Next Article in Journal
DNA Dyes—Highly Sensitive Reporters of Cell Quantification: Comparison with Other Cell Quantification Methods
Next Article in Special Issue
Application of Capillary and Free-Flow Zone Electrophoresis for Analysis and Purification of Antimicrobial β-Alanyl-Tyrosine from Hemolymph of Fleshfly Neobellieria bullata
Previous Article in Journal
Analysis of Molar Substitution of Hydroxybutyl Group by Zeisel Reaction in Starch Ethers
Previous Article in Special Issue
Identification of Bacteria Associated with Post-Operative Wounds of Patients with the Use of Matrix-Assisted Laser Desorption/Ionization Time-of-Flight Mass Spectrometry Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Breath Biomarkers in Diagnostic Applications

1
Department of Sensory Analytics and Technologies, Fraunhofer Institute for Process Engineering and Packaging IVV, Giggenhauser Straße 35, 85354 Freising, Germany
2
Department of Chemistry and Pharmacy, Chair of Aroma and Smell Research, Friedrich-Alexander-Universität Erlangen-Nürnberg, Henkestraße 9, 91054 Erlangen, Germany
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(18), 5514; https://doi.org/10.3390/molecules26185514
Submission received: 6 August 2021 / Revised: 7 September 2021 / Accepted: 8 September 2021 / Published: 11 September 2021

Abstract

:
The detection of chemical compounds in exhaled human breath presents an opportunity to determine physiological state, diagnose disease or assess environmental exposure. Recent advancements in metabolomics research have led to improved capabilities to explore human metabolic profiles in breath. Despite some notable challenges in sampling and analysis, exhaled breath represents a desirable medium for metabolomics applications, foremost due to its non-invasive, convenient and practically limitless availability. Several breath-based tests that target either endogenous or exogenous gas-phase compounds are currently established and are in practical and/or clinical use. This review outlines the concept of breath analysis in the context of these unique tests and their applications. The respective breath biomarkers targeted in each test are discussed in relation to their physiological production in the human body and the development and implementation of the associated tests. The paper concludes with a brief insight into prospective tests and an outlook of the future direction of breath research.

1. Introduction

The human metabolome represents the entirety of low molecular weight molecules present in the human body for a given physiological state and environmental conditions [1]. Metabolomics research undertakes to characterise these metabolic profiles that manifest in relation to health, disease, or environmental burden, with the aim of their prospective use in diagnostic applications or for drug-based interventions [2]. The field has evolved rapidly in recent years, driven foremost by technological advancements in analytical instrumentation (higher resolution/sensitivity, lower detection limits) and the development of sophisticated data treatment tools (multivariate chemometric methods).
Various biological sample types can be used to extract metabolic profiles, with their collection proceeding either invasively, e.g., blood serum or tissue biopsies, or non-invasively, for instance faeces, urine, sputum or breath. Invasive sampling is intrinsically associated with an element of discomfort (that can lead to low patient compliance), limited sample medium and collection frequency (e.g., blood is not exhaustive and can be sampled only intermittently) and typically higher costs. Non-invasive sampling overcomes many of these hurdles, although stool, urine and sputum are not limitless in supply and patient uptake can be compromised due to embarrassment for the former two, and discomfort for the latter. With these considerations in mind, exhaled breath represents an ideal biological fluid for metabolomics research, offering a practically unlimited supply and little to no discomfort to the patient, which encourages cooperation. In addition, exhaled breath can be sampled without the need for privacy or medical personnel and it typically does not generate infectious waste (notwithstanding airborne pathogens [3]), making breath analysis an attractive approach for various applications, from disease detection to exposure assessments [4,5,6,7].
Despite these benefits, the apparent simplicity of breath analysis conceals a complex enterprise. Breath is a rich medium comprising gas-phase inorganic and organic compounds, as well as aerosols in the form of water vapour and particles. Focussing on the gas phase, breath contains several inorganic species and several hundred volatile organic compounds (VOCs) of a diverse chemical nature, the latter being present only in trace quantities [8]. Accordingly, a major challenge in breath analysis is the sensitive detection of individual compounds and the determination of their unique specificity to the disease under scrutiny. The principle reasons for the hitherto moderate pace in progress of breath biomarker discoveries lies in the technological limitations associated with reliably capturing breath and the analytical intricacy of extracting potential biomarkers from complex datasets [9,10]. Moreover, the lack of standardisation in breath analysis has led to a limited alignment of results between independent studies employing different approaches [11]. The need to develop and introduce standardised practices in breath research is widely accepted, and adopting such measures is expected to expedite progress in the field [11,12,13], as has been demonstrated in the successful development and implementation of the nitric oxide breath test for asthma, as discussed below [14].
Overall, despite the aforementioned limitations and challenges, breath-based metabolomics has been successfully implemented for specific applications and shows promise as a complementary approach to traditional diagnostic tests for characterising health disorders, infections and environmental exposure [15]. This review outlines the case for using breath analysis as a rapid and practical tool in metabolomics research and presents and discusses the gas-phase breath-based tests that have been established and are in (routine) use. The paper commences with a general overview of relevant tests, then divides these topically between tests using compounds of either endogenous or exogenous origin, such as exhaled nitric oxide or breath ethanol, respectively. Each breath test is reviewed in relation to the discovery of the respective biomarker, the associated biochemical or physiological processes of biomarker production and/or metabolism in the human body and its subsequent excretion via breath, and ultimately the practical or clinical implementation of the test. An in-depth discussion of the benefits and limitations of breath analysis and a critical discourse of the different sampling and analytical approaches have been reviewed at length in the scientific literature [4,10,16,17,18]; thus, they are not treated here. Similarly, this review focusses on the compounds present in the gas phase of exhaled breath; the tests based on exhaled breath condensate (EBC), aerosols (EBA) or particles are not covered in this paper.

2. Approved and/or Established Breath-Based Tests

The volatile (gas-phase) fraction of breath comprises compounds of different origins, which are broadly categorised as endogenous or exogenous. Endogenous compounds derive from the host metabolism, particularly from routine metabolic processes, but also from an imbalance in the body relating to disease or (organ) dysfunction [19]. These compounds reflect the momentary physiological state of an individual and are of potential interest and use as biomarkers [20]. The regular metabolism in a healthy individual drives the production, distribution and fate of endogenous compounds. When adverse changes in the body up- or downregulate these processes, the associated changes in compounds prospectively manifest in exhaled breath [21,22]. In contrast, exogenous VOCs represent compounds from the environment. These are ubiquitous, and their intake into the body proceeds either through the inhalation of ambient constituents and pollutants, via the ingestion of food or drugs, or by dermal absorption [8,17]. A large proportion of the VOCs hitherto detected in exhaled breath are—or derive from—exogenous compounds [8,9,23]. An additional source of origin of VOCs, referred to as biological/non-host [17], is represented by the microbiome [24,25,26]. The interaction of the host and commensal microorganisms, especially in the human gut microflora, contributes significant complexity to the human metabolome [27], whereas the metabolism of the lung (and oral/nasal) microbiome may also affect the exhaled VOC profile [28]. The VOCs metabolised or produced in vivo by bacteria are of clinical interest and potentially provide indications of pathogen infections [24]. The symbiotic nature of the microbiome raises the question as to whether these bacteria should be considered a part of the human organism or rather treated as separate entities; accordingly, there is no current consensus on how the compounds associated with these organisms should be designated, either as endogenous or as exogenous, hence the additional ‘biological’ categorisation.
From the considerations above it is clearly evident that the origin of any individual compound present in breath is rarely distinctive. Human breath composition has manifold contributing factors; therefore, determining the specific origins of individual biomarkers is challenging due to multiple potential sources [20,24]. Despite these difficulties, breath research has made steady advances in health-related applications in its relatively short recent history. Several breath-based tests have met with medical or regulatory approval, such as through the United States Food and Drug Administration (US FDA) or the European Union European Medicines Association (EU EMA), or are endorsed and/or recommended by professional associations and are now in regular or routine use [29]. These approved and established tests, which are discussed individually in the ensuing sections of this review, are summarised in Table 1.

3. Tests Targeting Endogenous Compounds

The endogenous compounds present in breath derive from regular internal metabolic production, with either subsequent systemic circulation and transition into the gas-phase after passing the alveolar-blood capillary membranes in the lungs, or direct release into exhaled gas in the case of localised airways production [19,30]. A variation in the volatile composition of exhaled breath can provide insights into the corresponding (adverse) changes in the body [20]. As such, the majority of breath research hitherto undertaken has focused on exploring endogenous disease biomarkers.
The applications of endogenous compound-based breath tests that are currently implemented range from widespread routine use to highly specific settings. The most common test uses carbon dioxide, a by-product of cellular respiration, as a breath-borne marker to monitor breathing (e.g., in intensive care or sedated patients), whereas other tests target specific compounds for particular cases, e.g., nitric oxide in patients with asthma, or carbon monoxide in infants at risk of neonatal jaundice, demonstrating the prospects of breath tests based on a symptomatic approach. The Heartsbreath test for heart allograft rejection further shows the applicability of breath testing of multiple compounds as a screening procedure for adverse events. This section reviews the breath tests that utilise endogenous compounds, starting with the most common and widely implemented procedure.

3.1. Capnography

The practice of measuring carbon dioxide (CO2) during respiration is known as capnography. Specifically, capnography refers to the continuous analysis of CO2 partial pressure in respiratory gas [31]. The widespread, routine implementation of capnography throughout the world makes CO2 the most widely exploited and commonly used breath biomarker. Research on CO2 in exhaled breath boasts an early history, with its presence in breath first discovered in 1784 by Lavoisier and Laplace through their investigations of breath from guinea pigs. Antoine Laurent Lavoisier (1743–1794) and Pierre Simon Laplace (1749–1827) devised a system that enabled pre-concentrated breath to be purged through a device containing a chemical solution that reacted with CO2 in the gas to form a precipitate and indicate its presence. Accordingly, these experiments allowed Lavoisier and Laplace to demonstrate for the first time that CO2 is a constituent of exhaled breath, which they associated with the respiratory metabolism (Figure 1) [32].
During the same period, researchers began to study the association between CO2 and plant life, which ultimately led to the development of the theory of photosynthesis in the 1770s by the physiologist Jan Ingen-Housz (1730–1799) [33]. Photosynthesis can be essentially considered the reverse of respiration that acts as a mechanism for plants to store energy, with an uptake of CO2 and the release of oxygen (O2) following a series of reactions. In 1803, John Dalton (1766–1844) posited that the molecular structure of CO2 consisted of one carbon atom and two oxygen atoms. Due to the asymmetric and polyatomic nature of CO2, it strongly absorbs light with wavelengths in the infrared (IR) spectrum, a phenomenon that was exploited in one of the earliest IR measurements of CO2 in breath when John Tyndall (1820–1893) constructed an apparatus that measured the absorption of various gases and vapours, including CO2 [34].
The physiological production of CO2 is now well understood. The process of aerobic cell respiration involves a series of metabolic reactions that breakdown glucose to CO2 and water (H2O) and generate the energy-intermediate adenosine triphosphate (ATP). In the presence of O2, pyruvate, which is an intermediate product of glycolysis, is oxidised and enters the Krebs cycle (citric acid cycle) as acetyl coenzyme A (acetyl-CoA), whereupon metabolic transformations lead to the release of CO2 as a by-product (Figure 1, right-hand side) [35], which is then transported via the bloodstream into the lungs and is ultimately expelled from the body in exhaled breath (Figure 1, left-hand side).
The clinical exploitation of exhaled CO2 in the form of capnography has been widely implemented for over three decades, being a staple procedure in clinical practice since the late 1980s [36,37]. Besides its basic assessment of lung ventilation, capnography has become an important component during general anaesthesia, conscious sedation, intubation, patient movement or transportation and other procedures, providing an essential indicator for patient safety and offering valuable physiological data on ventilation and perfusion matching in the lungs, cardiac output and metabolic rate [38,39]. A schematic figure of the characteristic CO2 partial pressure in breath during respiration, referred to as a capnogram, is depicted in Figure 2a.
Generally, instrumentation for capnography operates based on one of two technologies, either colorimetry or infrared sensors. Colorimetric CO2 devices utilise a pH-sensitive material that changes colour in relation to the amount of CO2 present, thereby offering a qualitative and semi-quantitative detection of CO2. Colorimetric detectors are portable and disposable, yet are prone to false positive readings through pH-influencing contaminants [40,41]. Accordingly, the use of IR sensor-based technologies is the preferred choice for both intubated and spontaneously breathing patients. As a spectroscopic method, IR-based systems detect the presence of CO2 through absorption bands in the electromagnetic spectrum and the CO2 waveform is subsequently displayed as a function of time or exhaled volume [31,38]. The compact nature of IR sensors has led to the development of palm-sized miniature CO2 monitors [42,43], making bedside observation practical, as well as being particularly tolerable for vulnerable patient groups, such as children or the elderly [44]. IR sensors are sensitive and accurate, although they are susceptible to damage when handled or through fouling due to expectorate secretions and circuit condensate [31]. Monitoring the CO2 in breath can be performed via mainstream or sidestream sampling. The former relates to the direct analysis of the entire breath fraction through the inline positioning of the capnograph, whereas the latter samples respiratory gas through a side port of the breathing circuit (mainly for intubated patients), and as a consequence, incurs a time lag of several seconds before displaying the results [31,45].
Despite its benefits and widespread use, capnography is not universally implemented in medical settings owing to a limited availability and accessibility of equipment in some areas, disparate views of medical personnel on its ability to impact or improve patient care, and the contentious interpretation of capnograms. Accordingly, exhaled CO2 monitoring is often underutilised, despite its high potential in ensuring patient safety [36,46,47,48]. Nevertheless, the importance and acceptance of capnography in the clinical setting is evident by the numerous standards, guidelines and advisory statements, as well as its recognition by several professional institutions and regulatory agencies, including the American Society for Anesthesia, the American Heart Association, the American Association for Respiratory Care, the Joint Commission and the Centers for Medicare and Medicaid Services, and the Association of Anaesthetists of Great Britain and Ireland, amongst others [37,49]. Based on the evidence of its efficacy in medical care and the overwhelming consensus of medical practitioners, capnography represents an important procedure to monitor patient stability, thus exhaled CO2 can be considered a key and widely exploited breath-borne biomarker in healthcare.

3.2. Nitric Oxide Breath Test for Asthma

Nitric oxide (NO) in exhaled breath has been successfully exploited as a biomarker for asthma screening and treatment management since the turn of the millennium. Following discrepancies in the observed concentrations of exhaled NO between early studies, it soon became apparent that the fraction of exhaled nitric oxide (FENO) exhibits a flow dependency [52,53], thus must be measured under defined and well-controlled conditions to generate reliable data to indicate inflammation of the airways [54]. These findings led to the development of comprehensive guidelines for accurate FENO detection that were indispensable for establishing a reliable breath test for asthma.
The history of utilising FENO measurements in healthcare dates back approximately two decades, yet the history of NO research extends back further. NO is an inorganic molecule that was first discovered in 1774 by the British chemist—and contemporary of Lavoisier—Joseph Priestley (1733–1804) [55]. Early research on NO in the last century focussed on its role in atmospheric chemistry, notably in the production of tropospheric ozone and its contribution to photochemical smog, and two centuries were to transpire since its discovery before an association to human physiology was made. In the late 1970s and early 1980s, independent physiological and pharmacological studies by Ferid Murad, Robert Furchgott and Louis Ignarro on vasodilation, mechanisms and modes of action of vasodilating and muscle relaxing drugs, and the component termed ‘endothelial-derived relaxing factor’ (EDRF), led to the discovery of NO as an endogenous mediator in health and disease (this work was later acknowledged with the Nobel Prize in Physiology or Medicine for the ‘discoveries concerning nitric oxide as a signalling molecule in the cardiovascular system’, awarded to the trio in 1998) [56,57,58]. The presence of NO in mammalian exhaled breath was first reported in 1991 in a seminal paper by Lars Gustafsson and colleagues, who demonstrated its endogenous production by NO synthase through the use of enzyme inhibitors [59,60]. This discovery was soon followed by evidence that the NO in exhaled breath was present at higher concentrations in asthmatics [61,62,63], which opened the path for the use of NO as a breath-borne biomarker. Another notable discovery was made in 1995, namely that epithelial cells in the paranasal sinuses produced NO at high concentrations [64], highlighting a potential confounding contribution to FENO from nasal air.
Nitric oxide is produced in the body by a group of enzymes, nitric oxide synthases (NOS), when l-arginine is oxidised to l-citrulline in the presence of oxygen and cofactors (Figure 3a) [65,66,67]. NO has diverse functions in blood vessels and the airways, such as for smooth muscle relaxation and vasodilatation for matching regional airflow and blood flow [68]. In the respiratory system, NO is produced in the lung alveoli, proximal and upper airways, as well as in the nasal cavity, diffusing over cell membranes via a concentration gradient and released into the airways [69]. The elevated levels of NO in asthmatics are associated with inducible NOS (iNOS), which is a variant of NOS that is expressed when primed by inflammatory stimuli [70,71] but exhibits inhibitory effects after the administration of corticosteroids [65,72].
The observations in the early 1990s that airway inflammation associated with asthma leads to an increase in FENO, and that an inhibition of iNOS via inhaled corticosteroids (ICS) treatment acts to lower FENO levels, soon placed exhaled NO as a promising inflammatory marker of asthma [69,73,74]. Several discoveries soon followed, including observations of reduced FENO levels in smokers [63,75], age-dependent concentrations in children [76], and the influence of exogenous factors, such as caffeine ingestion, on FENO levels [77]. Notably, the FENO concentration was found to depend on exhalation flow, whereby levels were observed to decrease at higher flow rates (Figure 4), thus there was an immediate need to establish standardised practices to ensure accurate and reliable readings [14]. This phenomenon can be explained by the flow-dependency of NO output, where the exhaled gas at low flow will have time to be enriched with NO from the upper airways, whereas the alveolar gas constitutes a large part of FENO at high flow; the linear portions of NO output at low and high flow intercept at an expiratory flow of around 50 mL/s, regardless of the FENO concentration; this is designated with the term FENO50 (Figure 4) [78].
To address this aspect, the European Respiratory Society (ERS) created a taskforce to gather and provide consensus recommendations amongst experts, which was published in 1997 [80], followed shortly afterwards by similar guidelines compiled by the American Thoracic Society (ATS) [81]; a joint guidelines paper of the two societies containing revised recommendations for standardised methods of measuring and reporting exhaled NO was published in 2005 [82]. These guideline documents were decisive in aligning practices on NO measurements (including specifying an expiratory flow of 50 mL/s) and in generating an understanding of NO in the human physiology, and therefore, played a pivotal role in establishing the FENO breath test for asthma. The concurrent development of a chemiluminescence-based NO analyser specifically for FENO measurements in breath resulted in the launch of the commercial NIOX® device by Aerocrine AB (Solna, Sweden), which received an FDA approval for its use in detecting and monitoring asthma in 2003 [29,83].
Measurement scenarios for FENO differ depending on the application and setting. Apart from offline methods in which exhaled breath is first collected in a reservoir for subsequent analysis, as often implemented in large-scale studies [84,85,86], online NO measurements use an NO analyser for direct sampling and immediate analysis [87,88,89]. Online measurements may result in higher data quality, whereas offline testing can be more practical in some scenarios [90]. FENO sampling can be achieved either through continuous tidal breathing or single-exhalation episodes, with the latter generally being the favourable approach. Due to the localised generation of NO in the nose, sampling nasal air should be avoided [65,91]. The standardised single-breath technique specifies that a patient should inhale to total lung capacity (TLC), then exhale at a constant flow rate of 50 mL/s against resistance (to exclude nasal NO by means of velum closure); the 3-second NO plateau at the end of the exhalation is designated as FENO (Figure 2b) [82]. If inhalation to TLC is difficult to perform, e.g., in patients with severe lung illnesses, small children or the elderly, non-TLC deep inhalation may be performed, albeit for an extended period in order to reach the plateau [78]; this alternative manoeuvre is more comfortable for the patient and has been adopted by clinicians, as well as in instrument designs of manufacturers [92]. Ventilated patients pose another challenge for single-breath measurements, with studies describing alternative methodologies for intubated subjects [93].
Nowadays, the most common methods for NO measurement are chemiluminescence, electrochemical sensors and laser-based technology, with chemiluminescence representing the gold standard due to its high sensitivity, low detection threshold (ppb level) and fast response time (0.5–0.7 s) [94]. FENO values can be influenced by several non-disease-related factors, such as genetics, sex, weight and height, diet (e.g., caffeine intake), drug use or smoking, thus a questionnaire accompanying NO measurement is recommended to establish the potential confounders. Due to this variability, the Global Initiative for Asthma (GINA) recommends the use of additional parameters for asthma screening rather than a diagnosis based solely on FENO, although the guidelines nevertheless recognise that the use of FENO for treatment can lead to fewer exacerbations compared to treatment based on the current guidelines [94,95]. Overall, the clinical utility and diagnostic value of FENO for asthma management exhibits a sensitivity that ranges between 79% and 86% and specificities between 85% and 89% [96].
Presently, FENO is becoming increasingly established in the general clinical guidelines for asthma diagnosis. The British National Institute for Health and Care Excellence and the Scottish guidelines for asthma diagnosis and management, for instance, recommend FENO testing in combination with other diagnostic tests to help diagnose asthma or support asthma management in symptomatic patients despite ICS treatment [94]. Today, FENO is used to predict and monitor the ICS response [97,98] and adherence [99,100], and to diagnose ICS-naive patients [95]. Overall, the FENO breath test for asthma diagnosis demonstrates the high potential for breath-based diagnostics, with NO representing the most successful breath-borne biomarker for disease indication; it should be stressed again here that this success is directly attributable to the exhaustive studies on FENO in relation to sampling and analysis and the ensuing extensive guidelines on the related required practices.

