; Obayya, S.S.A. Novel optical gas sensor based on photonic crystal fiber. In Proceedings of the Optical Components and Materials XVI, San Francisco, CA, USA, 2–7 February 2019; Volume 10914, pp. 354–360. [Google Scholar]
  91. Mishra, G.P.; Kumar, D.; Chaudhary, V.S.; Kumar, S. Design and sensitivity improvement of microstructured-core photonic crystal fiber based sensor for methane and hydrogen fluoride detection. IEEE Sens. J. 2021, 22, 1265–1272. [Google Scholar] [CrossRef]
  92. Wu, B.; Lu, Y.; Hao, C.; Duan, L.; Musideke, M.; Yao, J. A photonic crystal fiber sensor based on differential optical absorption spectroscopy for mixed gases detection. Optik 2014, 125, 2909–2911. [Google Scholar] [CrossRef]
  93. Abbaszadeh, A.; Makouei, S.; Meshgini, S. High sensitive triangular photonic crystal fiber sensor design applicable for gas detection. Adv. Electromagn. 2021, 10, 1–5. [Google Scholar] [CrossRef]
  94. Britto, E.C.; Nizar, S.M.; Krishnan, P. A highly sensitive photonic crystal fiber gas sensor for the detection of sulfur dioxide. Silicon 2022, 14, 12665–12674. [Google Scholar] [CrossRef]
  95. Islam, M.I.; Ahmed, K.; Sen, S.; Chowdhury, S.; Paul, B.K.; Islam, M.S.; Miah, M.B.A.; Asaduzzaman, S. Design and optimization of photonic crystal fiber based sensor for gas condensate and air pollution monitoring. Photonic Sens. 2017, 7, 234–245. [Google Scholar] [CrossRef]
  96. Rifat, A.A.; Ahmed, K.; Asaduzzaman, S.; Paul, B.K.; Ahmed, R. Development of photonic crystal fiber-based gas/chemical sensors. Comput. Photonic Sens. 2019, 287–317. [Google Scholar]
  97. Cubillas, A.M.; Unterkofler, S.; Euser, T.G.; Etzold, B.J.; Jones, A.C.; Sadler, P.J.; Wasserscheid, P.; Russell, P.S.J. Photonic crystal fibres for chemical sensing and photochemistry. Chem. Soc. Rev. 2013, 42, 8629–8648. [Google Scholar] [CrossRef]
  98. Hu, H.F.; Zhao, Y.; Zhang, Y.N.; Yang, Y. Characterization of infrared gas sensors employing hollow-core photonic crystal fibers. Instrum. Sci. Technol. 2016, 44, 495–503. [Google Scholar] [CrossRef]
  99. Yoon, J.W.; Lee, J.H. Toward breath analysis on a chip for disease diagnosis using semiconductor-based chemiresistors: Recent progress and future perspectives. Lab A Chip 2017, 17, 3537–3557. [Google Scholar] [CrossRef]
  100. Li, Z.; Sie, S.H.; Lee, J.L.; Chen, Y.R.; Chou, T.I.; Wu, P.C.; Chuang, Y.T.; Lin, Y.T.; Chen, I.C.; Lu, C.C.; et al. A miniature electronic nose for breath analysis. In Proceedings of the 2021 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 11–15 December 2021; IEEE: New York, NY, USA, 2021; pp. 35.2.1–35.2.4. [Google Scholar]
