Nitrous oxide abuse direct measurement for diagnosis and follow-up: update on kinetics and impact on metabolic pathways
-
Angèle Lucas
, Guillaume Garcon
, Sébastien Antherieu
Abstract
Recreational use of nitrous oxide (N2O) has become a major health issue worldwide, with a high number of clinical events, especially in neurology and cardiology. It is essential to be able to detect and monitor N2O abuse to provide effective care and follow-up to these patients. Current recommendations for detecting N2O in cases of recreational misuse and consumption markers are lacking. We aimed to update current knowledge through a review of the literature on N2O measurement and kinetics. We reviewed the outcomes of experiments, whether in preclinical models (in vitro or in vivo), or in humans, with the aim to identify biomarkers of intoxication as well as biomarkers of clinical severity, for laboratory use. Because N2O is eliminated 5 min after inhalation, measuring it in exhaled air is of no value. Many studies have found that urine and blood matrices concentrations are connected to ambient concentrations, but there is no similar data for direct exposure. There have been no studies on N2O measurement in direct consumers. Currently, patients actively abusing N2O are monitored using effect biomarkers (biomarkers related to the effects of N2O on metabolism), such as vitamin B12, homocysteine and methylmalonic acid.
Introduction
Nitrous oxide (N2O), since its discovery in 1772, has been mainly used for anesthesia and analgesia. As the least potent inhalational anesthetic [1], N2O is frequently combined with other agents such as oxygen. It can also be found in supermarkets for use as an agent in whipped cream dispensers. However, this form can also be used recreationally as ‘laughing gas’. Its wide availability in many countries explains why recreational use of N2O is becoming a major health issue. The Global Drug Survey (n=32,022) listed N2O as the 14th most commonly used recreational drug worldwide in 2021 [2]. The widespread recreational consumption of N2O globally, particularly attributed to its legal status (with notable exceptions like the United Kingdom and the Netherlands), and its accessibility, presents a unique public health concern. Unlike many psychoactive substances, N2O remains legal in numerous jurisdictions, largely due to its established medical applications and commercial uses in food preparation. This legality contributes to its widespread availability, making it easily accessible to the general public. Furthermore, the relatively low cost and ease of procurement, often via online platforms or in commercial outlets, in the form of whipped cream chargers, facilitate its misuse. The perception of N2O as a ‘legal high’ diminishes awareness of its potential health risks, leading to its increased popularity among recreational users. This situation is compounded by the challenge of regulating a substance that has a role in various legitimate industries, creating a complex scenario for public health authorities in balancing the substance’s legitimate uses against its potential for misuse.
According to the European Monitoring Centre for Drugs and Drug Addiction, recreational use of N2O is increasingly observed in the general population [3]. Data from different countries was gathered in a review that was released in 2022 in order to assess the growth of N2O use. In the Western world, N2O use is becoming increasingly common [4]. Epidemiological data show that it is used by mainly young people aged 18–25 years old. A survey by the French Monitoring Center for Addiction network showed that N2O is consumed more by men, which is consistent with an American study [5, 6]. Until recently whipped cream cartridges were used recreationally, but now it is consumed in much larger quantities via canisters (Figure 1).

Consumption of nitrous oxide. The containers used for nitrous oxide recreational use have changed throughout time. Originally used for whipped cream siphons, small cartridges (on the left) have been gradually supplanted by canisters (middle), which hold 80 to 100 cartridges each. There are also tanks (on the right), which hold 400–600 cartridges each.
The clinical effects of recreational N2O consumption have been widely reported for both the acute and chronic effects. Short-term adverse events may include digestive problems such as nausea or vomiting, dizziness, psychomotor retardation and ataxia, as well as cold burns if the skin comes into prolonged contact with the canisters. In the long term, neurological damage has been described, with subacute combined degeneration of the spinal cord and/or peripheral sensory and motor neuropathy, leading to sensory loss, limb weakness, ataxia, bladder and bowel disturbance and sexual dysfunction [7], [8], [9]. Central nervous disorders have also been reported, notably encephalopathy [10]. Several cases of thrombosis have also been reported, likely attributable to elevation of homocysteine level [11], [12], [13].