3.3. Exhaled Carbon Monoxide in Neonatal Jaundice

Carbon monoxide (CO), an inorganic gaseous molecule primarily associated with the combustion of fossil fuels or organic matter, was first discovered in the late eighteenth century in the era of Lavoisier and Priestley. Indeed, it was Priestley who first produced CO in 1772 through heating charcoal, which he termed ‘combined fixed air’ [55]. The presence of CO in blood was first demonstrated in studies at the turn of the twentieth century [101,102], followed by observations in 1949 of its endogenous production, which was associated with bilirubin production [103]. The predominant source of endogenous CO is heme degradation by heme oxygenase (HO) enzymes (EC: 1.14.14.18) [104,105,106]. This enzymatic oxidation is catalysed primarily by heme oxygenase-1 (HO-1), the inducible form of HO, though the constitutively expressed isozyme HO-2 may also contribute [107]. Subsequently, CO binds with high affinity to haemoglobin to form carboxyhaemoglobin, which is transported through the systemic circulation to the lungs, where CO transitions into the gas-phase at the alveoli and is then excreted from the body via exhaled breath (Figure 3b) [108]. Besides CO, heme catabolism during red blood cell turnover in normal human physiology produces bilirubin [109,110]. The accumulation of bilirubin in neonates through elevated HO activity leads to hyperbilirubinemia, commonly known as jaundice, which causes a yellow-orange pigmentation of the skin, sclerae and other tissues [110]. Bilirubin production in normal term neonates is known to be two to three times higher than in adults [111]. Due to the neurotoxic nature of bilirubin, severe jaundice, caused by its accumulation to excessively high concentrations, is considered a pathophysiological condition, causing permanent neuronal damage, and is referred to as bilirubin-induced neurologic dysfunction (BIND), or kernicterus [112,113,114,115].
Bilirubin accumulation is exacerbated by haemolysis for a variety of reasons, such as the breakdown of extravascular blood, maternal diabetes or infant prematurity, as well as hitherto unknown causes associated with ethnicity, and other conditions [111]. Accordingly, current preventive or therapeutic measures for severe hyperbilirubinemia, such as phototherapy or blood exchange transfusions, are nonspecific [110]. The concomitant production of CO with bilirubin during heme degradation and its subsequent excretion via breath have been studied extensively. Due to its stoichiometry, the measurement of CO in breath can be used effectively as an index of bilirubin accumulation and thus, hyperbilirubinemia in vivo [109,116]. Specifically, it has been shown that the (pulmonary) excretion rate of CO (VeCO) represents a reliable measure to estimate bilirubin accumulation. Early studies of CO in the exhaled breath of neonates made use of an incubator, whereby the airtight hood was fitted with a control module and a gas chromatography (GC) system equipped with a gas detector that allowed the infant’s VeCO to be measured over a 20–30 min period after an equilibration period of 40 min [116]. In 1984, however, a study reported a correlation between VeCO and end-tidal CO (ETCO) in infants without pulmonary disease and, therefore, demonstrated the possibility to use ETCO instead of VeCO as a measure of bilirubin accumulation, offering significant advantages in simplicity and rapidity over the VeCO-based approach [117]. The measurement of ETCO as a marker of haemolysis was traditionally performed using GC [118], but this has transitioned to the contemporary use of more portable and non-invasive systems, such as the CoSense® CO monitor (Capnia, Inc., Redwood City, CA, USA), which exhibit sufficient accuracy and precision for the detection of haemolysis in neonates. Studies on ETCO have shown that concentrations ≥ 2.5 ppm (corrected for ambient CO) are indicative of the presence of significant haemolysis [119,120,121,122], and that this breath-based test exhibits a higher accuracy compared to the direct anti-globulin test (DAT) [123]. Moreover, a study from 2001 found a correlation between the ETCO level associated with neonatal jaundice, even in infants without haemolytic diseases, with high sensitivity, specificity and negative predictive values (86, 80 and 97%, respectively) [109]. Accordingly, this approach has been demonstrated to be an accurate method for the early identification of haemolysis in new-born individuals that offers the opportunity for the timely initiation of phototherapy, thereby reducing the risk of readmission, neuronal toxicity and kernicterus [124]. The efficacy of the exhaled CO breath test for neonatal jaundice in infants led to the procedure receiving an FDA waiver in 2018. Although not in widespread use, this breath-based test represents an important non-invasive approach for screening the most vulnerable members of society that allows for early action in treating this potentially fatal condition.

3.4. Heartsbreath Test in Cardiac Transplant Rejection

Cardiac transplantation is a life-saving treatment for patients with severe heart conditions and heart failure. Despite the scarcity of universal epidemiological data on heart failure, population-based studies have shown representative trends in recent years. The steady increase in admissions of patients suffering congestive cardiac failure has been linked to an ageing population and an increasing prevalence of obesity in younger adults, indicating the growing importance and necessity of transplantation in severe cases [125]. Following the surgical procedure itself, heart transplantation suffers from a high incidence of allograft rejection, despite progress in immunosuppressive therapy, and this represents a leading cause of death in these patient groups [126,127]. Post-surgery, recipients must undertake regular screenings for signs of organ rejection, but associated symptoms, including fatigue, malaise and oedema, amongst others, are uncommon, making rejection episodes difficult to detect [128]. Currently, an endomyocardial biopsy is the gold standard for the diagnosis of heart transplant rejection; this is an invasive procedure that is performed weekly for the first six weeks following allograft surgery, then periodically for at least one year post-surgery [129,130]. The biopsy procedure is itself associated with health risks, especially in immunosuppressed patients, following the original surgery, and oftentimes biopsy results are unremarkable and do not lead to an alteration in patient treatment, thus their benefit is questionable [126,128].
Non-invasive alternatives to screen for heart allograft rejection have been increasingly studied over the last few decades, including antibody tests, magnetic resonance imaging, echocardiography and the use of serum markers [131,132], yet these approaches pose deficiencies in their respective accuracy, making them unsuitable for personalised medicine. In the early 2000s, contemporary breath research pioneer Michael Phillips explored the potential to exploit the oxidative stress associated with organ transplantation and the ensuing manifestation of alkanes in exhaled breath [128,133]. Allograft rejection is accompanied by oxidative stress due to the increased production of reactive oxygen species (ROS), which degrade the membrane polyunsaturated fatty acids (PUFAs) caused by free-radical induced lipid peroxidation and ultimately lead to the generation of alkanes and methylalkanes, with their subsequent transition into breath [83,128,134]. Based on this premise, a model was developed that utilises a combination of nine C4-C20 alkanes and mono-methylated alkanes—termed the breath methylated alkane contour (BMAC)—as markers of oxidative stress [135]. The breath test is accomplished by capturing exhaled VOCs onto a sorbent trap using a portable breath collection apparatus (BCA, Breath Meter Technology, Inc., Cleveland, OH, USA) and subsequently analysing the sample through thermal desorption and gas chromatography-mass spectrometry (TD-GC-MS) [136]. The studies observed an increase in the BMAC in heart transplant patients with grade 0, 1 or 2 rejection, which is attributed to increased myocardial oxidative stress in the heart in the period between the post-mortem removal from the donor and the subsequent allograft in the recipient [137]. In contrast, a paradoxical contour reversal of the BMAC markers observed in patients with grade 3 rejection was observed, indicating that the abundance of these markers decreased in this group. It is hypothesised that this phenomenon is associated with self-induced catabolic processes, in which increased levels of alkanes trigger higher activity in the inducible cytochrome P450 (CYP450) enzymes that deplete the oxidative stress markers in these patients [128]. Overall, this breath-based test for heart transplant rejection—named Heartsbreath—is able to predict grade 3 organ rejection with high accuracy compared to the concordant set of International Society for Heart and Lung Transplantation (ISHLT) grades in biopsies (sensitivity 78.6%, specificity 62.4%, positive predictive value 5.6%, negative predictive value 97.2%) [128,133]. A particular benefit of the Hearthsbreath screening test is that the high negative predictive value allows rejection episodes to be largely ruled out, thereby reducing the need for burdensome biopsies. The Heartsbreath test, developed by Menssana Research (Newark, NJ, USA), received FDA approval through a humanitarian device exemption in 2004, which caters for devices intended to benefit individuals with diseases or conditions affecting fewer than 8000 patients annually in the US [138]. Despite FDA approval, however, the test is currently not in use because it is not covered in the public or private medical insurance programmes in the USA.

4. Tests Exploiting Exogenous Compounds

The exogenous compounds in exhaled breath represent substances that originate exterior to—and are taken up by—the body, which are subsequently eliminated via respiration. The primary sources include constituents of ambient air that are inhaled and food-derived compounds or medication ingredients that are ingested or administered. In breath analysis, xenobiotic compounds, i.e., chemical compounds that are foreign to the human organism, have been largely considered as confounding factors in the search for endogenous biomarker compounds [139]. Indeed, a persistent challenge in breath biomarker discovery research has been how to identify these compounds and exclude them from datasets [11], although in certain scenarios these can be utilised to explore aspects relating to the human exposome [140]. In terms of the inhaled compounds, the collection of ambient air samples concurrent to breath sampling provides a baseline dataset of environment constituents that can be used retrospectively to assess potential confounding VOCs in exhaled breath [20,141,142]. Such measures are particularly important in some settings, such as the clinical environment in which the abundant and persistent use of detergents and disinfectants creates a high background of related compounds [143]. An important point to note, however, is that the fate of a VOC of exogenous origin varies according to multiple factors, including the nature of the compound (lipophilicity, volatility), the degree to which it can be metabolised, as well as the concentration and duration of exposure. Consequently, collecting and analysing an ambient air sample at the time of the breath test represents only a snapshot of the immediate exposure but will not throw light on prior exposure, nor will it identify the metabolic by-products of inhaled compounds. In view of the latter, however, this process can be exploited in a targeted way, as discussed below.
Although exogenous compounds present extraneous confounders in breath studies that target endogenous biomarkers, conversely they represent an opportunity for assessing environmental exposure and pharmacokinetics [139,144]. Observations of their reduced levels in breath post-uptake as a result of metabolism, for example, might be as equally insightful as finding raised levels [19]. Accordingly, exogenous VOCs from daily exposure or through active introduction to induce a perturbation can be used to assess specific (disease-associated) processes or to identify population subgroups [144,145]. Several breath tests exist that exploit exogenous compounds, and all of these make use of the ingestion route. One such breath test, which does not target disease per se, is the well-known and widely implemented breath alcohol ‘breathalyser’ test, as used in law enforcement to identify drink-drivers [146]. Other targeted tests focus on the analysis of metabolic by-products from administered substrates, including the breath tests for hypolactasia [147], Helicobacter pylori infection [148], gastroparesis [149] and liver function [150] (depicted in Figure 5a–d). This section presents a discourse on the current breath tests that exploit exogenous substances or substrates for various means, commencing with the most common approach, before discussing the more specialised niche procedures.

4.1. Breath Alcohol Testing in Law Enforcement

The (over-)consumption of alcohol impairs the cognitive and psychomotor functions of the body, which has adverse effects on the execution of skilled tasks, such as driving a motor vehicle. Due to its physicochemical properties and related pharmacokinetics, ethanol can be readily detected in exhaled breath after alcohol consumption. The concentrations of alcohol in the blood and breath are highly correlated, which is utilised by law enforcement worldwide to identify drivers suspected of driving under the influence (DUI) of alcohol. Next to capnography, the breath alcohol test in law enforcement—colloquially known as the breathalyser test—is the most well-established and widely implemented breath-borne biomarker test [146,151].
The first breath alcohol analysis dates back to the mid-1860s, when British physician Francis Anstie made the first quantitative measurement of ethanol exhaled in human breath [152]. Using a breath trap containing chromic acid, which shifts in hue from red-brown to green in the presence of ethanol, Anstie demonstrated that only a small fraction of the consumed alcohol could be recovered in exhaled air. This led to the conclusion that the majority of the alcohol is metabolised in the body.
After alcohol is ingested, it is absorbed from the gut and transported to the liver, where approximately 90% is metabolised by oxidation with alcohol dehydrogenase (ADH) enzymes. Studies have shown that a maximum of 10% of the dose ingested is ultimately excreted unaltered via breath, urine and/or perspiration (Figure 6) [153,154,155].
Extensive studies have shown that the breath alcohol concentration (BrAC) is highly correlated with the concentration of alcohol in arterial blood circulation (blood alcohol concentration, BAC) over a wide range of ethanol dosages [156,157]. The elimination rates from the body are similar in both media [158]; this relationship demonstrates that BrAC is a suitable proxy to determine BAC [159]. This gives both the scientific and physiological premise for using an ethanol breath test for alcohol intoxication in law enforcement.
Despite an excellent statistical correlation, however, there are certain limitations to the alcohol breath test in forensic practice, for example, samples should not be taken within at least 15 min after the end of drinking to avoid falsely overestimating BrAC (and, correspondingly, BAC) as a result of residual, unabsorbed ethanol present in the mouth [160,161]. Furthermore, a person’s breathing pattern, the temperature of the breath and the exhaled volume impact on the resulting BrAC [162,163]. The phase of ethanol metabolism also warrants consideration, as does specific health conditions, such as pulmonary functions [164,165]. These limitations with breath alcohol testing are well known and can be minimised to a large extent by following certain recommendations for quality assurance and, if necessary, the analysis of ethanol in alternative biological matrices, such as blood or urine, in some circumstances [160].
Diverse breath alcohol testing systems have been developed since the first devices emerged in the 1930s–1950s [166], transitioning from slow, manually operated wet-chemical instruments, incorporated in the Drunkometer [167] or the first Breathalyzer [168], to the more compact and easily operated infrared spectroscopy-based systems [146]. The current state-of-the-art breath alcohol instrumentation offers contact-free breath sampling (Servotek AB, Arlöv, Sweden) and exhibits an accuracy of >97% and a precision (coefficient of variation) of <1% [153,169,170]. Moreover, these contemporary instruments do not require a mouthpiece to be attached, because they simultaneously measure the exhaled water vapour concentrations, and the ethanol content is then standardised to the fully saturated water vapour concentration of alveolar air at body temperature [171]. These technical advances, together with empirical data on the relationship between BrAC and BAC, have led to widespread acceptance of breath alcohol analysis for legal purposes, with handheld devices for screening driver sobriety at the roadside, to evidential quality instruments, the results from which are used in criminal prosecutions [146,172,173]. Furthermore, also available are small and compact devices for use by the general population for self-control of sobriety before driving.
The implementation of BrAC screening for a suspected DUI—with subsequent confirmatory BAC analyses—dates back to the 1950s, with the development of the first commercial Breathalyzer instrument, which became accepted and used in Australia, Canada and the USA [153]. Three decades later and extensive research established a quantitative relationship between BrAC and BAC and the worldwide acceptance of breath alcohol test results for evidential purposes. Instead of converting BrAC into BAC, however, most countries adopted a statutory BrAC limit, above which it has been made an offence to drive [153,174].
The legal permissible BrAC limits vary between countries and currently range from 0.10 to 0.40 mg/L, which equates to between approximately 0.2 and 0.8 g/kg (0.2–0.8‰) BAC [153]. The approval of use for breath alcohol devices in law enforcement varies depending on the country; in the USA, approval is granted by the National Highway Traffic Safety Administration of the US Department of Transportation. In relation to FDA approval, the first approved device for breath ethanol testing was the AlcoMate CA2000 Digital Alcohol semiconductor oxide sensor detector (KHN Solutions LLC, San Francisco, CA, USA), issued in 2004, with various other devices being approved since then [29]. Today, breath alcohol tests are in widespread and frequent use worldwide. The ‘breathalyser’ test does not represent a disease diagnostic procedure, yet it demonstrates the high utility of a breath-based approach and its potential for practical applications beyond healthcare. The breath alcohol test is the most common breath-borne biomarker-based procedure that is encountered in everyday life.