  101. Kononov, A.; Korotetsky, B.; Jahatspanian, I.; Gubal, A.; Vasiliev, A.; Arsenjev, A.; Nefedov, A.; Barchuk, A.; Gorbunov, I.; Kozyrev, K.; et al. Online breath analysis using metal oxide semiconductor sensors (electronic nose) for diagnosis of lung cancer. J. Breath Res. 2019, 14, 016004. [Google Scholar] [CrossRef]
  102. Wilson, A.D. Advances in electronic-nose technologies for the detection of volatile biomarker metabolites in the human breath. Metabolites 2015, 5, 140–163. [Google Scholar] [CrossRef]
  103. Scarlata, S.; Pennazza, G.; Santonico, M.; Pedone, C.; Antonelli Incalzi, R. Exhaled breath analysis by electronic nose in respiratory diseases. Expert Rev. Mol. Diagn. 2015, 15, 933–956. [Google Scholar] [CrossRef]
  104. Abbaszadeh, A.; Makouei, S.; Meshgini, S. Ammonia measurement in exhaled human breath using PCF sensor for medical applications. Photonics Nanostruct.-Fundam. Appl. 2021, 44, 100917. [Google Scholar] [CrossRef]
  105. Chow, K.K.; Short, M.; Lam, S.; McWilliams, A.; Zeng, H. A Raman cell based on hollow core photonic crystal fiber for human breath analysis. Med. Phys. 2014, 41, 092701. [Google Scholar] [CrossRef]
  106. Rifat, A.A.; Haider, F.; Ahmed, R.; Mahdiraji, G.A.; Mahamd Adikan, F.R.; Miroshnichenko, A.E. Highly sensitive selectively coated photonic crystal fiber-based plasmonic sensor. Opt. Lett. 2018, 43, 891–894. [Google Scholar] [CrossRef]
  107. Hanf, S.; Bögözi, T.; Keiner, R.; Frosch, T.; Popp, J. Fast and highly sensitive fiber-enhanced Raman spectroscopic monitoring of molecular H2 and CH4 for point-of-care diagnosis of malabsorption disorders in exhaled human breath. Anal. Chem. 2015, 87, 982–988. [Google Scholar] [CrossRef]
  108. Mortazavi, S.; Makouei, S.; Garamaleki, S.M. Hollow core photonic crystal fiber based carbon monoxide sensor design applicable for hyperbilirubinemia diagnosis. Opt. Eng. 2023, 62, 066105. [Google Scholar] [CrossRef]
  109. Akowuah, E.K.; Gorman, T.; Ademgil, H.; Haxha, S.; Robinson, G.K.; Oliver, J.V. Numerical analysis of a photonic crystal fiber for biosensing applications. IEEE J. Quantum Electron. 2012, 48, 1403–1410. [Google Scholar] [CrossRef]
  110. De, M.; Gangopadhyay, T.K.; Singh, V.K. Prospects of photonic crystal fiber as physical sensor: An overview. Sensors 2019, 19, 464. [Google Scholar] [CrossRef] [PubMed]
  111. Al Mahfuz, M.; Mollah, M.A.; Momota, M.R.; Paul, A.K.; Masud, A.; Akter, S.; Hasan, M.R. Highly sensitive photonic crystal fiber plasmonic biosensor: Design and analysis. Opt. Mater. 2019, 90, 315–321. [Google Scholar] [CrossRef]
  112. Mehaney, A.; Elsayed, H.A.; Ahmed, A.M. Detection of Isoprene Traces in Exhaled Breath by Using Photonic Crystals as a Biomarker For Chronic Liver Fibrosis Disease. 2021. Available online: https://scispace.com/pdf/detection-of-isoprene-traces-in-exhaled-breath-by-using-39ago6x9ep.pdf (accessed on 21 August 2025).