With regards to the risk of neurological harm and to the limited understanding of long-term outcomes, further research and standardized clinical care are needed [14]. In addition, N2O has been increasingly implicated in some medico-legal matters, such as car accidents, especially if consumed while driving, and healthcare workers who develop symptoms after exposure [3]. It is therefore important to be able to biologically monitor N2O consumption in selected individuals.
The kinetic properties of N2O have been explored. The physical features of distribution between different media and tissues have been studied in vitro and in animal models. Because past exposure was primarily environmental, the main aim of clinical studies was to find a method of measuring this exposure. As a result, several investigations have focused on sampling urine. We first aimed to review the various technologies available for N2O measurement, to establish whether these methods are suitable for routine use in medical laboratories. Then, we reviewed available toxicokinetic data in the literature regarding absorption, distribution, metabolism, and elimination in order to evaluate biological diagnosis and monitoring of recreational use of N2O.
The analytical methods currently able to measure N2O will be discussed in order to validate analytical measurement of nitrous oxide. Secondly, we will discuss the biological kinetics of N2O after consumption, to determine the optimum choice of biological medium for its measurement. Finally, we will focus on the consequences on metabolism, so as to be able to propose possible biological markers of interest. To carry out this literature review, we used the PubMed and Medline databases, using keywords related to N2O and measurement methods. We excluded most studies related to the environment (non-biological studies), but retained studies involving animal exposure, given the small number of human studies available.
Pharmacological effects
N2O is known for its analgesic, anesthetic, anxiolytic, and antidepressant properties. These pharmacological effects depend on N2O concentration [15]. N2O is the least potent among the inhaled anesthetic and its effect involves the non-competitive inhibition of the NMDA receptors [16]. This blockage also partly explains N2O’s analgesic properties. The analgesic effect depends on the activation of opioid receptors, especially kappa opioid receptor [17], and to a lesser extent adrenoreceptors (α1 and α2B) activation. The anxiolytic effect of N2O entails activation of the GABAergic system, in particular the GABA type A receptor, which is also involved in the action of benzodiazepines, major anxiolytics. The antidepressant potential of N2O is much less known and more recently described. It is comparable to ketamine as similarly brief, and mediated via non-competitive inhibition of NMDA receptors [18, 19].
Measurements methods
There are several methods for measuring N2O, but most are used for environmental measurements to estimate global warming and ozone depletion, making them difficult to transpose to the medical laboratory. There are few publications focusing on measurement in biological matrices from organisms such as humans. Here, we summarize the methods that may be of potential interest to the medical laboratory, and have not focused on methods that cannot be applied to clinical chemistry, such as chamber methods for measuring gas flow [20], air flow measurement methods [21], [22], [23] and optical methods [24, 25], which are better suited for environmental applications. Methods of potential interest to laboratory medicine are chromatography and infrared spectroscopy.
Gas chromatography-mass spectrometry
Most N2O analytical methods rely on gas chromatographic techniques. Environmental studies have shown that gas chromatography-isotope ratio mass spectrometry [26] and gas chromatography-mass spectrometry (GC-MS) [27], [28], [29] are used. These mass spectrometry-based approaches have also been used in other fields of study, such as investigation of biological pathways [30, 31]. GC coupled with a thermal conductivity detector (TC) and a micro-ionization cross-section detector (MICS) are alternative methods for measuring N2O [32]. Methodological shortcomings include limited sensitivity and the use of internal standards that are not optimally selective at the molecular level for N2O measurement. Due to the lack of acceptable internal standards, external calibration is required, which is not the best strategy for gas analysis given the potential risk of leaks during sampling, extraction, and analysis [33, 34]. Even though previous published GC-MS techniques have improved sensitivity, the use of pre-loaded hermetic gas syringes should alleviate any quantification difficulties caused by probable leakage.