4.2. Hydrogen Breath Test in Hypolactasia

Lactose intolerance, or hypolactasia, is a condition in which lactose cannot be readily digested due to insufficient lactase enzymes in the gut. Specifically, the malabsorption of this sugar is caused by a reduced expression of lactase in the small intestine, leading to patient discomfort with abdominal complaints, such as diarrhoea, flatulence or pain [175,176]. Several clinical tests exist to diagnose lactose malabsorption [177], including the measurement of serum glucose, genetic tests and breath testing [178,179,180]. In the past, a lactose activity assay by jejunal biopsy was proposed as the gold standard [175,181], but this has been superseded by the non-invasive hydrogen breath test (HBT), which is one of the few breath-based approaches that is applied for routine clinical purposes and the most widely used method for the determination of lactose malabsorption [182,183,184].
Hydrogen breath tests are based on the principle that hydrogen gas is not endogenously produced other than by intestinal bacteria as they metabolise carbohydrates, such as lactose, that are insufficiently absorbed in the colon; accordingly, the HBT can be considered a biological/non-host-driven test (see Section 2). Following its production by intestinal bacteria, intracolonic hydrogen (H2) diffuses through the colon wall into the systemic circulation and is carried to the lungs, where it is excreted from the body via exhaled breath (Figure 5a) [182,185,186,187].
As the ingested substrate (lactose) is considered a common ingredient of food consumed in large quantities on a daily basis, the HBT itself is exempt from approval by the FDA; although, the first HBT instrument, Micro H2, using sensor technology (Micro direct, Inc., ME, USA), received FDA approval in 1997, followed by other HBT testing devices in subsequent years [29]. Various techniques have been put forward to perform the HBT, which differ according to the type of substrate (lactose or milk), its quantity (physiological or tolerance test), and the duration of the test (one to six hours) and its sampling intervals [181,182,188,189,190]. Nowadays, the HBT involves an oral challenge with a standardised dose of lactose (usually 20–50 g, but up to 100 g, corresponding to a lactose content of approximately 400–2000 mL bovine milk), with subsequent monitoring of exhaled H2, usually over a period of about two hours; lactose intolerance is then determined when breath H2 remains at significantly elevated concentrations within the respective test period (cut-off value > 20 ppm) [147,175,180,183].
Although HBTs are in routine use in clinical practice, the interpretation of the test results can be challenging due to several factors that influence its accuracy. The premise of the HBT is the presence of hydrogen-producing bacteria, yet there is a considerable proportion of patients in whom the colonic flora does not produce hydrogen, referred to as ‘H2-non-producers’, which leads to false-negative results [180,186,191]. Another possible cause for false-negative results could be a full colonic adaptation to lactose ingestion due to the favoured growth of lactose metabolising bacteria without hydrogen production [192,193]. Usually, patients produce either H2 or methane gas (CH4); however, about 30% of the adult population possess methanogens that produce methane at the expense of hydrogen in the gut. In these cases, complementary CH4 detection in breath can enhance the HBT and, therefore, might improve the correct diagnosis of malabsorption issues [194,195]. Hydrogen-methane breath testing has been hitherto underutilised due to the lack of low-cost, easy-to-operate instruments, but recent technical advances have led to the emergence of commercial devices [194], such as the BreathTrackerTM SC analyser by QuinTron Instrument Company, Inc. (Milwaukee, WI, USA), which separates the compounds in an alveolar gas sample using a fast GC combined with a solid-state sensor and is used to identify carbohydrate malabsorption and small intestinal bacterial overgrowth (SIBO) sufferers [196]. Other devices target multiple components by incorporating electrochemical H2 sensors and IR sensors for CH4 and CO2 detection (Lactotest 202 by Medical Electrionic Construction, Nivelles, Belgium) [197], or IR sensors for the simultaneous measurement of H2, CH4 and O2, with additional electrochemical sensors for CH4 and H2 detection (GastroCH4ECK by Bedfont Scientific Ltd., Maidstone, UK) [195]. Methane breath testing is, therefore, an emerging and promising complement to the HBT. An additional potential source of error in the HBT is carbohydrate malabsorption in relation to chronic pancreatitis and coeliac disease [198,199]. Conversely, false positives can arise due to smoking, oral bacterial flora or SIBO, or from a high intake of dietary fibres on the day prior to testing [180,185,200]. Accordingly, tracking the nutritional history and symptoms of patients, as well as measuring the blood glucose levels and exhaled methane, are helpful in correctly interpreting HBT results [178,183].

4.3. 13C-Breath Tests in Clinical Applications

Elemental carbon naturally occurs in three isotopes, with 12C and 13C being stable. In the past, breath tests were developed using the radioactive isotope 14C for testing exocrine pancreatic function; however, its radiation hazard and long half-life (5730 years) led to the increasing use of 13C as a substitute for diagnostic tests [148,201]. In 1973, Lacroix and co-workers first reported the use of 13C as a tracer compound in a human metabolomics study, whereby the abundance of exhaled 13CO2 was followed after the ingestion of 13C-labelled glucose [202]. Nowadays, 13CO2 is used in several diagnostic tests as an in vivo biomarker in exhaled breath after ingesting specific 13C-substrates, pushing the way forward in personalised medicine [83]. In contrast to the HBT, the field of application of 13CO2 breath tests is wider due to the diversity in available substrates. The choice of the substrate determines the target of the respective breath test, from potential gastric bacterial Helicobacter pylori infection, gastric emptying, liver and pancreatic function, to the assessment of other enzyme activities. The common basis of these breath tests is the use of 13C-labelled tracer probes that undergo metabolism via a pathway of interest and produce 13CO2 as a metabolite, which is subsequently expelled through respiration and detectable in exhaled breath [203]. By monitoring the unidirectional decomposition to 13CO2, the turnover of the substrate can be assessed; this is based on the premise that the process under investigation is driving the excretion rate of 13CO2, since other metabolic processes are negligibly fast or not variable. It is worth noting that the tracer compound cannot be fully recovered in breath, since a part of it is stored in the carbon pool of the human body; thus, such approaches are only semi-quantitative [186]. Despite substantial research in 13CO2-based breath testing, only three tests are currently approved by the FDA/EMA, namely the 13C-urea breath test (UBT) for Helicobacter pylori infection, 13C-spirulina for gastroparesis, and 13C-methacetin for liver maximum (LiMax) function assessment [203], which are presented in the following sections.

4.3.1. 13C-Urea Breath Test for Diagnosis of Helicobacter Pylori Infection

The most prominent application of stable isotope substrates and subsequent 13CO2 monitoring is the 13C-urea breath test (UBT), which detects a Helicobacter pylori (H. pylori) infection in the gut. H. pylori is a common bacterium infecting at least 50% of the world’s population [204,205]. The microbe produces large amounts of the enzyme urease, which hydrolyses urea in the stomach [148]. Even though the majority of infected people are asymptomatic, an H. pylori infection has been associated with different diseases, such as peptic ulcer disease, non-ulcer dyspepsia and gastric cancer [206,207,208]. The close association of an H. pylori infection with gastric cancer has led to the classification of this bacterium by the World Health Organization (WHO) as a group 1 carcinogen [209,210]. The assessment of possible infection is, therefore, crucial to enable suitable antibiotic therapy.
The UBT has undergone several evolutions since its first reported use in 1987 [211], namely adjustments in the test protocol relating to the fasting state, the type and quantity of the substrate and the sampling intervals [212,213,214,215]. Generally in this test, orally administered 13C-urea is hydrolysed by the bacterial urease activity in the stomach to form 13C-labelled CO2 (and ammonia, NH3), which is absorbed through the mucus layer of the stomach, transported to the lungs with the bloodstream and excreted via exhaled breath (Figure 5b) [148]. After an initial baseline breath sample collection, a patient will typically ingest 75 mg of 13C-urea and will provide a second breath sample 20 min later. An enrichment in exhaled 13CO2 of >2.4‰, or about 26 ppm, as determined with isotope ratio mass spectrometry, is evidence of an active infection, which is expressed as delta over baseline (DOB) [203].
Various factors suppressing bacterial growth can affect the test results, including the intake of proton pump inhibitors, H2 antagonists or antibiotics, which can reduce sensitivity and might cause false negative results [186,216]. Furthermore, it is common practice for patients to receive a standardised test meal to ensure a consistent diagnostic accuracy of the UBT for pre- and post-treatment for an H. pylori infection. Multiple studies and a meta-analysis that included more than 3500 patients reported the high sensitivity (>95%) and specificity (>95%) of the UBT compared to the histology [212,217,218,219]. Its accuracy has been compared to the invasive diagnostic rapid urease test, histology and culture after a gastric biopsy during an endoscopy [220] and the non-invasive faecal antigen test [208], and it has been claimed to be superior to the previous methods [186].
In the past, an H. pylori infection was diagnosed by a non-invasive stool antigen test and serology, or via invasive means, such as histology, rapid urease test, and culture as biopsy-based endoscopic tests [201,221]. Since its approval by the FDA in 1996, the non-invasive UBT has been instrumental in reducing endoscopic procedures (and the associated costs) in clinical screenings. A commercial UBT kit for the diagnosis of an H. pylori infection was launched on the market in 1997 by Meretek Diagnostics Inc. (Lafayette, CO, USA), followed by the approval of additional UBT devices, such as the Exalenz BreathID Hp System (Exalenz Biosciences Ltd., Modi’in, Israel) in 2011, or more recently, the successful premarket approval in 2020 for the PyloPlus UBT system (ARJ Medical, Oldsmar, FL, USA) [221].

4.3.2. Gastric Emptying Breath Test for Gastroparesis

Gastroparesis is a muscular disorder that delays the gastric emptying of food, commonly resulting in symptoms of early satiety, postprandial fullness, nausea, vomiting, belching and bloating. Since the condition is part of a catalogue of gastric neuromuscular dysfunctions with overlapping symptoms, it is essential to be able to distinguish between them [222]. In 1833, Beaumont first observed the emptying of gastric contents through a gastric fistula [223]. Since then, various methods have been developed to gauge the degree of gastric emptying, but no single technique has the ability to describe comprehensively the gastrointestinal transit after the ingestion of a heterogeneous meal comprising solid and liquid components [224]. This shortcoming relates to the distinctive emptying between solids and liquids, smaller and larger solid particles, and the lipid and aqueous phases of gastric contents [225]. Scintigraphy currently represents the gold standard reference method for determining the gastric emptying of solid and liquid meals [226,227]. The procedure, which must be executed by highly qualified staff, involves the use of a test meal that incorporates potentially harmful radioactive isotopes; thus, alternative methods are highly desirable [224,228,229]. The gastric emptying breath test (GEBT) is a comparatively safe technique to assess digestive function and provides a reliable diagnosis of gastroparesis [149]. In contrast to scintigraphy, the GEBT makes use of 13C-substrates, either 13C-octanoic acid [230] or 13C-enriched Spirulina platensis (S. platensis) (an edible blue-green alga) for solid gastric emptying, or 13C-sodium acetate for liquid gastric emptying [231], all of which are suitable and safe for critically ill patients, pregnant women and children [201,221]. These substrates are quickly absorbed in the proximal small intestines and are subsequently metabolised by the liver, leading to the production and ultimately excretion of 13CO2, which is eliminated from the body via respiration (Figure 5c). The time limiting step from substrate ingestion to the appearance of elevated 13CO2 concentrations in the breath provides a measure for the rate of gastric emptying [232], although the detection of alterations in the gastric emptying of solids is generally more sensitive than for liquids [233].
The GEBT procedure requires three to four hours, whereby breath samples are collected once pre-ingestion and at five to six intervals post-ingestion of the test meal labelled with the 13C substrate, such as 13C-spirulina [221]. A post-meal enrichment of 13C in relation to 12C is determined in exhaled CO2 using isotope ratio mass spectrometry, which allows the rate change of 13C in the breath (in µmol/L/min) to be calculated using the DOB approach [234]. A suppression of gastric emptying is indicated through a delayed increase in 13CO2 and a subsequent slower recovery rate of 12CO2, and measures of these phenomena are then used to calculate the gastric emptying half-time and lag phase duration [228], which reflect the overall emptying functions and the ability of the stomach to triturate solid food to smaller particles that can be emptied [235].
Despite the aforementioned benefits, the GEBT is not widely implemented as an alternative to scintigraphy, largely due to a lack of awareness of this approach by physicians and patients. Nevertheless, validation studies have demonstrated that the GEBT has a specificity and sensitivity of 89–98% and 24–64%, respectively, with positive predictive values of 73–97% and negative predictive values of 65–90%, depending on the chosen time point for diagnosis [236]. In 1981, the FDA acknowledged 13C-S. platensis as a ‘legally marketed’ food with health benefits due to its high nutritional qualities, thereby bypassing the need for specific approval for the adminstration of the substrate. The GEBT for solids using this algae, developed by Cairn Diagnostics (Brentwood, TN, USA), received FDA approval in 2015 [221]. The GEBT is a validated and standardised procedure that has been demonstrated to be highly reliable compared to the gold standard method, thus representing a suitable, non-invasive alternative to the hitherto established diagnostic approach for gastroparesis.

4.3.3. Maximum Liver Function Capacity Breath Test

Hepatologists have been studying potential 13C-breath tests to examine the cytosolic, mitochondrial and microsomal hepatic function associated with various liver diseases for over four decades [203,237]. In contemporary medicine, hepatectomy represents the treatment of choice for liver malignancies, yet the procedure carries a significant risk of post-operative liver failure—and death—following the resection of the hepatic tissue, a risk that could be lowered through knowledge on the pre- and post-surgery liver function capacity [238,239]. Various tests to evaluate liver dysfunction exist, including methods based on protein synthesis in the liver (e.g., prothrombin, albumin), hepatocellular integrity (transaminases), cholestasis and excretion (bilirubin, alkaline phosphatase, gamma-glutamyltransferase), as well as other techniques [240]. In addition, scoring indices for assessing liver damage severity and monitoring patients, such as the model for end-stage liver disease (MELD) and Child–Turcotte–Pugh (CTP), are widely used. A major shortcoming of these tests, however, is that they only provide an assessment of liver dysfunction (or injury) rather than function. Due to the capacity of the liver to regrow and rejuvenate, it is critical for haematologists to evaluate function rather than injury or dysfunction, since even a compromised liver can exhibit sufficient functionality. Further, on the one hand, there is a need to evaluate the liver function of patients undergoing a hepatic resection in order to identify eligible candidates for liver transplantation [241], and on the other hand, the early detection of post-operative liver failure is essential [239,242].
Breath-based tests that utilise 13C-labelled substances have been examined for different applications, from evaluating the residual liver functional capacity or function after liver transplantation, to assessing the severity of liver fibrosis from early stages up to liver cirrhosis [201,243]. A variety of 13C-labelled substrates has been studied for their use in assessing hepatic function, including 13C-phenylalanine, 13C-aminopyrine, 13C-erythromycin and other compounds [201,244,245,246,247]. One specific substrate is the paracetamol prodrug 13C-methacetin, which targets the CYP450-dependent enzymatic system and is used to determine maximum liver capacity, or LiMAx [150,221]. 13C-Methacetin is exclusively metabolised by CYP450 1A2 (CYP1A2), whereby the cleaved methyl group is oxidised to formic acid, before entering the C1 pool and subsequently exhaled as 13CO2 (Figure 5d) [186,248]. The enzyme is not influenced by genetic variations or drugs and is ubiquitous in the liver, making it the ideal target to assess liver function capacity [239].
As with the 13C-based tests discussed above, the LiMAx breath test involves determining the patient’s DOB via an initial measurement of the 13CO2/12CO2 ratio. 13C-Methacetin is then administered intravenously at a body-weight-adjusted quantity and the 13CO2/12CO2 ratio in the breath—which changes as a function of liver capacity—is monitored in real-time over the course of an hour [248]. Liver function is classified into three levels based on the metabolising capacity of CYP1A2, as determined from the latter ratio, namely normal liver function, with a lower cut-off of 315 µg/kg/h, intermediate liver function, with 140–315 µg/kg/h, and strongly impaired hepatic function, with a LiMAx < 140 µg/kg/h. The LiMAx test has been demonstrated to be highly reproducible in subjects with normal liver function, with a correlation coefficient of the repeat LiMAx test of 0.85 (95% confidence interval 0.69–0.93) [239]; further, age, sex and obesity have not been observed to influence the test. Due to the invasive (intravenous) nature of the substrate administration, the LiMAx test has yet to receive FDA approval and its implementation in the US market is challenging, yet an EMA approval of the test in 2017/2018 for its clinical use in Germany, Austria and the United Kingdom [221] has led to its current practice in clinical diagnostics in more than 20 hospitals in Europe [248].