  113. Zhou, D.; Wang, Q.; Lan, Z.; Chen, Y.; Peng, Z.; Zhang, L.; Liu, Y. Liquid-crystal-based fiber laser sensor for non-invasive gas detection. Opt. Lett. 2023, 48, 4508–4511. [Google Scholar] [CrossRef] [PubMed]
  114. Chen, D.; Liu, S.; Shen, Z.; Niu, P.; Cheng, W.; Xu, F. Respiration monitoring using antiresonant reflecting guidance in selectively infiltrated hollow core photonic crystal fiber. IEEE Sens. J. 2023, 23, 26004–26011. [Google Scholar] [CrossRef]
  115. Nizar, S.M.; Kesavaraman, B.; Priyanka, E.; Jayasri, R. Detection of immunoglobulin G (IgG) and immunoglobulin M (IgM) antibodies using circular photonic crystal fiber sensor. J. Phys. Conf. Ser. 2021, 1717, 012039. [Google Scholar] [CrossRef]
  116. Shao, M.; Sun, H.; Liang, J.; Han, L.; Feng, D. In-fiber Michelson interferometer in photonic crystal fiber for humidity measurement. IEEE Sens. J. 2020, 21, 1561–1567. [Google Scholar] [CrossRef]
  117. Chi, Z.; Li, M.; Xu, J.; Yang, L. A photonic crystal fiber–based fluorescence sensor for simultaneous and sensitive detection of lactic acid enantiomers. Anal. Bioanal. Chem. 2022, 414, 1641–1649. [Google Scholar] [CrossRef]
  118. Shirmohammad, M.; Short, M.A.; Zeng, H. A new gas analysis method based on single-beam excitation stimulated Raman scattering in hollow core photonic crystal fiber enhanced Raman spectroscopy. Bioengineering 2023, 10, 1161. [Google Scholar] [CrossRef]
  119. Ferdous, A.H.M.I.; Mynuddin, M.; Noor, K.S. High-performance sulphur dioxide sensor: Unveiling the potential of photonic crystal fibre technology. IET Nanodielectr. 2024, 7, 262–272. [Google Scholar] [CrossRef]
  120. Singh, S.; Kumar, D.; Sahu, A.; Chaudhary, V.S.; Singh, G.; Kumar, S. Photonic Crystal Fiber Based Sensors for Various Cancer Detection in Human Body-A Review. IEEE Sens. J. 2025, 25, 5956–5968. [Google Scholar] [CrossRef]
  121. Sharif, V.; Saberi, H.; Pakarzadeh, H. Designing a terahertz optical sensor based on helically twisted photonic crystal fiber for toxic gas sensing. Sci. Rep. 2025, 15, 2268. [Google Scholar] [CrossRef] [PubMed]
  122. Chao, X.; Zeng, F.; Cheng, H.; Jiang, X.; Huang, Z.; Yao, Q.; Tang, J. Performance of Hollow-Core Photonic Crystal Fiber-Based Trace C2H2 Detection System. IEEE Trans. Instrum. Meas. 2025, 74, 9503410. [Google Scholar] [CrossRef]
Figure 2. Volatile organic compounds (VOCs) in exhaled breath as non-invasive biomarkers. This figure shows the respiratory system and lists prominent VOCs that can be detected in breath, along with their potential diagnostic utility for a range of diseases and conditions.
Figure 2. Volatile organic compounds (VOCs) in exhaled breath as non-invasive biomarkers. This figure shows the respiratory system and lists prominent VOCs that can be detected in breath, along with their potential diagnostic utility for a range of diseases and conditions.
Photonics 12 00848 g002
Figure 3. Schematic overview of various sensor technologies utilized in exhaled breath analysis.
Figure 3. Schematic overview of various sensor technologies utilized in exhaled breath analysis.
Photonics 12 00848 g003
Figure 4. Exhaled breath analysis: VOC interaction with the sensor generates an electrical signal, which is amplified and displayed for biomarker detection.
Figure 4. Exhaled breath analysis: VOC interaction with the sensor generates an electrical signal, which is amplified and displayed for biomarker detection.
Photonics 12 00848 g004
Figure 5. Diagram illustrating the suggested experimental configuration for gas sensing with the PCF [91].
Figure 5. Diagram illustrating the suggested experimental configuration for gas sensing with the PCF [91].
Photonics 12 00848 g005
Table 2. Recent advancements in nanomaterial-based sensors.
Table 2. Recent advancements in nanomaterial-based sensors.
ApplicationSensor TypeNanomaterialsTarget GasKey FeaturesReferences
Early-stage disease diagnosisChemoresistive VOC sensorsNanomaterials (general)VOCsHigh sensitivity, real-time, non-invasive diagnosis[45]
Lung cancer detectionResistive gas sensorZnO nanosheetsDiethyl ketone, acetone, isopropanolHigh sensitivity, smartphone integration[41]
Diabetes diagnosisChemiresistive sensorsZnO-Bi2O3 nanosheetsAcetoneHigh sensitivity and selectivity[38]
General disease diagnosisVarious sensor technologiesNanomaterials (general)VOCsNon-invasive, highly selective, sensitive, robust sensors[46]
Non-invasive medical diagnosticsMetal oxide semiconductor sensorsWO3 nanowiresVOCsEnhanced response with UV-light irradiation, selective towards VOCs[47]
VOC detection for health and environmental applicationsGas sensorsAg-doped ZnOPropanol, acetone, methaneHigh sensitivity at low operating temperatures[48]
Diabetic diagnosis via exhaled breathGas sensorsZn2SnO4 nanoparticlesAcetoneHigh sensitivity, good selectivity, and stability[49]
General disease diagnosis via breath analysisNM-based gas sensorsNanomaterials (general)VOCsHigh surface-to-volume ratio, controllable morphology, potential for miniaturization[43]
Early and non-invasive disease diagnosisGas sensorsNanomaterials (general)VOCsAccurate detection, potential for commercial use as disease self-test kits[50]
VOC detection in human breath for diabetes and respiratory diseasesBinary nanocompositesReduced graphene oxide/SnO2AcetoneEnhanced acetone sensing performance distinguishes between healthy and diabetic subjects[51]
Portable, low-cost sensors for environmental and health applicationsHybrid nanomaterial sensorsConducting polymers, metal oxides, graphenesVOCsSuperior sensitivity, low detection limits, potential for miniaturization, and versatility[52]
Table 3. Comparison of sensor technologies in eNose applications.
Table 3. Comparison of sensor technologies in eNose applications.
Sensor TypePrinciple of OperationAdvantagesDisadvantagesReferences
Metal Oxide Sensor (MOS)Change in electrical conductivity upon VOC adsorptionHigh sensitivity, fast response, low cost, mature technologyLimited selectivity, cross-sensitivity, susceptibility to humidity and temperature variations[68,69]
Conducting Polymer (CP) SensorChange in electrical conductivity upon VOC adsorptionFlexibility in design, ease of fabrication, miniaturization potentialLow sensitivity compared to MOSs, potential for drift and aging[70]
Quartz Crystal Microbalance (QCM) SensorChange in resonant frequency due to mass change upon VOC adsorptionHigh sensitivity, ability to detect trace amounts of VOCsRequires selective coatings, susceptibility to interference from other gases[72]
Mass Spectrometry (MS) SensorIonization and separation of VOCs based on mass-to-charge ratioUnparalleled selectivity and sensitivityBulky, expensive, complex operation[74]
Table 4. Comparison of sensor technologies.
Table 4. Comparison of sensor technologies.
Sensor TypeAdvantagesDisadvantagesCostFeaturesGas DetectionLimitationsResponse TimeSensitivityReferences
Nanomaterial-BasedHigh sensitivity and selectivity, low cost, portable, non-invasive, low power consumptionSensitive to humidity, may require pre-treatment, stability issuesLowUses metal oxides, carbon nanotubes, graphene; can be miniaturizedNO, NH3, H2S, acetone, other VOCsHumidity sensitivity, stability issues in varying conditionsSeconds to minutesHigh sensitivity (ppb level)[99]
Chemiresistive GasHigh sensitivity, capable of detecting multiple gases, pattern recognition, portableMay require calibration, influenced by environmental conditions, moderate costModerateUses an array of sensors integrated with AI and pattern recognition systemsH2S, NH3, NO, VOCsEnvironmental sensitivity, potential need for recalibrationSeconds to minutesHigh sensitivity (ppb level)[100,101]
Electronic Nose (E-nose)High sensitivity, can detect multiple gases, pattern recognition, portableMay require calibration, influenced by environmental conditions, moderate costModerateUses an array of sensors integrated with AI and pattern recognition systemsH2S, NH3, NO, VOCsEnvironmental sensitivity, potential need for recalibrationSeconds to minutesHigh sensitivity (ppb level)[100,101]
Infrared SpectroscopyHigh precision and accuracy, can detect a wide range of gases, non-invasiveExpensive, requires skilled operation, large equipmentVery highHigh precision, capable of detecting a wide range of gasesMultiple VOCs, CO2, CH4Expensive, not easily portableSeconds to minutesHigh sensitivity (ppb to ppm level)[102]
ChromatographyHigh accuracy, can detect very low concentrations of gases, gold standard for analysisExpensive, time-consuming, requires skilled operation, non-portableVery highHigh accuracy and sensitivity, capable of comprehensive gas analysisWide range of VOCsExpensive, non-portable, requires skilled operationMinutes to hoursVery high sensitivity (ppb level)[103]
Photonic Crystal FiberHigh sensitivity, selectivity, low interference from humidity, non-invasive, highly stableHigh cost, requires specialized equipment and knowledge for operationModerateUses photonic crystals, highly selective and stable, less affected by humidityMultiple VOCs, NOxHigh cost, requires specialized operationSeconds to minutesVery high sensitivity (ppb level)[43]
Table 5. PCF sensors for disease detection using EBA.