5. Summary and Outlook

Metabolomics is an emerging discipline that offers a promising opportunity to examine the human physiology in relation to different scenarios, from screening for disease, to targeting drug metabolism, or estimating the burden of environmental stressors; it thereby serves as a potential tool for objective diagnostics. In comparison to other metabolic profiling approaches, the analysis of human exhaled breath represents a dynamic and multifaceted technique with high prospective value. The benefits of breath-based tests are indisputable: breath can be provided non-invasively, on demand and repeatedly, offering patient comfort, flexibility and avoiding the specific need for privacy or highly skilled personnel; the analysis of certain gas-phase biomarkers can be performed directly (online) and deliver immediate results; further, aside from airborne pathogens, exhaled breath is typically non-infectious and does not generate hazardous waste. Despite these benefits, breath analysis has its limitations: most gas-phase volatile chemical compounds are present in the breath at ultra-trace concentrations and exhibit a degree of lability, making many compounds challenging to sample efficiently and analyse accurately; their high qualitative and quantitative variability, their commonality in different diseases, and their ubiquity in the environment present further challenges in identifying representative, disease-specific biomarkers.
Nevertheless, several breath-based tests have emerged to become established in clinical practice or other settings. These tests vary in their approach in targeting compounds that are either of endogenous or exogenous origin. Beyond the breath-based tests reviewed here, current breath research is exploring several promising avenues. Notably, breath analysis has received unprecedented attention recently in relation to the severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2) outbreak and the associated coronavirus disease 2019 (COVID-19) pandemic. The potential for exhaled breath to either detect this airborne virus directly or to diagnose infection is currently being investigated as a comfortable alternative to existing approaches that collect mucus secretions via nasopharyngeal or oropharyngeal swabs, or serological samples [249,250,251]. While no breath test has yet been developed that allows a reliable detection of the infection, studies have reported potential breath-borne VOC biomarkers (detected via gas chromatography ion mobility spectrometry, GC-IMS) [252] or specific breathprints (using proton transfer reaction time-of-flight mass spectrometry, PTR-TOF-MS) [253] for COVID-19, as well as evidence for tests using nanomaterial-based sensors [254]. Further, although not treated in this review, exhaled breath condensate (EBC) has been explored as an alternative medium to mucus collected by nasopharyngeal swabs to detect the presence of COVID-19 via reverse transcriptase-polymerase chain reaction (RT-PCR) [255]. There is much research currently being undertaken on detecting COVID-19 in exhaled breath, and the data collected from early pilot studies suggest that a non-invasive breath test is a viable future approach to screening for infection.
Besides COVID-19, breath testing has a promising future in other areas of application. One particularly innovative approach relates to the potential diagnosis of liver cirrhosis in relation to exhaled limonene. Limonene in the breath is widely accepted to be of exogenous origin, deriving primarily from the diet (e.g., being present at high concentrations in citrus fruits and juices), or to a lesser extent through inhalation due to its ubiquitous presence in the environment. In a 2015 study that investigated the exhaled breath composition of patients with a cirrhotic liver, it was observed that concentrations of limonene (as well as methanol and 2-pentanone) in the patient group were significantly elevated compared to healthy controls [256]. These observations were explained by the inability of the compromised liver to breakdown this exogenous compound, compared with its efficient metabolism by CYP450 enzymes in the healthy cohort. Follow-up studies have further explored this phenomenon, with the inclusion of patient groups with hepatocellular carcinoma (HCC) and/or hepatic encephalography (HE) [257,258]. Similar observations were made, although no differences were found between liver cirrhosis with or without HCC, and the degree of HE severity could not be linked to limonene concentrations, although limonene accumulation was speculated to be itself a causative pathway for HE. The innovative aspect of these first pilot studies is that they exploit a dietary/environmental constituent that is not expressly administered to represent a marker of organ function, which is a hitherto unexplored area of breath research. Further, these studies serve as the basis for the development of tests using defined substrates—exogenous VOC (EVOC) probes—to examine the metabolic breakdown of a target compound to ascertain compromised enzymatic activity. This approach is currently the subject of investigation [145].
Looking at the established and/or approved breath tests, as discussed in this review, three of the eight tests make use of 13C-labelled substrates, which underlines the potential of such targeted approaches in breath analysis. Indeed, it can be expected that different 13C-labelled tracer probes will join the collection of breath tests for clinical routine purpose in future. The current developments that focus on the analysis of 13CO2 include a test that utilises an oral administration of 13C-methacetin as a substrate as opposed to intravenous administration (see Section 4.3.3; Figure 5d) to assess metabolism and liver capacity, a breath-based glucose tolerance test using 13C-labelled glucose, and a substrate-based bacterial overgrowth assessment, as well as a pancreatic inefficiency test, as reviewed in the literature [221]. Further, numerous studies have observed associations between breath volatiles and sputum or blood inflammatory cells, which highlight the prospects of breathomics as a future clinical tool for disease phenotyping and personalised medicine [259,260,261].
A key lesson learnt from the FENO test for asthma is that standardisation is essential to establish a reliable, reproducible and meaningful test. The lack of standardised practices in breath research has been a major impediment to progress in the field [11]. Accordingly, the current initiatives on standardisation and benchmarking are worthy of mention here. Specifically, the International Association of Breath Research (IABR) has instigated a pilot study to explore the establishment of a benchmarking protocol for the measurement of VOCs in the breath in order to allow for comparison between sampling and analysis approaches. The ‘Peppermint Experiment’ is a pharmacokinetic-based undertaking that focusses on the release and excretion of volatile peppermint oil constituents after the ingestion of a capsule containing the oil [262]. In short, measuring the washout curves of the specified compounds in exhaled breath over a defined timeframe allow for mean washout times to be calculated for any one sampling and/or analytical approach, allowing for comparison with other datasets. The consortium of participating research laboratories currently numbers 16 from seven countries, with the initial pilot studies comprising 1200 breath samples collected from 200 participants, and with analyses performed using GC-MS [263,264], secondary electrospray ionisation mass spectrometry (SESI-MS) [264,265], as well as PTR-MS and selected ion flow tube-mass spectrometry (SIFT-MS) [266], and with the publication of datasets from other approaches, such as GC-IMS, pending. Although the comprehensive datasets from these feasibility studies have highlighted the necessity to improve and refine the experimental protocol, the approach taken represents a first concerted effort within the broader breath analysis community to establish a method to allow for quality assurance checks of breath data. Ultimately, a compartmentalised approach to standardising practices in breath sampling and analysis is needed in order to cater for the broad spectrum of methods and target diseases [12].
The existing breath tests reviewed in this paper demonstrate the efficacy of breath-based diagnostics. Breath analysis remains an innovative and compelling approach for companion diagnostics and personalised medicine. Continual advancements in sampling approaches and analytical technologies, as well as data mining tools, will serve to generate cogent evidence for breath-based testing. The field holds much promise, especially for intervention-based approaches or personalised monitoring, but for new breath tests to succeed and transition to routine applications, the corresponding data must be cross-validated and exhibit high sensitivity, selectivity and accuracy, as well as offer a benefit over conventional diagnostic approaches.

Author Contributions

Conceptualisation J.B. and Y.L.P.; writing—original draft preparation, Y.L.P.; writing—review and editing, J.B.; visualisation, Y.L.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly funded by the German Ministry of Education and Research (BMBF), grant number 01KT1803A, as part of the ERA-NET TRANSCAN-2 project Airborne Biomarkers for Colorectal Cancer (ABC-Cancer).