Table 5. PCF sensors for disease detection using EBA.
TitleGas/AnalyteDisease DiagnosedSensitivityAdvantages & FeaturesWavelength/BandReferences
Ammonia Measurement via PCFAmmoniaKidney disease63.18%Moisture-resistant, FEM-optimized1.544 µm[104]
CO Detection via Hollow-Core PCFCarbon monoxide (CO)Hyperbilirubinemia64.28%Early jaundice detection1.567 µm[108]
Isoprene Detection via PCIsopreneLiver fibrosis0.321 nm/ppmTamm plasmon-based, non-invasiveVisible-NIR[112]
Liquid Crystal PCF SensorAcetoneDiabetes65 ppm (LOD)Compact, temp-compensatedN/A[113]
Antiresonant PCF for RespirationWater vaporRespiration monitoringWearable, tracks breathing~1.5 µm[114]
Circular PCF for IgG/IgMAntibodiesCOVID-19High (qualitative)Low loss, blood-serum compatibleNot stated[115]
PCF Humidity SensorHumidityRespiratory rate−0.166 dB/%RHLow temp cross-sensitivity~1.55 µm[116]
PCF Fluorescence for Lactic AcidLactic acidSepsis/cancer0.8–9.5 µM (LOD)Dual-channel, enzymaticFluorescence[117]
Raman-Enhanced HC-PCFVOCs (propene, H2, CO2)Lung cancerSRS-basedHigh enhancement, broadbandVisible–NIR[118]
High-Performance SO2 PCF SensorSulfur dioxideRespiratory risk87.39%Low loss, THz-compatible1.8 THz[119]
PCF Sensors for Cancer Detection (Review)Various biomarkersMultiple cancers— (review)SPR/SERS/Interferometry OverviewVarious[120]
Helically Twisted PCF SensorCO, NOx, SOxToxic gas exposure3000 RIU−1100% relative sensitivity, simple build0.2–3.0 THz[121]
Trace C2H2 Detection via HC-PCFAcetylene (C2H2)VOCs in breath49 ppm (LOD)Linear response, real-time sensing1–500 Hz[122]
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

Mortazavi, S.; Makouei, S.; Abbasian, K.; Danishvar, S. Exhaled Breath Analysis (EBA): A Comprehensive Review of Non-Invasive Diagnostic Techniques for Disease Detection. Photonics 2025, 12, 848. https://doi.org/10.3390/photonics12090848

AMA Style

Mortazavi S, Makouei S, Abbasian K, Danishvar S. Exhaled Breath Analysis (EBA): A Comprehensive Review of Non-Invasive Diagnostic Techniques for Disease Detection. Photonics. 2025; 12(9):848. https://doi.org/10.3390/photonics12090848

Chicago/Turabian Style

Mortazavi, Sajjad, Somayeh Makouei, Karim Abbasian, and Sebelan Danishvar. 2025. "Exhaled Breath Analysis (EBA): A Comprehensive Review of Non-Invasive Diagnostic Techniques for Disease Detection" Photonics 12, no. 9: 848. https://doi.org/10.3390/photonics12090848

APA Style

Mortazavi, S., Makouei, S., Abbasian, K., & Danishvar, S. (2025). Exhaled Breath Analysis (EBA): A Comprehensive Review of Non-Invasive Diagnostic Techniques for Disease Detection. Photonics, 12(9), 848. https://doi.org/10.3390/photonics12090848

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