Acknowledgments

We dedicate this review to—and congratulate—Bogusław Buszewski on the occasion of his 70th birthday. Buszewski has been a resounding presence in breath research for more than fifteen years. His contributions to the field include innumerable research and review articles on a wide range of topics, from novel separation and analytical approaches for examining breath gas matrices, to clinical trials for cancer diagnostics via exhaled breath, and even setting the analytical chemist’s best friend (GC-MS) aside to work with man’s best friend to explore the ability of canine olfaction to detect volatile disease biomarkers in biological specimen. His hosting and organisation of the 2014 IABR Breath Summit in Toruń, Poland, were a testament to his passion for interdisciplinary science, networking and debate, and, as is worthy of mention, demonstrated his adeptness at putting on a great show. We are indebted to Marieann Högman, Wayne Jones, Anil Modak, Michael Phillips and Steve Ryter for their critical appraisal and input of individual sections of this review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Snowden, S.; Dahlén, S.E.; Wheelock, C.E. Application of metabolomics approaches to the study of respiratory diseases. Bioanalysis 2012, 4, 2265–2290. [Google Scholar] [CrossRef] [Green Version]
  2. Dunn, W.B.; Ellis, D.I. Metabolomics: Current analytical platforms and methodologies. Trends Anal. Chem. 2005, 24, 285–294. [Google Scholar]
  3. Pleil, J.D.; Beauchamp, J.D.; Dweik, R.A.; Risby, T.H. Breath research in times of a global pandemic and beyond: The game changer. J. Breath Res. 2020, 14, 040202. [Google Scholar] [CrossRef]
  4. Beale, D.J.; Jones, O.A.; Karpe, A.V.; Dayalan, S.; Oh, D.Y.; Kouremenos, K.A.; Ahmed, W.; Palombo, E.A. A review of analytical techniques and their application in disease diagnosis in breathomics and salivaomics research. Int. J. Mol. Sci. 2016, 18, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Pleil, J.D. Breath biomarkers in toxicology. Arch. Toxicol. 2016, 90, 2669–2682. [Google Scholar] [CrossRef] [PubMed]
  6. Pleil, J.D.; Lindstrom, A.B. Exhaled human breath measurement method for assessing exposure to halogenated volatile organic compounds. Clin. Chem. 1997, 43, 723–730. [Google Scholar] [CrossRef] [Green Version]
  7. Beauchamp, J.; Davis, C.; Pleil, J. Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
  8. 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]
  9. Lawal, O.; Ahmed, W.M.; Nijsen, T.M.E.; Goodacre, R.; Fowler, S.J. Exhaled breath analysis: A review of ‘breath-taking’ methods for off-line analysis. Metabolomics 2017, 13, 1–16. [Google Scholar] [CrossRef]
  10. Rattray, N.J.W.; Hamrang, Z.; Trivedi, D.K.; Goodacre, R.; Fowler, S.J. Taking your breath away: Metabolomics breathes life in to personalized medicine. Trends Biotechnol. 2014, 32, 538–548. [Google Scholar] [CrossRef] [PubMed]
  11. Beauchamp, J.D.; Pleil, J.D. Simply breath-taking? Developing a strategy for consistent breath sampling. J. Breath Res. 2013, 7, 042001. [Google Scholar] [CrossRef] [Green Version]
  12. Herbig, J.; Beauchamp, J. Towards standardization in the analysis of breath gas volatiles. J. Breath Res. 2014, 8, 037101. [Google Scholar] [CrossRef] [PubMed]
  13. Risby, T.H. Critical issues for breath analysis. J. Breath Res. 2008, 2, 030302. [Google Scholar] [CrossRef]
  14. Grob, N.M.; Dweik, R.A. Exhaled nitric oxide in asthma: Progress since the introduction of standardized methodology. J. Breath Res. 2008, 2, 037002. [Google Scholar] [CrossRef]
  15. Pleil, J.D.; Stiegel, M.A.; Beauchamp, J.D. Chapter 1—Breath biomarkers and the exposome. In Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Beauchamp, J., Davis, C., Pleil, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 3–21. [Google Scholar]
  16. Lourenço, C.; Turner, C. Breath analysis in disease diagnosis: Methodological considerations and applications. Metabolites 2014, 4, 465–498. [Google Scholar] [CrossRef]
  17. Davis, M.D.; Fowler, S.J.; Montpetit, A.J. Exhaled breath testing—a tool for the clinician and researcher. Paediatr. Respir. Rev. 2019, 29, 37–41. [Google Scholar] [CrossRef]
  18. Wallace, M.A.G.; Pleil, J.D. Evolution of clinical and environmental health applications of exhaled breath research: Review of methods and instrumentation for gas-phase, condensate, and aerosols. Anal. Chim. Acta 2018, 1024, 18–38. [Google Scholar] [CrossRef]
  19. Basanta, M.; Ibrahim, B.; Dockry, R.; Douce, D.; Morris, M.; Singh, D.; Woodcock, A.; Fowler, S.J. Exhaled volatile organic compounds for phenotyping chronic obstructive pulmonary disease: A cross-sectional study. Respir. Res. 2012, 13, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Pleil, J.D.; Stiegel, M.A.; Risby, T.H. Clinical breath analysis: Discriminating between human endogenous compounds and exogenous (environmental) chemical confounders. J. Breath Res. 2013, 7, 017107. [Google Scholar] [CrossRef] [Green Version]
  21. Basanta, M.; Jarvis, R.M.; Xu, Y.; Blackburn, G.; Tal-Singer, R.; Woodcock, A.; Singh, D.; Goodacre, R.; Thomas, C.P.; Fowler, S.J. Non-invasive metabolomic analysis of breath using differential mobility spectrometry in patients with chronic obstructive pulmonary disease and healthy smokers. Analyst 2010, 135, 315–320. [Google Scholar] [CrossRef]
  22. Ibrahim, B.; Basanta, M.; Cadden, P.; Singh, D.; Douce, D.; Woodcock, A.; Fowler, S.J. Non-invasive phenotyping using exhaled volatile organic compounds in asthma. Thorax 2011, 66, 804–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Bruderer, T.; Gaisl, T.; Gaugg, M.T.; Nowak, N.; Streckenbach, B.; Müller, S.; Moeller, A.; Kohler, M.; Zenobi, R. On-Line Analysis of Exhaled Breath: Focus Review. Chem. Rev. 2019, 119, 10803–10828. [Google Scholar] [CrossRef] [PubMed]
  24. Boots, A.W.; Bos, L.D.; van der Schee, M.P.; van Schooten, F.J.; Sterk, P.J. Exhaled molecular fingerprinting in diagnosis and monitoring: Validating volatile promises. Trends Mol. Med. 2015, 21, 633–644. [Google Scholar] [CrossRef] [PubMed]
  25. Bos, L.D.J.; Sterk, P.J.; Schultz, M.J. Volatile metabolites of pathogens: A systematic review. PLoS Pathog. 2013, 9, e1003311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Schulz, S.; Dickschat, J.S. Bacterial volatiles: The smell of small organisms. Nat. Prod. Rep. 2007, 24, 814–842. [Google Scholar] [CrossRef]
  27. Goodacre, R. Metabolomics of a superorganism. J. Nutr. 2007, 137, 259S–266S. [Google Scholar] [CrossRef] [Green Version]
  28. Fowler, S.J.; Basanta-Sanchez, M.; Xu, Y.; Goodacre, R.; Dark, P.M. Surveillance for lower airway pathogens in mechanically ventilated patients by metabolomic analysis of exhaled breath: A case-control study. Thorax 2015, 70, 320–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Paschke, K.M.; Mashir, A.; Dweik, R.A. Clinical applications of breath testing. F1000 Med. Rep. 2010, 2. [Google Scholar] [CrossRef]
  30. O’Hara, M.E.; Clutton-Brock, T.H.; Green, S.; Mayhew, C.A. Endogenous volatile organic compounds in breath and blood of healthy volunteers: Examining breath analysis as a surrogate for blood measurements. J. Breath Res. 2009, 3, 027005. [Google Scholar] [CrossRef]
  31. Siobal, M.S. Monitoring exhaled carbon dioxide. Respir. Care 2016, 61, 1397–1416. [Google Scholar] [CrossRef] [Green Version]
  32. Phillips, M. Breath tests in medicine. Sci. Am. 1992, 267, 74–79. [Google Scholar] [CrossRef]
  33. Gest, H. Bicentenary homage to Dr Jan Ingen-Housz, MD (1730–1799), pioneer of photosynthesis research. Photosynth. Res. 2000, 63, 183–190. [Google Scholar] [CrossRef] [PubMed]
  34. Jaffe, M.B. Infrared measurement of carbon dioxide in the human breath: “Breathe-through” devices from Tyndall to the present day. Anesth. Analg. 2008, 107, 890–904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Berg, J.M.; Stryer, L.; Tymoczko, J.L. Stryer Biochemie; Springer Spektrum: Berlin, Germany, 2018. [Google Scholar]
  36. Whitaker, D.K. Time for capnography-everywhere. Anaesthesia 2011, 66, 544–549. [Google Scholar] [CrossRef]
  37. Checketts, M.R.; Alladi, R.; Ferguson, K.; Gemmell, L.; Handy, J.M.; Klein, A.A.; Love, N.J.; Misra, U.; Morris, C.; Nathanson, M.H.; et al. Recommendations for standards of monitoring during anaesthesia and recovery 2015: Association of Anaesthetists of Great Britain and Ireland. Anaesthesia 2016, 71, 85–93. [Google Scholar] [CrossRef]
  38. Nassar, B.S.; Schmidt, G.A. Capnography during critical illness. Chest 2016, 149, 576–585. [Google Scholar] [CrossRef]
  39. Bhende, M. End-tidal carbon dioxide monitoring in pediatrics-clinical applications. J. Postgrad. Med. 2001, 47, 215. [Google Scholar]
  40. Varon, A.J.; Morrina, J.; Civetta, J.M. Clinical utility of a colorimetric end-tidal CO2 detector in cardiopulmonary resuscitation and emergency intubation. J. Clin. Monit. 1991, 7, 289–293. [Google Scholar] [CrossRef]
  41. MacLeod, B.A.; Heller, M.B.; Gerard, J.; Yealy, D.M.; Menegazzi, J.J. Verification of endotracheal tube placement with colorimetric end-tidal CO2 detection. Ann. Emerg. Med. 1991, 20, 267–270. [Google Scholar] [CrossRef]
  42. Kim, K.W.; Choi, H.R.; Bang, S.R.; Lee, J.w. Comparison of end-tidal CO2 measured by transportable capnometer (EMMA™ capnograph) and arterial pCO2 in general anesthesia. J. Clin. Monit. Comput. 2016, 30, 737–741. [Google Scholar] [CrossRef]
  43. Kameyama, M.; Uehara, K.; Takatori, M.; Tada, K. Clinical usefulness of EMMA for monitoring end-tidal carbon dioxide. Masui 2013, 62, 477–480. [Google Scholar]
  44. Langhan, M.L.; Chen, L. Current utilization of continuous end-tidal carbon dioxide monitoring in pediatric emergency departments. Pediatr. Emerg. Care 2008, 24, 211–213. [Google Scholar] [CrossRef] [Green Version]
  45. Bhavani-Shankar, K.; Philip, J.H. Defining Segments and Phases of a Time Capnogram. Anesth. Analg. 2000, 91, 973–977. [Google Scholar] [CrossRef]
  46. Langhan, M.L.; Kurtz, J.C.; Schaeffer, P.; Asnes, A.G.; Riera, A. Experiences with capnography in acute care settings: A mixed-methods analysis of clinical staff. J. Crit. Care 2014, 29, 1035–1040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Erasmus, P.D. The use of end-tidal carbon dioxide monitoring to confirm endotracheal tube placement in adult and paediatric intensive care units in Australia and New Zealand. Anaesth. Intensive Care 2004, 32, 672–675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Deiorio, N. Continuous end-tidal carbon dioxide monitoring for confirmation of endotracheal tube placement is neither widely available nor consistently applied by emergency physicians. Emerg. Med. J. 2005, 22, 490–493. [Google Scholar] [CrossRef] [PubMed]
  49. Neumar, R.W.; Otto, C.W.; Link, M.S.; Kronick, S.L.; Shuster, M.; Callaway, C.W.; Kudenchuk, P.J.; Ornato, J.P.; McNally, B.; Silvers, S.M. Part 8: Adult advanced cardiovascular life support: 2010 American Heart Association guidelines for cardiopulmonary resuscitation and emergency cardiovascular care. Circulation 2010, 122, 729–767. [Google Scholar] [CrossRef] [Green Version]
  50. Beauchamp, J.D.; Miekisch, W. Chapter 2—Breath sampling and standardization. In Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Beauchamp, J., Davis, C., Pleil, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 23–41. [Google Scholar]
  51. Högman, M.; Lehtimäki, L. Chapter 4—Exhaled nitric oxide physiology and modeling. In Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Beauchamp, J., Davis, C., Pleil, J., Eds.; Elsevier: Boston, MA, USA, 2020; pp. 63–77. [Google Scholar]
  52. Hogman, M.; Stromberg, S.; Schedin, U.; Frostell, C.; Hedenstierna, G.; Gustafsson, L. Nitric oxide from the human respiratory tract efficiently quantified by standardized single breath measurements. Acta Physiol. Scand. 1997, 159, 345–346. [Google Scholar] [CrossRef]
  53. Silkoff, P.E.; McClean, P.A.; Slutsky, A.S.; Furlott, H.G.; Hoffstein, E.; Wakita, S.; Chapman, K.R.; Szalai, J.P.; Zamel, N. Marked flow-dependence of exhaled nitric oxide using a new technique to exclude nasal nitric oxide. Am. J. Respir. Crit. Care Med. 1997, 155, 260–267. [Google Scholar] [CrossRef]
  54. Grob, N.M.; Dweik, R.A. Exhaled nitric oxide in asthma: From diagnosis, to monitoring, to screening: Are we there yet? Chest 2008, 133, 837–839. [Google Scholar] [CrossRef]
  55. West, J.B. Joseph Priestley, oxygen, and the enlightenment. Am. J. Physiol. Lung Cell. Mol. Physiol. 2014, 306, L111–L119. [Google Scholar] [CrossRef] [Green Version]
  56. Ignarro, L.J.; Buga, G.M.; Wood, K.S.; Byrns, R.E.; Chaudhuri, G. Endothelium-derived relaxing factor produced and released from artery and vein is nitric oxide. Proc. Natl. Acad. Sci. USA 1987, 84, 9265–9269. [Google Scholar] [CrossRef] [Green Version]
  57. Furchgott, R.F.; Zawadzki, J.V. The obligatory role of endothelial cells in the relaxation of arterial smooth muscle by acetylcholine. Nature 1980, 288, 373–376. [Google Scholar] [CrossRef]
  58. Arnold, W.P.; Mittal, C.K.; Katsuki, S.; Murad, F. Nitric oxide activates guanylate cyclase and increases guanosine 3′: 5′-Cyclic monophosphate levels in various tissue preparations. Proc. Natl. Acad. Sci. USA 1977, 74, 3203–3207. [Google Scholar] [CrossRef] [Green Version]
  59. Gustafsson, L.E.; Leone, A.M.; Persson, M.G.; Wiklund, N.P.; Moncada, S. Endogenous nitric oxide is present in the exhaled air of rabbits, guinea pigs and humans. Biochem. Biophys. Res. Commun. 1991, 181, 852–857. [Google Scholar] [CrossRef]
  60. Högman, M.; Risby, T. The unique contribution of Professor Lars E Gustafsson to the field of breath research. J. Breath Res. 2017, 11, 040201. [Google Scholar] [CrossRef]
  61. Alving, K.; Weitzberg, E.; Lundberg, J.M. Increased amount of nitric oxide in exhaled air of asthmatics. Eur. Respir. J. 1993, 6, 1368–1370. [Google Scholar] [PubMed]
  62. Kharitonov, S.A.; Yates, D.; Robbins, R.A.; Barnes, P.J.; Logan-Sinclair, R.; Shinebourne, E.A. Increased nitric oxide in exhaled air of asthmatic patients. Lancet 1994, 343, 133–135. [Google Scholar] [CrossRef]
  63. Persson, M.G.; Gustafsson, L.E.; Zetterström, O.; Agrenius, V.; Ihre, E. Single-breath nitric oxide measurements in asthmatic patients and smokers. Lancet 1994, 343, 146–147. [Google Scholar] [CrossRef]
  64. Lundberg, J.; Farkas-Szallasi, T.; Weitzberg, E.; Rinder, J.; Lidholm, J.; Änggåard, A.; Hökfelt, T.; Lundberg, J.; Alving, K. High nitric oxide production in human paranasal sinuses. Nat. Med. 1995, 1, 370–373. [Google Scholar] [CrossRef]
  65. Pijnenburg, M.W.H.; De Jongste, J.C. Exhaled nitric oxide in childhood asthma: A review. Clin. Exp. Allergy 2008, 38, 246–259. [Google Scholar] [CrossRef]
  66. Moncada, S.; Higgs, A. The L-arginine-nitric oxide pathway. N. Engl. J. Med. 1993, 329, 2002–2012. [Google Scholar]
  67. Förstermann, U.; Sessa, W.C. Nitric oxide synthases: Regulation and function. Eur. Heart J. 2012, 33, 829–837. [Google Scholar] [CrossRef] [Green Version]
  68. Hamad, A.M.; Clayton, A.; Islam, B.; Knox, A.J. Guanylyl cyclases, nitric oxide, natriuretic peptides, and airway smooth muscle function. Am. J. Physiol. Lung Cell. Mol. Physiol. 2003, 285, 973–983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Malinovschi, A.; Michils, A.; Högman, M. Chapter 5—Exhaled nitric oxide in clinical practice. In Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Beauchamp, J., Davis, C., Pleil, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 81–92. [Google Scholar]
  70. Aktan, F. iNOS-mediated nitric oxide production and its regulation. Life Sci. 2004, 75, 639–653. [Google Scholar] [CrossRef] [PubMed]
  71. Chibana, K.; Trudeau, J.B.; Mustovitch, A.T.; Hu, H.; Zhao, J.; Balzar, S.; Chu, H.W.; Wenzel, S.E. IL-13 induced increases in nitrite levels are primarily driven by increases in inducible nitric oxide synthase as compared with effects on arginases in human primary bronchial epithelial cells. Clin. Exp. Allergy 2008, 38, 936–946. [Google Scholar] [CrossRef]
  72. Nathan, C.; Xie, Q.W. Nitric oxide synthases: Roles, tolls, and controls. Cell 1994, 78, 915–918. [Google Scholar] [CrossRef]
  73. Menzies-Gow, A.; Mansur, A.H.; Brightling, C.E. Clinical utility of fractional exhaled nitric oxide in severe asthma management. Eur. Respir. J. 2020, 55. [Google Scholar] [CrossRef]
  74. Ricciardolo, F.L.; Silkoff, P.E. Perspectives on exhaled nitric oxide. J. Breath Res. 2017, 11, 047104. [Google Scholar] [CrossRef] [Green Version]
  75. Robbins, R.A.; Millatmal, T.; Lassi, K.; Rennard, S.; Daughton, D. Smoking cessation is associated with an increase in exhaled nitric oxide. Chest 1997, 112, 313–318. [Google Scholar] [CrossRef] [Green Version]
  76. Kissoon, N.; Duckworth, L.J.; Blake, K.V.; Murphy, S.P.; Taylor, C.L.; DeNicola, L.R.; Silkoff, P.E. Exhaled nitric oxide concentrations: Online versus offline values in healthy children. Pediatr. Pulmonol. 2002, 33, 283–292. [Google Scholar] [CrossRef] [PubMed]
  77. Bruce, C.; Yates, D.H.; Thomas, P.S. Caffeine decreases exhaled nitric oxide. Thorax 2002, 57, 361–363. [Google Scholar] [CrossRef] [Green Version]
  78. Högman, M.; Drca, N.; Ehrstedt, C.; Meriläinen, P. Exhaled nitric oxide partitioned into alveolar, lower airways and nasal contributions. Respir. Med. 2000, 94, 985–991. [Google Scholar] [CrossRef] [Green Version]
  79. Högman, M.; Thornadtsson, A.; Hedenstierna, G.; Meriläinen, P. A practical approach to the theoretical models to calculate NO parameters of the respiratory system. J. Breath Res. 2014, 8, 016002. [Google Scholar] [CrossRef] [PubMed]
  80. Kharitonov, S.; Alving, K.; Barnes, P. Exhaled and nasal nitric oxide measurements: Recommendations. The European Respiratory Society task force. Eur. Respir. J. 1997, 10, 1683–1693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. American Thoracic Society. Recommendations for standardized procedures for the on-line and off-line measurement of exhaled lower respiratory nitric oxide and nasal nitric oxide in adults and children-1999. This official statement of the American Thoracic Society was adopted by the ATS Board of Directors. Am. J. Respir. Crit. Care Med. 1999, 160, 2104–2117. [Google Scholar]
  82. American Thoracic Society/European Respiratory Society. ATS/ERS recommendations for standardized procedures for the online and offline measurement of exhaled lower respiratory nitric oxide and nasal nitric oxide. Am. J. Respir. Crit. Care Med. 2005, 171, 912–930. [Google Scholar] [CrossRef]
  83. Modak, A.S. Breath biomarkers for personalized medicine. Per. Med. 2010, 7, 643–653. [Google Scholar] [CrossRef] [PubMed]
  84. Silkoff, P.E.; Stevens, A.; Pak, J.; Bucher-Bartelson, B.; Martin, R.J. A Method for the Standardized Offline Collection of Exhaled Nitric Oxide. Chest 1999, 116, 754–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Van Der Heijden, H.H.A.C.M.; Brouwer, M.L.; Hoekstra, F.; Van Der Pol, P.; Merkus, P.J.F.M. Reference values of exhaled nitric oxide in healthy children 1–5 years using off-line tidal breathing. Pediatr. Pulmonol. 2014, 49, 291–295. [Google Scholar] [CrossRef]
  86. Zhang, Y.; Eckel, S.; Berhane, K.; Garcia, E.; Muchmore, P.; Molshatzki, N.B.A.; Rappaport, E.; Linn, W.S.; Habre, R.; Gilliland, F. Long-term Exposures to Air Pollutants Affect FeNO in Children: A Longitudinal Study. MedRxiv 2021. [Google Scholar] [CrossRef]
  87. Buchvald, F.; Baraldi, E.; Carraro, S.; Gaston, B.; De Jongste, J.; Pijnenburg, M.W.; Silkoff, P.E.; Bisgaard, H. Measurements of exhaled nitric oxide in healthy subjects age 4 to 17 years. J. Allergy Clin. Immunol. 2005, 115, 1130–1136. [Google Scholar] [CrossRef]
  88. Gehring, U.; Oldenwening, M.; Brunekreef, B.; Wieringa, M.H.; Kerkhof, M.; Smit, H.A.; Van Der Ent, C.K.; De Jongste, J.C. The impact of ambient NO on online measurements of exhaled and nasal NO: The PIAMA study. Pediatric Allergy Immunol. 2009, 20, 665–672. [Google Scholar] [CrossRef]
  89. Ma’pol, A.; Hashim, J.H.; Norbäck, D.; Weislander, G.; Hashim, Z.; Isa, Z.M. FeNO level and allergy status among school children in Terengganu, Malaysia. J. Asthma 2020, 57, 842–849. [Google Scholar] [CrossRef]
  90. Linn, W.S.; Berhane, K.T.; Rappaport, E.B.; Bastain, T.M.; Avol, E.L.; Gilliland, F.D. Relationships of online exhaled, offline exhaled, and ambient nitric oxide in an epidemiologic survey of schoolchildren. J. Expo. Sci. Environ. Epidemiol. 2009, 19, 674–681. [Google Scholar] [CrossRef] [Green Version]
  91. 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] [PubMed]
  92. Horváth, I.; Barnes, P.J.; Loukides, S.; Sterk, P.J.; Högman, M.; Olin, A.C.; Amann, A.; Antus, B.; Baraldi, E.; Bikov, A. A European Respiratory Society technical standard: Exhaled biomarkers in lung disease. Eur. Respir. J. 2017, 49, 1600965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Tornberg, D.C.; Bjorne, H.; Lundberg, J.O.; Weitzberg, E. Multiple single-breath measurements of nitric oxide in the intubated patient. Am. J. Respir. Crit. Care Med. 2003, 168, 1210–1215. [Google Scholar] [CrossRef] [PubMed]
  94. Heffler, E.; Carpagnano, G.E.; Favero, E.; Guida, G.; Maniscalco, M.; Motta, A.; Paoletti, G.; Rolla, G.; Baraldi, E.; Pezzella, V. Fractional Exhaled Nitric Oxide (FENO) in the management of asthma: A position paper of the Italian Respiratory Society (SIP/IRS) and Italian Society of Allergy, Asthma and Clinical Immunology (SIAAIC). Multidiscip. Respir. Med. 2020, 15, 36. [Google Scholar] [CrossRef] [Green Version]
  95. Rupani, H.; Chauhan, A.J. Measurement of FeNO in asthma: What the hospital doctor needs to know. Br. J. Hosp. Med. 2019, 80, 99–104. [Google Scholar] [CrossRef] [PubMed]
  96. Spahn, J.D.; Malka, J.; Szefler, S.J. Current application of exhaled nitric oxide in clinical practice. J. Allergy Clin. Immunol. 2016, 138, 1296–1298. [Google Scholar] [CrossRef] [Green Version]
  97. Price, D.B.; Buhl, R.; Chan, A.; Freeman, D.; Gardener, E.; Godley, C.; Gruffydd-Jones, K.; McGarvey, L.; Ohta, K.; Ryan, D. Fractional exhaled nitric oxide as a predictor of response to inhaled corticosteroids in patients with non-specific respiratory symptoms and insignificant bronchodilator reversibility: A randomised controlled trial. Lancet Respir. Med. 2018, 6, 29–39. [Google Scholar] [CrossRef] [Green Version]
  98. Bjermer, L.; Alving, K.; Diamant, Z.; Magnussen, H.; Pavord, I.; Piacentini, G.; Price, D.; Roche, N.; Sastre, J.; Thomas, M. Current evidence and future research needs for FeNO measurement in respiratory diseases. Respir. Med. 2014, 108, 830–841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Heaney, L.G.; Busby, J.; Bradding, P.; Chaudhuri, R.; Mansur, A.H.; Niven, R.; Pavord, I.D.; Lindsay, J.T.; Costello, R.W. Remotely monitored therapy and nitric oxide suppression identifies nonadherence in severe asthma. Am. J. Respir. Crit. Care Med. 2019, 199, 454–464. [Google Scholar] [CrossRef] [Green Version]
  100. Price, D.; Ryan, D.; Burden, A.; Von Ziegenweidt, J.; Gould, S.; Freeman, D.; Gruffydd-Jones, K.; Copland, A.; Godley, C.; Chisholm, A. Using fractional exhaled nitric oxide (FeNO) to diagnose steroid-responsive disease and guide asthma management in routine care. Clin. Transl. Allergy 2013, 3, 1–10. [Google Scholar] [CrossRef] [Green Version]
  101. Gréhant, N. Les Gaz du Sang: Applications à L’hygiène Expérimentale; Forgotten Books: Paris, French, 1894. (In French) [Google Scholar]
  102. Lepine, R.; Boulud, J. Sur l’existence d’oxyde de carbone dans le sang des anemiques. CR Soc. Biol. 1905, 57, 55. (In French) [Google Scholar]
  103. Sjöstrand, T. Endogenous formation of carbon monoxide in man under normal and pathological conditions. Scand. J. Clin. Lab. Investig. 1949, 1, 201–214. [Google Scholar] [CrossRef]
  104. Tenhunen, R.; Marver, H.S.; Schmid, R. Microsomal heme oxygenase characterization of the enzyme. J. Biol. Chem. 1969, 244, 6388–6394. [Google Scholar] [CrossRef]
  105. Berk, P.D.; Rodkey, F.L.; Blaschke, T.F.; Collison, H.A.; Waggoner, J.G. Comparison of plasma bilirubin turnover and carbon monoxide production in man. J. Lab. Clin. Med. 1974, 83, 29–37. [Google Scholar] [PubMed]
  106. Landaw, S.A.; Callahan, E.W.; Schmid, R. Catabolism of heme in vivo: Comparison of the simultaneous production of bilirubin and carbon monoxide. J. Clin. Investig. 1970, 49, 914–925. [Google Scholar] [CrossRef] [Green Version]
  107. Ryter, S.W.; Choi, A.M.K. Targeting heme oxygenase-1 and carbon monoxide for therapeutic modulation of inflammation. Transl. Res. 2016, 167, 7–34. [Google Scholar] [CrossRef] [Green Version]
  108. Ryter, S.W. Chapter 6—Exhaled carbon monoxide. In Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Beauchamp, J., Davis, C., Pleil, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 93–107. [Google Scholar]
  109. Okuyama, H.; Yonetani, M.; Uetani, Y.; Nakamura, H. End-tidal carbon monoxide is predictive for neonatal non-hemolytic hyperbilirubinemia. Pediatr. Int. 2001, 43, 329–333. [Google Scholar] [CrossRef]
  110. Cohen, R.S.; Wong, R.J.; Stevenson, D.K. Understanding neonatal jaundice: A perspective on causation. Pediatr. Neonatol. 2010, 51, 143–148. [Google Scholar] [CrossRef] [Green Version]
  111. Stevenson, D.K.; Vreman, H.J. Carbon monoxide and bilirubin production in neonates. Pediatrics 1997, 100, 252–259. [Google Scholar] [CrossRef]
  112. Johnson, L.; Bhutani, V.K. The clinical syndrome of bilirubin-induced neurologic dysfunction. Semin. Perinatol. 2011, 35, 101–113. [Google Scholar] [CrossRef] [PubMed]
  113. Tidmarsh, G.F.; Wong, R.J.; Stevenson, D.K. End-tidal carbon monoxide and hemolysis. J. Perinatol. 2014, 34, 577–581. [Google Scholar] [CrossRef] [PubMed]
  114. Watchko, J.F.; Tiribelli, C. Bilirubin-induced neurologic damage—Mechanisms and management approaches. N. Engl. J. Med. 2013, 369, 2021–2030. [Google Scholar] [CrossRef]
  115. Kaplan, M.; Bromiker, R.; Hammerman, C. Hyperbilirubinemia, hemolysis, and increased bilirubin neurotoxicity. Semin. Perinatol. 2014, 38, 429–437. [Google Scholar] [CrossRef] [PubMed]
  116. Cohen, R.S.; Ostrander, C.R.; Cowan, B.E.; Stevens, G.B.; Hopper, A.O.; Stevenson, D.K. Pulmonary excretion rates of carbon monoxide using a modified technique: Differences between premature and full-term infants. Neonatology 1982, 41, 289–293. [Google Scholar] [CrossRef] [PubMed]
  117. Smith, D.W.; Hopper, A.O.; Shahin, S.M.; Cohen, R.S.; Ostrander, C.R.; Ariagno, R.L.; Stevenson, D.K. Neonatal bilirubin production estimated from “end-tidal” carbon monoxide concentration. J. Pediatr. Gastroenterol. Nutr. 1984, 3, 77–80. [Google Scholar] [CrossRef]
  118. Vreman, H.J.; Mahoney, J.J.; Stevenson, D.K. Carbon monoxide and carboxyhemoglobin. Adv. Pediatr. 1995, 42, 303–334. [Google Scholar]
  119. Bhutani, V.K.; Maisels, M.J.; Schutzman, D.L.; Castillo Cuadrado, M.E.; Aby, J.L.; Bogen, D.L.; Christensen, R.D.; Watchko, J.F.; Wong, R.J.; Stevenson, D.K. Identification of risk for neonatal haemolysis. Acta Paediatr. 2018, 107, 1350–1356. [Google Scholar] [CrossRef] [PubMed]
  120. Bhutani, V.K.; Srinivas, S.; Castillo Cuadrado, M.E.; Aby, J.L.; Wong, R.J.; Stevenson, D.K. Identification of neonatal haemolysis: An approach to predischarge management of neonatal hyperbilirubinemia. Acta Paediatr. 2016, 105, 189–194. [Google Scholar] [CrossRef] [PubMed]
  121. Castillo Cuadrado, M.E.; Bhutani, V.K.; Aby, J.L.; Vreman, H.J.; Wong, R.J.; Stevenson, D.K. Evaluation of a new end-tidal carbon monoxide monitor from the bench to the bedside. Acta Paediatr. 2015, 104, 279–282. [Google Scholar] [CrossRef]
  122. Christensen, R.D.; Malleske, D.T.; Lambert, D.K.; Baer, V.L.; Prchal, J.T.; Denson, L.E.; Gerday, E.; Lewis, K.A.W.; Shepherd, J.G. Measuring end-tidal carbon monoxide of jaundiced neonates in the birth hospital to identify those with hemolysis. Neonatology 2016, 109, 1–5. [Google Scholar] [CrossRef]
  123. Elsaie, A.L.; Taleb, M.; Nicosia, A.; Zangaladze, A.; Pease, M.E.; Newton, K.; Schutzman, D.L. Comparison of end-tidal carbon monoxide measurements with direct antiglobulin tests in the management of neonatal hyperbilirubinemia. J. Perinatol. 2020, 1–5. [Google Scholar] [CrossRef]
  124. Bhatia, A.; Chua, M.C.; dela Puerta, R.; Rajadurai, V.S. Noninvasive Detection of Hemolysis with ETCOc Measurement in Neonates at Risk for Significant Hyperbilirubinemia. Neonatology 2020, 117, 612–618. [Google Scholar] [CrossRef] [PubMed]
  125. Conrad, N.; Judge, A.; Tran, J.; Mohseni, H.; Hedgecott, D.; Crespillo, A.P.; Allison, M.; Hemingway, H.; Cleland, J.G.; McMurray, J.J. Temporal trends and patterns in heart failure incidence: A population-based study of 4 million individuals. Lancet 2018, 391, 572–580. [Google Scholar] [CrossRef] [Green Version]
  126. McMinn, J.F.; Lang, N.N.; McPhadden, A.; Payne, J.R.; Petrie, M.C.; Gardner, R.S. Biomarkers of acute rejection following cardiac transplantation. Biomark. Med. 2014, 8, 815–832. [Google Scholar] [CrossRef]
  127. Cleland, J.G.; van Veldhuisen, D.J.; Ponikowski, P. The year in cardiology 2018: Heart failure. Eur. Heart J. 2019, 40, 651–661. [Google Scholar] [CrossRef] [Green Version]
  128. Phillips, M.; Boehmer, J.P.; Cataneo, R.N.; Cheema, T.; Eisen, H.J.; Fallon, J.T.; Fisher, P.E.; Gass, A.; Greenberg, J.; Kobashigawa, J.; et al. Heart allograft rejection: Detection with breath alkanes in low levels (the HARDBALL study). J. Heart Lung Transplant. 2004, 23, 701–708. [Google Scholar] [CrossRef]
  129. Costanzo, M.R.; Dipchand, A.; Starling, R.; Anderson, A.; Chan, M.; Desai, S.; Fedson, S.; Fisher, P.; Gonzales-Stawinski, G.; Martinelli, L. The International Society of Heart and Lung Transplantation Guidelines for the care of heart transplant recipients. J. Heart Lung Transplant. 2010, 29, 915–956. [Google Scholar] [CrossRef]
  130. Weber, B.N.; Kobashigawa, J.A.; Givertz, M.M. Evolving areas in heart transplantation. JACC: Heart Fail. 2017, 5, 869–878. [Google Scholar] [CrossRef]
  131. Khaw, B.A.; Narula, J. Antibody imaging in the evaluation of cardiovascular diseases. J. Nucl. Cardiol. 1994, 1, 457–476. [Google Scholar] [CrossRef]
  132. Moran, A.M.; Lipshultz, S.E.; Rifai, N.; O’Brien, P.; Mooney, H.; Perry, S.; Perez-Atayde, A.; Lipsitz, S.R.; Colan, S.D. Non-invasive assessment of rejection in pediatric transplant patients: Serologic and echocardiographic prediction of biopsy-proven myocardial rejection. J. Heart Lung Transplant. 2000, 19, 756–764. [Google Scholar] [CrossRef]
  133. Phillips, M.; Boehmer, J.P.; Cataneo, R.N.; Cheema, T.; Eisen, H.J.; Fallon, J.T.; Fisher, P.E.; Gass, A.; Greenberg, J.; Kobashigawa, J. Prediction of heart transplant rejection with a breath test for markers of oxidative stress. Am. J. Cardiol. 2004, 94, 1593–1594. [Google Scholar] [CrossRef] [PubMed]
  134. Kneepkens, C.F.; Lepage, G.; Roy, C.C. The potential of the hydrocarbon breath test as a measure of lipid peroxidation. Free Radic. Biol. Med. 1994, 17, 127–160. [Google Scholar] [CrossRef]
  135. Phillips, M.; Cataneo, R.N.; Greenberg, J.; Gunawardena, R.; Naidu, A.; Rahbari-Oskoui, F. Effect of age on the breath methylated alkane contour, a display of apparent new markers of oxidative stress. J. Lab. Clin. Med. 2000, 136, 243–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Phillips, M. Method for the collection and assay of volatile organic compounds in breath. Anal. Biochem. 1997, 247, 272–278. [Google Scholar] [CrossRef] [Green Version]
  137. Cargnoni, A.; Ceconi, C.; Bernocchi, P.; Parrinello, G.; Benigno, M.; Boraso, A.; Curello, S.; Ferrari, R. Changes in oxidative stress and cellular redox potential during myocardial storage for transplantation: Experimental studies. J. Heart Lung Transplant. 1999, 18, 478–487. [Google Scholar] [CrossRef]
  138. Food and Drug Administration, U.S. Humanitarian Device Exemption (HDE) Program. Guidance for Industry and Food and Drug Administration Staff. Available online: https://www.fda.gov/media/74307/download (accessed on 13 July 2021).
  139. Beauchamp, J. Inhaled today, not gone tomorrow: Pharmacokinetics and environmental exposure of volatiles in exhaled breath. J. Breath Res. 2011, 5, 037103. [Google Scholar] [CrossRef]
  140. Pleil, J.D.; Stiegel, M.A.; Sobus, J.R. Breath biomarkers in environmental health science: Exploring patterns in the human exposome. J. Breath Res. 2011, 5, 046005. [Google Scholar] [CrossRef] [Green Version]
  141. Phillips, M.; Altorki, N.; Austin, J.H.; Cameron, R.B.; Cataneo, R.N.; Greenberg, J.; Kloss, R.; Maxfield, R.A.; Munawar, M.I.; Pass, H.I. Prediction of lung cancer using volatile biomarkers in breath. Cancer Biomark. 2007, 3, 95–109. [Google Scholar] [CrossRef] [Green Version]
  142. Bikov, A.; Paschalaki, K.; Logan-Sinclair, R.; Horváth, I.; Kharitonov, S.A.; Barnes, P.J.; Usmani, O.S.; Paredi, P. Standardised exhaled breath collection for the measurement of exhaled volatile organic compounds by proton transfer reaction mass spectrometry. BMC Pulm. Med. 2013, 13, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Kischkel, S.; Miekisch, W.; Sawacki, A.; Straker, E.; Trefz, P.; Amann, A.; Schubert, J. Breath biomarkers for lung cancer detection and assessment of smoking related effects—Confounding variables, influence of normalization and statistical algorithms. Clin. Chim. Acta 2010, 411, 1637–1644. [Google Scholar] [CrossRef]
  144. 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] [PubMed]
  145. Gaude, E.; Nakhleh, M.K.; Patassini, S.; Boschmans, J.; Allsworth, M.; Boyle, B.; van der Schee, M.P. Targeted breath analysis: Exogenous volatile organic compounds (EVOC) as metabolic pathway-specific probes. J. Breath Res. 2019, 13, 032001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Wigmore, J.G.; Langille, R.M. Six generations of breath alcohol testing instruments: Changes in the detection of breath alcohol since 1930. An historical overview. Can. Soc. Forensic Sci. J. 2009, 42, 276–283. [Google Scholar] [CrossRef]
  147. Eisenmann, A.; Amann, A.; Said, M.; Datta, B.; Ledochowski, M. Implementation and interpretation of hydrogen breath tests. J. Breath Res. 2008, 2, 046002. [Google Scholar] [CrossRef]
  148. Savarino, V.; Vigneri, S.; Celle, G. The 13C urea breath test in the diagnosis of Helicobacter pylori infection. Gut 1999, 45, I18–I22. [Google Scholar] [CrossRef]
  149. Bharucha, A.E.; Camilleri, M.; Veil, E.; Burton, D.; Zinsmeister, A.R. Comprehensive assessment of gastric emptying with a stable isotope breath test. Neurogastroenterol. Motil. 2013, 25, 60–69. [Google Scholar] [CrossRef]
  150. Jara, M.; Bednarsch, J.; Valle, E.; Lock, J.F.; Malinowski, M.; Schulz, A.; Seehofer, D.; Jung, T.; Stockmann, M. Reliable assessment of liver function using LiMAx. J. Surg. Res. 2015, 193, 184–189. [Google Scholar] [CrossRef] [PubMed]
  151. Kalant, H. Research on tolerance: What can we learn from history? Alcohol. Clin. Exp. Res. 1998, 22, 67–76. [Google Scholar] [CrossRef] [PubMed]
  152. Anstie, F.E. Prognosis and treatment of certain acute diseases. Lancet 1867, 28, 385–387. [Google Scholar]
  153. Jones, A.W. Chapter 29—Breath analysis in law enforcement. In Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Beauchamp, J., Davis, C., Pleil, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 475–492. [Google Scholar]
  154. Liljestrand, G.; Linde, P.T. Über die Ausscheidung des Alkohols mit der Exspirationsluft. Skand. Arch. Für Physiol. 1930, 60, 273–298. [Google Scholar] [CrossRef]
  155. Kalant, H. Pharmacokinetics of ethanol: Absorption, distribution, and elimination. In The Pharmacology of Alcohol and Alcohol Dependence; Begleiter, H., Kissin, B., Eds.; Oxford University Press, Inc.: Oxford, NY, USA, 1996; pp. 15–58. [Google Scholar]
  156. Cowan, J.M.; Burris, J.M.; Hughes, J.R.; Cunningham, M.P. The relationship of normal body temperature, end-expired breath temperature, and BAC/BrAC ratio in 98 physically fit human test subjects. J. Anal. Toxicol. 2010, 34, 238–242. [Google Scholar] [CrossRef]
  157. Jones, A.W. Variability of the blood:breath alcohol ratio in vivo. J. Stud. Alcohol Drugs 1978, 39, 1931–1939. [Google Scholar] [CrossRef]
  158. Jones, A.W.; Andersson, L. Comparison of ethanol concentrations in venous blood and end-expired breath during a controlled drinking study. Forensic Sci. Int. 2003, 132, 18–25. [Google Scholar] [CrossRef] [Green Version]
  159. Taylor, M.D.; Hodgson, B.T. Blood/breath correlations: Intoxilyzer® 5000C, Alcotest® 7110, and Breathalyzer® 900A breath alcohol analyzers. Can. Soc. Forensic Sci. J. 1995, 28, 153–164. [Google Scholar]
  160. Jurič, A.; Fijačko, A.; Bakulić, L.; Orešić, T.; Gmajnički, I. Evaluation of breath alcohol analysers by comparison of breath and blood alcohol concentrations. Arh. Hig. Rada Toksikol. 2018, 69, 69–76. [Google Scholar] [CrossRef] [Green Version]
  161. Sterling, K. The rate of dissipation of mouth alcohol in alcohol positive subjects. J. Forensic Sci. 2012, 57, 802–805. [Google Scholar] [CrossRef]
  162. Anderson, J.C.; Hlastala, M.P. The alcohol breath test in practice: Effects of exhaled volume. J. Appl. Physiol. 2019, 126, 1630–1635. [Google Scholar] [CrossRef]
  163. Jones, A.W. How breathing technique can influence the results of breath-alcohol analysis. Med. Sci. Law 1982, 22, 275–280. [Google Scholar] [CrossRef] [PubMed]
  164. Kovačić, Z.; Nestić, M.; Stemberga, V.; Bosnar, A.; Petrovečki, M.; Sutlović, D. Reliability of breath alcohol testing with Dräger Alcotest 7410Plus analyzer in a court process. Med. Jad. 2008, 38, 47–51. [Google Scholar]
  165. Trafford, D.J.H.; Makin, H.L.J. Breath-alcohol concentration may not always reflect the concentration of alcohol in blood. J. Anal. Toxicol. 1994, 18, 225–228. [Google Scholar] [CrossRef] [PubMed]
  166. Harger, R.N.; Lamb, E.B.; Hulpieu, H.R. A rapid chemical test for intoxication employing breath: A new reagent for alcohol and a procedure for estimating the concentration of alcohol in the body from the ratio of alcohol to carbon dioxide in the breath. J. Am. Med Assoc. 1938, 110, 779–785. [Google Scholar] [CrossRef]
  167. Harger, R.N. Debunking the Drunkometer. J. Crim. L. Criminol. 1949, 40, 497. [Google Scholar] [CrossRef]
  168. Borkenstein, R.F.; Smith, H.W. The breathalyzer and its applications. Med. Sci. Law 1961, 2, 13–22. [Google Scholar] [CrossRef]
  169. Grubb, D.; Rasmussen, B.; Linnet, K.; Olsson, S.; Lindberg, L. Breath alcohol analysis incorporating standardization to water vapour is as precise as blood alcohol analysis. Forensic Sci. Int. 2012, 216, 88–91. [Google Scholar] [CrossRef]
  170. Grubb, D.; Frigyesi, A.; Finnhult, M.; Dencker, D.; Olsson, S.G.; Lindberg, L. Breath alcohol analysis by standardization to water vapour enables contact free sampling with preserved high accuracy and precision as compared with mouthpiece sampling. J. Forensic Investig. 2014, 2, 1–6. [Google Scholar]
  171. 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]
  172. Ferris, J.; Mazerolle, L.; King, M.; Bates, L.; Bennett, S.; Devaney, M. Random breath testing in Queensland and Western Australia: Examination of how the random breath testing rate influences alcohol related traffic crash rates. Accid. Anal. Prev. 2013, 60, 181–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Kriikku, P.; Wilhelm, L.; Jenckel, S.; Rintatalo, J.; Hurme, J.; Kramer, J.; Jones, A.W.; Ojanperä, I. Comparison of breath-alcohol screening test results with venous blood alcohol concentration in suspected drunken drivers. Forensic Sci. Int. 2014, 239, 57–61. [Google Scholar] [CrossRef] [PubMed]
  174. Dubowski, K.M. Quality assurance in breath-alcohol analysis. J. Anal. Toxicol. 1994, 18, 306–311. [Google Scholar] [CrossRef] [Green Version]
  175. Metz, G.; Peters, T.; Jenkins, D.A.; Newman, A.; Blendis, L. Breath hydrogen as a diagnostic method for hypolactasia. Lancet 1975, 305, 1155–1157. [Google Scholar] [CrossRef]
  176. Satta, P.U.; Anania, C.; Astegiano, M.; Miceli, E.; Montalto, M.; Tursi, A. H2-breath testing for carbohydrate malabsorption. Aliment. Pharm. Ther. 2009, 29, 14–18. [Google Scholar]
  177. Solomons, N.W.; Garcia-Ibanez, R.; Viteri, F.E. Hydrogen breath test of lactose absorption in adults: The application of physiological doses and whole cow’s milk sources. Am. J. Clin. Nutr. 1980, 33, 545–554. [Google Scholar] [CrossRef]
  178. Hovde, Ø.; Farup, P.G. A comparison of diagnostic tests for lactose malabsorption-which one is the best? BMC Gastroenterol. 2009, 9, 1–7. [Google Scholar] [CrossRef] [Green Version]
  179. Arola, H. Diagnosis of hypolactasia and lactose malabsorption. Scand. J. Gastroenterol. 1994, 29, 26–35. [Google Scholar] [CrossRef]
  180. Misselwitz, B.; Pohl, D.; Fruhauf, H.; Fried, M.; Vavricka, S.R.; Fox, M. Lactose malabsorption and intolerance: Pathogenesis, diagnosis and treatment. United Eur. Gastroenterol. J. 2013, 1, 151–159. [Google Scholar] [CrossRef]
  181. Newcomer, A.D.; McGill, D.B.; Thomas, P.J.; Hofmann, A.F. Prospective comparison of indirect methods for detecting lactase deficiency. N. Engl. J. Med. 1975, 293, 1232–1236. [Google Scholar] [CrossRef]
  182. Casellas, F.; Malagelada, J.R. Applicability of short hydrogen breath test for screening of lactose malabsorption. Dig. Dis. Sci. 2003, 48, 1333–1338. [Google Scholar]
  183. Ruzsanyi, V.; Heinz-Erian, P.; Entenmann, A.; Karall, D.; Müller, T.; Schimkowitsch, A.; Amann, A.; Scholl-Bürgi, S. Diagnosing lactose malabsorption in children: Difficulties in interpreting hydrogen breath test results. J. Breath Res. 2016, 10, 016015. [Google Scholar] [CrossRef]
  184. Stefano, M.D.; Certo, M.; Colecchia, A.; Sorge, M.; Perri, F. H2-breath tests: Methodological audits in adults and children. Aliment. Pharm. Ther. 2009, 29, 8–13. [Google Scholar]
  185. Simren, M.; Stotzer, P.O. Use and abuse of hydrogen breath tests. Gut 2006, 55, 297–303. [Google Scholar] [CrossRef]
  186. Braden, B. Methods and functions: Breath tests. Best Pract. Res. Clin. Gasteroenterol. 2009, 23, 337–352. [Google Scholar] [CrossRef]
  187. Bond, J.H.; Levitt, M.D. Quantitative measurement of lactose absorption. Gastroenterology 1976, 70, 1058–1062. [Google Scholar]
  188. Calloway, D.H.; Murphy, E.L.; Bauer, D. Determination of lactose intolerance by breath analysis. Am. J. Gastroenterol. 1969, 14, 811–815. [Google Scholar] [CrossRef]
  189. Sategna-Guidetti, C.; Cruto, E.; Capobianco, P. Breath hydrogen excretion after lactose and whole milk ingestion: A prospective comparison in lactase deficiency. J. Clin. Gastroenterol. 1989, 11, 287–289. [Google Scholar] [CrossRef] [PubMed]
  190. Abramowitz, A.; Granot, E.; Tamir, I.; Deckelbaum, R.J. Two-hour lactose breath hydrogen test. J. Pediatr. Gastroenterol. Nutr. 1986, 5, 130–133. [Google Scholar] [PubMed]
  191. Corazza, G.; Strocchi, A.; Sorge, M.; Benati, G.; Gasbarrini, G. Prevalence and consistency of low breath H2 excretion following lactulose ingestion. Dig. Dis. Sci. 1993, 38, 2010–2016. [Google Scholar] [PubMed]
  192. Szilagyi, A.; Cohen, A.; Vinokuroff, C.; Ahman, D.; Nathwani, U.; Yesovitch, S. Deadaption and readaptation with lactose, but no cross-adaptation to lactulose: A case of occult colonic bacterial adaptation. Can. J. Gastroenterol. 2004, 18, 677–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Hertzler, S.R.; Savaiano, D.A. Colonic adaptation to daily lactose feeding in lactose maldigesters reduces lactose intolerance. Am. J. Clin. Nutr. 1996, 64, 232–236. [Google Scholar] [CrossRef]
  194. de Lacy Costello, B.P.; Ledochowski, M.; Ratcliffe, N.M. The importance of methane breath testing: A review. J. Breath Res. 2013, 7, 024001. [Google Scholar] [CrossRef]
  195. Materacki, L.; Lee, S.M.; Laidler, P.; Yong, K.; Betteridge, F.; Murugiah, D.; Colleypriest, B. PWE-098 Is methane testing a useful adjunct to hydrogen breath testing? Gut 2018, 67, A167. [Google Scholar]
  196. Gao, F.; Wang, M.; Zhang, X.; Zhang, J.; Xue, Y.; Wan, H.; Wang, P. Simultaneous detection of hydrogen and methane in breath for the diagnosis of small intestinal bacterial overgrowth by fast gas chromatography. Anal. Methods 2018, 10, 4329–4338. [Google Scholar] [CrossRef]
  197. Vernia, P.; Cesarini, M.; de Carolis, A.; Vernia, F. Early hydrogen excretion peaks during breath tests. Small intestinal bacterial overgrowth or accelerated transit? Dig. Liver Dis. 2021, 53, 442–444. [Google Scholar] [CrossRef] [PubMed]
  198. Corazza, G.R.; Strocchi, A.; Gasbarrini, G. Fasting breath hydrogen in celiac disease. Gastroenterology 1987, 93, 53–58. [Google Scholar] [CrossRef]
  199. Kerlin, P.; Wong, L.; Harris, B.; Capra, S. Rice flour, breath hydrogen, and malabsorption. Gastroenterology 1984, 87, 578–585. [Google Scholar] [CrossRef]
  200. Levitt, M.D. Production and excretion of hydrogen gas in man. N. Engl. J. Med. 1969, 281, 122–127. [Google Scholar] [CrossRef] [PubMed]
  201. Braden, B.; Lembcke, B.; Kuker, W.; Caspary, W.F. 13C-breath tests: Current state of the art and future directions. Dig. Liver Dis. 2007, 39, 795–805. [Google Scholar] [CrossRef]
  202. Lacroix, M.; Mosora, F.; Pontus, M.; Lefebvre, P.; Luyckx, A.; Lopez-Habib, G. Glucose naturally labeled with carbon-13: Use for metabolic studies in man. Science 1973, 181, 445–446. [Google Scholar] [CrossRef]
  203. Modak, A.S. Stable isotope breath tests in clinical medicine: A review. J. Breath Res. 2007, 1, 014003. [Google Scholar] [CrossRef]
  204. Peek, R.M.; Blaser, M.J. Helicobacter pylori and gastrointestinal tract adenocarcinomas. Nat. Rev. Cancer 2002, 2, 28–37. [Google Scholar] [CrossRef] [PubMed]
  205. Hooi, J.K.Y.; Lai, W.Y.; Ng, W.K.; Suen, M.M.Y.; Underwood, F.E.; Tanyingoh, D.; Malfertheiner, P.; Graham, D.Y.; Wong, V.W.S.; Wu, J.C.Y. Global prevalence of Helicobacter pylori infection: Systematic review and meta-analysis. Gastroenterology 2017, 153, 420–429. [Google Scholar] [CrossRef] [Green Version]
  206. Nocon, M.; Kuhlmann, A.; Leodolter, A.; Roll, S.; Vauth, C.; Willich, S.N.; Greiner, W. Efficacy and cost-effectiveness of the 13C-urea breath test as the primary diagnostic investigation for the detection of Helicobacter pylori infection compared to invasive and non-invasive diagnostic tests. GMS Health Technol. Assess. 2009, 5. [Google Scholar] [CrossRef] [Green Version]
  207. Malfertheiner, P.; Megraud, F.; O’Morain, C.; Bazzoli, F.; El-Omar, E.; Graham, D.; Hunt, R.; Rokkas, T.; Vakil, N.; Kuipers, E.J. Current concepts in the management of Helicobacter pylori infection: The Maastricht III Consensus Report. Gut 2007, 56, 772–781. [Google Scholar] [CrossRef] [PubMed]
  208. Crowe, S.E. Helicobacter pylori infection. N. Engl. J. Med. 2019, 380, 1158–1165. [Google Scholar] [CrossRef]
  209. Nomura, A.; Stemmermann, G.N.; Chyou, P.H.; Kato, I.; Perez-Perez, G.I.; Blaser, M.J. Helicobacter pylori infection and gastric carcinoma among Japanese Americans in Hawaii. N. Engl. J. Med. 1991, 325, 1132–1136. [Google Scholar] [CrossRef]
  210. Parsonnet, J.; Friedman, G.D.; Vandersteen, D.P.; Chang, Y.; Vogelman, J.H.; Orentreich, N.; Sibley, R.K. Helicobacter pylori infection and the risk of gastric carcinoma. N. Engl. J. Med. 1991, 325, 1127–1131. [Google Scholar] [CrossRef] [PubMed]
  211. Graham, D.Y.; Evans JR, D.J.; Alpert, L.C.; Klein, P.D.; Evans, D.G.; Opekun, A.R.; Boutton, T.W. Campylobacter pylori detected noninvasively by the 13C-urea breath test. Lancet 1987, 329, 1174–1177. [Google Scholar] [CrossRef]
  212. Gatta, L.; Ricci, C.; Tampieri, A.; Osborn, J.; Perna, F.; Bernabucci, V.; Vaira, D. Accuracy of breath tests using low doses of 13C-urea to diagnose Helicobacter pylori infection: A randomised controlled trial. Gut 2006, 55, 457–462. [Google Scholar] [CrossRef]
  213. Leodolter, A.; Dominguez-Munoz, J.E.; Von Arnim, U.; Manes, G.; Malfertheiner, P. 13C-urea breath test for the diagnosis of Helicobacter pylori infection: A further simplification for clinical practice. Scand. J. Gastroenterol. 1998, 33, 267–270. [Google Scholar]
  214. Ellenrieder, V.; Glasbrenner, B.; Stoffels, C.; Weiler, S.; Bode, G.; Möller, P.; Adler, G. Qualitative and semi-quantitative value of a modified 13C-urea breath test for identification ofHelicobacter pyloriinfection. Eur. J. Gastroenterol. Hepatol. 1997, 9, 1085–1089. [Google Scholar] [CrossRef] [PubMed]
  215. Graham, D.Y.; Malaty, H.M.; Cole, R.A.; Martin, R.F.; Klein, P.D. Simplified 13C-urea breath test for detection of Helicobacter pylori infection. Am. J. Gastroenterol. 2001, 96, 1741–1745. [Google Scholar] [CrossRef]
  216. Klein, P.D.; Graham, D.Y. Minimum analysis requirements for the detection of Helicobacter pylori infection by the 13C-urea breath test. Am. J. Gastroenterol. 1993, 88, 1865–1869. [Google Scholar]
  217. Vaira, D.; Vakil, N. Blood, urine, stool, breath, money, and Helicobacter pylori. Gut 2001, 48, 287–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Gisbert, J.P.; Pajares, J.M. 13C-urea breath test in the diagnosis of Helicobacter pylori infection—A critical review. Aliment. Pharm. Ther. 2004, 20, 1001–1017. [Google Scholar] [CrossRef] [PubMed]
  219. O’Hara, S.; Kato, M.; Saito, M.; Fukuda, S.; Kato, C.; Hamada, S.; Nagashima, R.; Obara, K.; Suzuki, M.; Honda, H. Comparison between a new 13C-urea breath test, using a film-coated tablet, and the conventional 13C-urea breath test for the detection of Helicobacter pylori infection. J. Gastroenterol. 2004, 39, 621–628. [Google Scholar] [CrossRef]
  220. Calvet, X.; Sánchez-Delgado, J.; Montserrat, A.; Lario, S.; Ramírez-Lázaro, M.J.; Quesada, M.; Casalots, A.; Suárez, D.; Campo, R.; Brullet, E. Accuracy of diagnostic tests for Helicobacter pylori: A reappraisal. Clin. Infect. Dis. 2009, 48, 1385–1391. [Google Scholar] [CrossRef] [Green Version]
  221. Modak, A.S. Chapter 17—13C breath tests. In Breathborne Biomarkers and the Human Volatilome, 2nd ed.; Beauchamp, J., Davis, C., Pleil, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 273–287. [Google Scholar]
  222. Camilleri, M.; Chedid, V.; Ford, A.C.; Haruma, K.; Horowitz, M.; Jones, K.L.; Low, P.A.; Park, S.Y.; Parkman, H.P.; Stanghellini, V. Gastroparesis. Nat. Rev. Dis. Primers 2018, 4, 1–19. [Google Scholar] [CrossRef]
  223. Beaumont, W.; Osler, W. Experiments and Observations on the Gastric Juice and the Physiology of Digestion; Dover Publications, Inc.: Mineola, NY, USA, 1996. [Google Scholar]
  224. Scarpignato, C. Gastric emptying measurement in man. In Clinical Investigation of Gastric Function; Scarpignato, C., Bianchi Porro, G., Eds.; Karger Publishers: Basel, Switzerland, 1990; Volume 17, pp. 198–246. [Google Scholar]
  225. Heading, R.C. Gastric emptying: A clinical perspective. Clin. Sci. 1982, 63, 231–235. [Google Scholar] [CrossRef] [Green Version]
  226. Abell, T.L.; Camilleri, M.; Donohoe, K.; Hasler, W.L.; Lin, H.C.; Maurer, A.H.; McCallum, R.W.; Nowak, T.; Nusynowitz, M.L.; Parkman, H.P. Consensus recommendations for gastric emptying scintigraphy: A joint report of the American Neurogastroenterology and Motility Society and the Society of Nuclear Medicine. J. Nucl. Med. Technol. 2008, 36, 44–54. [Google Scholar] [CrossRef]
  227. Lin, H.C.; Prather, C.; Fisher, R.S.; Meyer, J.H.; Summers, R.W.; Pimentel, M.; Mccallum, R.W.; Akkermans, L.M.; Loening-Baucke, V. Measurement of gastrointestinal transit. Dig. Dis. Sci. 2005, 50, 989–1004. [Google Scholar] [CrossRef] [PubMed]
  228. DiBaise, J.K.; Park, F.L.; Lyden, E.; Brand, R.E.; Brand, R.M. Effects of low doses of erythromycin on the 13C Spirulina platensis gastric emptying breath test and electrogastrogram: A controlled study in healthy volunteers. Am. J. Gastroenterol. 2001, 96, 2041–2050. [Google Scholar] [CrossRef] [PubMed]
  229. Hauser, B.; Roelants, M.; De Schepper, J.; Veereman, G.; Caveliers, V.; Devreker, T.; De Greef, E.; Vandenplas, Y. Gastric emptying of solids in children: Reference values for the 13C-octanoic acid breath test. Neurogastroenterol. Motil. 2016, 28, 1480–1487. [Google Scholar] [CrossRef] [PubMed]
  230. Bruno, G.; Lopetuso, L.R.; Ianiro, G.; Laterza, L.; Gerardi, V.; Petito, V.; Poscia, A.; Gasbarrini, A.; Ojetti, V.; Scaldaferri, F. 13C-octanoic acid breath test to study gastric emptying time. Eur. Rev. Med. Pharmacol. Sci. 2013, 17, 59–64. [Google Scholar] [PubMed]
  231. Bertram, F.; Andresen, V.; Layer, P.; Keller, J. Simultaneous non-invasive measurement of liquid gastric emptying and small bowel transit by combined 13C-acetate and H2-lactulose breath test. J. Breath Res. 2014, 8, 046007. [Google Scholar] [CrossRef] [PubMed]
  232. Ghoos, Y.F.; Maes, B.D.; Geypens, B.J.; Mys, G.; Hiele, M.I.; Rutgeerts, P.J.; Vantrappen, G. Measurement of gastric emptying rate of solids by means of a carbon-labeled octanoic acid breath test. Gastroenterology 1993, 104, 1640–1647. [Google Scholar] [CrossRef]
  233. González, A.; Mugueta, C.; Parra, D.; Labayen, I.; Martinez, A.; Varo, N.; Monreal, I.; Gil, M.J. Characterisation with stable isotopes of the presence of a lag phase in the gastric emptying of liquids. Eur. J. Nutr. 2000, 39, 224–228. [Google Scholar] [CrossRef]
  234. Viramontes, B.E.; Kim, D.Y.; Camilleri, M.; Lee, J.S.; Stephens, D.; Burton, D.D.; Thomforde, G.M.; Klein, P.D.; Zinsmeister, A.R. Validation of a stable isotope gastric emptying test for normal, accelerated or delayed gastric emptying. Neurogastroenterol. Motil. 2001, 13, 567–574. [Google Scholar] [CrossRef]
  235. Lee, J.S.; Camilleri, M.; Zinsmeister, A.R.; Burton, D.D.; Kost, L.J.; Klein, P.D. A valid, accurate, office based non-radioactive test for gastric emptying of solids. Gut 2000, 46, 768–773. [Google Scholar] [CrossRef]
  236. Food and Drug Administration, U.S. Summary of Safety and Effectiveness Data (PMA P110015). Available online: https://www.accessdata.fda.gov/cdrh_docs/pdf11/P110015b.pdf (accessed on 13 July 2021).
  237. Ilan, Y. The assessment of liver function using breath tests. Aliment. Pharm. Ther. 2007, 26, 1293–1302. [Google Scholar] [CrossRef]
  238. Stockmann, M.; Lock, J.F.; Malinowski, M.; Niehues, S.M.; Seehofer, D.; Neuhaus, P. The LiMAx test: A new liver function test for predicting postoperative outcome in liver surgery. HPB 2010, 12, 139–146. [Google Scholar] [CrossRef] [Green Version]
  239. Stockmann, M.; Lock, J.F.; Riecke, B.; Heyne, K.; Martus, P.; Fricke, M.; Lehmann, S.; Niehues, S.M.; Schwabe, M.; Lemke, A.J.; et al. Prediction of postoperative outcome after hepatectomy with a new bedside test for maximal liver function capacity. Ann. Surg. 2009, 250, 119–125. [Google Scholar] [CrossRef]
  240. Sakka, S.G. Assessing liver function. Curr. Opin. Crit. Care 2007, 13, 207–214. [Google Scholar] [CrossRef]
  241. Schneider, P.D. Preoperative assessment of liver function. Surg. Clin. N. Am. 2004, 84, 355–373. [Google Scholar] [CrossRef]
  242. Lalazar, G.; Ilan, Y. Assessment of liver function in acute or chronic liver disease by the methacetin breath test: A tool for decision making in clinical hepatology. J. Breath Res. 2009, 3, 047001. [Google Scholar] [CrossRef]
  243. Festi, D.; Capodicasa, S.; Vestito, A.; Mazzella, G.; Roda, E.; Vitacolonna, E.; Petrolati, A.; Angelico, M.; Colecchia, A. Breath tests with stable isotopes: Have they a role in liver transplantation? Eur. Rev. Med. Pharmacol. Sci. 2004, 8, 55–58. [Google Scholar]
  244. Hepner, G.W.; Vesell, E.S. Assessment of aminopyrine metabolism in man by breath analysis after oral administration of 14C-aminopyrine: Effects of phenobarbital, disulfiram and portal cirrhosis. N. Engl. J. Med. 1974, 291, 1384–1388. [Google Scholar] [CrossRef]
  245. Burke, P.A.; Stack, J.A.; Wagner, D.; Lewis, D.W.; Jenkins, R.L.; Forse, R.A. L-[1-13C] Phenylalanine oxidation as a measure of hepatocyte functional capacity in end-stage liver disease. Am. J. Surg. 1997, 173, 270–273. [Google Scholar] [CrossRef]
  246. Watkins, P.B.; Murray, S.A.; Winkelman, L.G.; Heuman, D.M.; Wrighton, S.A.; Guzelian, P.S. Erythromycin breath test as an assay of glucocorticoid-inducible liver cytochromes P-450. Studies in rats and patients. J. Clin. Investig. 1989, 83, 688–697. [Google Scholar] [CrossRef] [Green Version]
  247. Saadeh, S.; Behrens, P.W.; Parsi, M.A.; Carey, W.D.; Connor, J.T.; Grealis, M.; Barnes, D.S. The utility of the 13C-galactose breath test as a measure of liver function. Aliment. Pharm. Ther. 2003, 18, 995–1002. [Google Scholar] [CrossRef]
  248. Rubin, T.M.; Heyne, K.; Luchterhand, A.; Bednarsch, J.; Vondran, F.W.; Polychronidis, G.; Malinowski, M.; Nikolic, A.; Tautenhahn, H.M.; Jara, M. Kinetic validation of the LiMAx test during 10 000 intravenous 13C-methacetin breath tests. J. Breath Res. 2017, 12, 016005. [Google Scholar] [CrossRef]
  249. Gould, O.; Ratcliffe, N.; Król, E.; de Lacy Costello, B. Breath analysis for detection of viral infection, the current position of the field. J. Breath Res. 2020, 14, 041001. [Google Scholar] [CrossRef]
  250. Giovannini, G.; Haick, H.; Garoli, D. Detecting COVID-19 from Breath: A Game Changer for a Big Challenge. ACS Sens. 2021, 6, 1408–1417. [Google Scholar] [CrossRef]
  251. Davis, C.E.; Schivo, M.; Kenyon, N.J. A breath of fresh air–the potential for COVID-19 breath diagnostics. EBioMedicine 2021, 63, 103183. [Google Scholar] [CrossRef]
  252. Chen, H.; Qi, X.; Ma, J.; Zhang, C.; Feng, H.; Yao, M. Breath-borne VOC biomarkers for COVID-19. MedRxiv 2020. [Google Scholar] [CrossRef]
  253. Grassin-Delyle, S.; Roquencourt, C.; Moine, P.; Saffroy, G.; Carn, S.; Heming, N.; Fleuriet, J.; Salvator, H.; Naline, E.; Couderc, L.J.; et al. Metabolomics of exhaled breath in critically ill COVID-19 patients: A pilot study. EBioMedicine 2021, 63, 103154. [Google Scholar] [CrossRef]
  254. Shan, B.; Broza, Y.Y.; Li, W.; Wang, Y.; Wu, S.; Liu, Z.; Wang, J.; Gui, S.; Wang, L.; Zhang, Z. Multiplexed nanomaterial-based sensor array for detection of COVID-19 in exhaled breath. ACS Nano 2020, 14, 12125–12132. [Google Scholar] [CrossRef]
  255. Sawano, M.; Takeshita, K.; Ohno, H.; Oka, H. RT-PCR diagnosis of COVID-19 from exhaled breath condensate: A clinical study. J. Breath Res. 2021, 15, 037103. [Google Scholar] [CrossRef]
  256. Fernández del Río, R.; O’Hara, M.E.; Holt, A.; Pemberton, P.; Shah, T.; Whitehouse, T.; Mayhew, C.A. Volatile Biomarkers in Breath Associated With Liver Cirrhosis—Comparisons of Pre- and Post-liver Transplant Breath Samples. EBioMedicine 2015, 2, 1243–1250. [Google Scholar] [CrossRef] [Green Version]
  257. Ferrandino, G.; Orf, I.; Smith, R.; Calcagno, M.; Thind, A.K.; Debiram-Beecham, I.; Williams, M.; Gandelman, O.; de Saedeleer, A.; Kibble, G.; et al. Breath Biopsy assessment of liver disease using an exogenous volatile organic compound—Toward improved detection of liver impairment. Clin. Transl. Gastroenterol. 2020, 11, 1–7. [Google Scholar] [CrossRef]
  258. O’Hara, M.E.; Fernández Del Río, R.; Holt, A.; Pemberton, P.; Shah, T.; Whitehouse, T.; Mayhew, C.A. Limonene in exhaled breath is elevated in hepatic encephalopathy. J. Breath Res. 2016, 10, 046010. [Google Scholar] [CrossRef]
  259. De Vries, R.; Dagelet, Y.W.F.; Spoor, P.; Snoey, E.; Jak, P.M.C.; Brinkman, P.; Dijkers, E.; Bootsma, S.K.; Elskamp, F.; De Jongh, F.H.C. Clinical and inflammatory phenotyping by breathomics in chronic airway diseases irrespective of the diagnostic label. Eur. Respir. J. 2018, 51, 1701817. [Google Scholar] [CrossRef] [Green Version]
  260. Schleich, F.N.; Zanella, D.; Stefanuto, P.H.; Bessonov, K.; Smolinska, A.; Dallinga, J.W.; Henket, M.; Paulus, V.; Guissard, F.; Graff, S. Exhaled volatile organic compounds are able to discriminate between neutrophilic and eosinophilic asthma. Am. J. Respir. Crit. Care Med. 2019, 200, 444–453. [Google Scholar] [CrossRef]
  261. Ahmed, W.; Brinkman, P.; Fowler, S. Clinical phenotyping. In Breathborne Biomarkers and the Human Volatilome; Elsevier: Amsterdam, The Netherlands, 2020; pp. 321–334. [Google Scholar]
  262. Henderson, B.; Ruszkiewicz, D.M.; Wilkinson, M.; Beauchamp, J.D.; Cristescu, S.M.; Fowler, S.J.; Salman, D.; Di Francesco, F.; Koppen, G.; Langejürgen, J. A benchmarking protocol for breath analysis: The peppermint experiment. J. Breath Res. 2020, 14, 046008. [Google Scholar] [CrossRef]
  263. Wilkinson, M.; White, I.; Hamshere, K.; Holz, O.; Schuchardt, S.; Bellagambi, F.G.; Lomonaco, T.; Biagini, D.; Di, F.F.; Fowler, S.J. The peppermint breath test: A benchmarking protocol for breath sampling and analysis using GC-MS. J. Breath Res. 2021, 15, 026006. [Google Scholar]
  264. Gisler, A.; Lan, J.; Singh, K.D.; Usemann, J.; Frey, U.; Zenobi, R.; Sinues, P. Real-time breath analysis of exhaled compounds upon peppermint oil ingestion by secondary electrospray ionization-high resolution mass spectrometry: Technical aspects. J. Breath Res. 2020, 14, 046001. [Google Scholar] [CrossRef]
  265. Lan, J.; Gisler, A.; Bruderer, T.; Sinues, P.; Zenobi, R. Monitoring peppermint washout in the breath metabolome by secondary electrospray ionization-high resolution mass spectrometry. J. Breath Res. 2021, 15, 026003. [Google Scholar] [CrossRef]
  266. Henderson, B.; Slingers, G.; Pedrotti, M.; Pugliese, G.; Ghimenti, S.; Moreno, S.; Harren, F.; Schubert, J.; Mayhew, C.A.; Wilde, M.; et al. The peppermint breath test benchmark for PTR-MS and SIFT-MS. J. Breath Res. 2021, 15, 046005. [Google Scholar] [CrossRef]
Figure 1. Exchange of respiratory gases—oxygen (O2) and carbon dioxide (CO2)—within the alveoli during aerobic cell respiration; the process of aerobic cell respiration and CO2 production is depicted on the right: pyruvate, generated from glycolysis of glucose, and O2 enter the Krebs cycle to form CO2 and produce the energy-intermediate adenosine triphosphate (ATP), as well as the nicotinamide adenine dinucleotide anion (NADH) and the hydroquinone form of flavin adenine dinucleotide (FADH2). During oxidative phosphorylation, electrons are transferred from NADH and FADH2 to O2 by a series of electron carriers to form ATP and water (H2O). Created with BioRender.com.
Figure 1. Exchange of respiratory gases—oxygen (O2) and carbon dioxide (CO2)—within the alveoli during aerobic cell respiration; the process of aerobic cell respiration and CO2 production is depicted on the right: pyruvate, generated from glycolysis of glucose, and O2 enter the Krebs cycle to form CO2 and produce the energy-intermediate adenosine triphosphate (ATP), as well as the nicotinamide adenine dinucleotide anion (NADH) and the hydroquinone form of flavin adenine dinucleotide (FADH2). During oxidative phosphorylation, electrons are transferred from NADH and FADH2 to O2 by a series of electron carriers to form ATP and water (H2O). Created with BioRender.com.
Molecules 26 05514 g001
Figure 2. Schematic depictions of (a) the respiration cycle as described in relation to carbon dioxide (CO2) partial pressure—referred to as a capnogram—comprising four transitional phases, namely: phase I, end of inspiration and dead-space gas; phase II, mixed-airway and alveolar gas; phase III, end-tidal volume; and phase IV, inspiration; note that CO2 is offset from the vertical axis for a clearer depiction (figure adapted from [50]); and (b) fraction of exhaled nitric oxide (FENO) during respiration through the mouth at a flow rate of 50 mL/s, with depicted phases representing: phase I, inspiration with air; phase II, mixed-expiratory gas, including the initial peak of exhaled NO after inhalation via the nose (from local nasal NO production); phase III, steady flow region (the 3 second plateau from which the FENO value is extracted is indicated); and phase IV, inspiration (figure adapted from [51]).
Figure 2. Schematic depictions of (a) the respiration cycle as described in relation to carbon dioxide (CO2) partial pressure—referred to as a capnogram—comprising four transitional phases, namely: phase I, end of inspiration and dead-space gas; phase II, mixed-airway and alveolar gas; phase III, end-tidal volume; and phase IV, inspiration; note that CO2 is offset from the vertical axis for a clearer depiction (figure adapted from [50]); and (b) fraction of exhaled nitric oxide (FENO) during respiration through the mouth at a flow rate of 50 mL/s, with depicted phases representing: phase I, inspiration with air; phase II, mixed-expiratory gas, including the initial peak of exhaled NO after inhalation via the nose (from local nasal NO production); phase III, steady flow region (the 3 second plateau from which the FENO value is extracted is indicated); and phase IV, inspiration (figure adapted from [51]).
Molecules 26 05514 g002
Figure 3. (a) Synthesis of nitric oxide (NO) from l-arginine in the presence of oxygen (O2), with nitric oxide synthases (NOS) and the cofactors reduced nicotinamide adenine dinucleotide phosphate (NADPH), flavin adenine dinucleotide (FAD), flavin mononucleotide (FMN), and tetrahydrobiopterin (BH4); (b) Heme degradation pathway: heme metabolism by heme oxygenase enzymes requires O2 and NADPH (with NADPH CYP450 reductase) with heme oxygenase-1 (HO-1) to generate equimolar carbon monoxide (CO), iron (Fe2+) and biliverdin, which is reduced to bilirubin by NAD(P)H biliverdin reductase.
Figure 3. (a) Synthesis of nitric oxide (NO) from l-arginine in the presence of oxygen (O2), with nitric oxide synthases (NOS) and the cofactors reduced nicotinamide adenine dinucleotide phosphate (NADPH), flavin adenine dinucleotide (FAD), flavin mononucleotide (FMN), and tetrahydrobiopterin (BH4); (b) Heme degradation pathway: heme metabolism by heme oxygenase enzymes requires O2 and NADPH (with NADPH CYP450 reductase) with heme oxygenase-1 (HO-1) to generate equimolar carbon monoxide (CO), iron (Fe2+) and biliverdin, which is reduced to bilirubin by NAD(P)H biliverdin reductase.
Molecules 26 05514 g003
Figure 4. The fraction of exhaled of nitric oxide (FENO) and NO output as a function of expiratory flow, whereby 50 mL/s is recommended as the standardised flow for the FENO breath test (specified as FENO50; see text). The FENO value is derived from the intercept of the regression lines for low and high NO output, as illustrated. (Figure adapted from [79]).
Figure 4. The fraction of exhaled of nitric oxide (FENO) and NO output as a function of expiratory flow, whereby 50 mL/s is recommended as the standardised flow for the FENO breath test (specified as FENO50; see text). The FENO value is derived from the intercept of the regression lines for low and high NO output, as illustrated. (Figure adapted from [79]).
Molecules 26 05514 g004
Figure 5. (a) Assessment of hypolactasia via the hydrogen breath test (HBT): colonic fermentation of ingested lactose after malabsorption leads to the production of molecular hydrogen (H2), carbon dioxide (CO2) and short-chain fatty acids (SCFA): the detection of H2 in exhaled breath is possible after its venous transport to the lungs; (b) Diagnosis of H. pylori infection via the 13C-urea breath test (UBT): orally ingested 13C-urea is metabolised by urease enzyme to form 13CO2 (and ammonia, NH3) in the stomach, which is transported to the lungs and can be detected in exhaled breath; (c) Stable isotope breath test for assessing solid gastric emptying using 13C-S. platensis: the ingested substrate is absorbed in the small intestine and metabolised to 13CO2 in the liver, which is transported to the lungs and can be detected in exhaled breath; (d) Assessing liver function capacity after intravenous injection of 13C-methacetin by continuous measurement of 13CO2/12CO2-ratio: metabolism of the substrate in the liver by CYP1A2 enzymes forms 13CO2, which can be detected in exhaled breath. Figures created with BioRender.com.
Figure 5. (a) Assessment of hypolactasia via the hydrogen breath test (HBT): colonic fermentation of ingested lactose after malabsorption leads to the production of molecular hydrogen (H2), carbon dioxide (CO2) and short-chain fatty acids (SCFA): the detection of H2 in exhaled breath is possible after its venous transport to the lungs; (b) Diagnosis of H. pylori infection via the 13C-urea breath test (UBT): orally ingested 13C-urea is metabolised by urease enzyme to form 13CO2 (and ammonia, NH3) in the stomach, which is transported to the lungs and can be detected in exhaled breath; (c) Stable isotope breath test for assessing solid gastric emptying using 13C-S. platensis: the ingested substrate is absorbed in the small intestine and metabolised to 13CO2 in the liver, which is transported to the lungs and can be detected in exhaled breath; (d) Assessing liver function capacity after intravenous injection of 13C-methacetin by continuous measurement of 13CO2/12CO2-ratio: metabolism of the substrate in the liver by CYP1A2 enzymes forms 13CO2, which can be detected in exhaled breath. Figures created with BioRender.com.
Molecules 26 05514 g005
Figure 6. Alcohol (ethanol) disposition in the body. Only a small fraction (<10%) is eliminated unchanged through exhaled breath, perspiration and urinary excretion. Created with BioRender.com.
Figure 6. Alcohol (ethanol) disposition in the body. Only a small fraction (<10%) is eliminated unchanged through exhaled breath, perspiration and urinary excretion. Created with BioRender.com.
Molecules 26 05514 g006
Table 1. Current established/approved breath-based tests that utilise endogenous or exogenous gas-phase compounds.
Table 1. Current established/approved breath-based tests that utilise endogenous or exogenous gas-phase compounds.
Application or Disease TargetTest NameTarget CompoundDetection MethodUnit of MeasurementDeterminant and Setting * of Use
Endogenous compoundsVentilation/breathingCapnographyCarbon dioxide (CO2)Colorimetric CO2 detector, mainstream and sidestream CO2 monitoring (IR spectroscopy)mmHg Routine, clinical
AsthmaFENO(Fraction of exhaled) nitric oxide (NO)Chemiluminescence analyser, electrochemical sensors and laser-based technologyppbSymptomatic, clinical/surgery
Neonatal jaundiceCOCarbon monoxide (CO)CO monitor with integrated IR breathing sensor and electrochemical sensorppmSymptomatic, clinical
Grade 3 heart transplant rejectionHeartsbreathAlkanesTD-GC-MSBreath methylated alkane contourTargeted, clinical
Exogenous compoundsAlcohol intakeBreath alcohol test (Breathalyser)Ethanol (CH3CH2OH)IR spectroscopy, electrochemical fuel cells, dual sensor devices (electrochemical oxidation and IR absorption) mg/L, ‰Targeted, mobile
Lactase deficiencyHydrogen breath testHydrogen (H2)Hydrogen breath analyser with integrated electrochemical gas sensorppm Symptomatic/targeted, surgery
Helicobacter pylori infectionUrea breath test (UBT)13CO2Isotope ratio mass spectrometry ppm, ‰Symptomatic/targeted, surgery
GastroparesisGastric emptying breath test (GEBT)13CO2Isotope ratio mass spectrometryµmol/L/min Symptomatic/targeted, surgery
Liver functionMaximum liver function capacity (LiMAx)13CO2Isotope ratio mass spectrometry µg/kg/hTargeted, surgery
* Settings referred to are clinical, i.e., in-patients in hospitals or medical centres; surgery, i.e., ambulant patients in the physician’s practice (or medical centres); mobile, i.e., in the field or in specific settings (only applicable for the breath alcohol test). Abbreviations: IR = infrared; TD-GC-MS = thermal desorption-gas chromatography-mass spectrometry.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pham, Y.L.; Beauchamp, J. Breath Biomarkers in Diagnostic Applications. Molecules 2021, 26, 5514. https://doi.org/10.3390/molecules26185514

AMA Style

Pham YL, Beauchamp J. Breath Biomarkers in Diagnostic Applications. Molecules. 2021; 26(18):5514. https://doi.org/10.3390/molecules26185514

Chicago/Turabian Style

Pham, Y Lan, and Jonathan Beauchamp. 2021. "Breath Biomarkers in Diagnostic Applications" Molecules 26, no. 18: 5514. https://doi.org/10.3390/molecules26185514

Article Metrics

Back to TopTop