A variant of GC-MS called Headspace-GC-MS (HS-GC-MS) is a method also used to measure N2O. The sample is placed in a hermetically sealed container for HS-GCMS analysis. To create a headspace, the sample is heated to boiling point. The GC injection needle draws the gaseous substance into the headspace and injects it into the chromatographic column. In 2009, Poli et al. used HS-GC-MS to measure N2O in numerous tissues on patients who had been deceased for several days [35]. During autopsy, blood, liver, bile, kidney, fat, and brain samples were collected; a urine sample was also taken from the eighth deceased patient [35]. All of the tissue samples were immediately weighed and then transferred to 10-mL headspace (HS) vials with airtight plugs. The gaseous samples from the sample bags and syringe were also transferred to HS vials, which were completely filled [35]. In, 2015, Giuliani et al. suggest a new HS-GC-MS method for the quantification of N2O by using hydrogen sulfide as internal standard [36]. Although the authors validated this method, no independent study incorporated HS-GC-MS as their method, and its implementation in medical laboratories appears challenging due to its preparation difficulty and lack of reproducibility. Hence, use of HS-GC-MS for N2O measurement could be promising but requires further technical development and clinical validation studies for use in laboratory medicine. A few medical laboratories already employ HS-GC-MS for measurement of alcohols and glycols, so these laboratories that have the equipment and expertise could certainly investigate measuring and optimizing N2O by HS-GC-MS without significant cost [37].
Infrared spectroscopy
Infrared spectroscopy (IR), based on the molecule’s ability to absorb light at a specific wavelength, is another technique for measuring N2O. As described by Heusler et al., IR techniques are ideal to measure in air, but not in biological fluids such as blood or urine, where GC-MS is the preferred choice [38]. Closed path methods are preferred with greater sensitivity. The advantage of open path systems is that the concentration can be measured continuously, which can be useful when using N2O for anesthesia to control flow rate or to prevent occupational exposure [39], [40], [41]. However, interest in using IR technology for measuring direct unique or recreational N2O consumption is more limited [42].
Fourier transform infrared spectrometers (FTIR) are well described. This technique uses broadband IR radiation covering the entire IR spectrum, enabling many gases to be analysed at the same time with high precision, which explains why this technique is so popular for atmospheric measurements [25, 43, 44]. Another IR technique that can be used to measure N2O is infrared laser-absorption spectroscopy [45]. In contrast to FTIR spectroscopy, which measures the complete spectrum, laser spectroscopy uses a single wavelength to boost sensitivity. Specificity can be enhanced by working at low pressure, which results in narrower lines [42].
We have not found data on the use of IR methods for the determination of N2O in biological fluids other than exhaled air.
Exposure monitoring: toxicokinetic of N2O
N2O assay methods are available. However, identifying it in biological matrices necessitates a thorough understanding of its toxicokinetic properties in order to know which matrices to look for and when to look for it after intake. The four phases of toxicokinetic are studied: absorption, distribution, metabolism and elimination. In order to study N2O, it is necessary to know the quantities absorbed by the body in relation to the quantities taken in by consumers. The study of distribution is essential in order to know in which tissue and/or fluid N2O can be easily found, and in a quantity large enough to be measured. As an exogenous substance can be transformed in the body by detoxifying metabolic systems, knowing the metabolites of N2O may be of interest in finding markers of exposure other than N2O itself. Finally, an understanding of the elimination pathways is crucial to find the right fluids for N2O assay, and the timescales within which we can carry out the assay.
Absorption
Consumption-dependent quantity
There are two types of N2O exposure: therapeutic and recreational. In cases of analgesia without loss of consciousness, an equimolar mixture of oxygen and N2O (EMONO) is utilized or diluted in oxygen at concentrations of 50–70 % [46]. The anesthesiologist adapts the flow rate to the patient ventilatory rate, but this practice is waning.
The average flow rate recommended according to an SFAR study (French Society of Anesthesiologists and Reanimators) was 1 L/min or less. Thus, patient receives N2O at a rate of about 500 mL/min [47], or at 0.9 g/min of N2O since the gaseous form has a density of 0.0018 g/mL (Figure 2). EMONO is still frequently used in dental surgery, particularly on young patients. The majority of treatments run no longer than 40 min, and the N2O concentration used ranges from 20 to 50 %, depending on the complexity of the intervention [48], [49], [50].

Summary of global toxicokinetic of nitrous oxide. Nitrous oxide is absorbed by the respiratory tract alone. It is then distributed to all tissues. Due to its low solubility, the body is rapidly saturated. It is not metabolized, and is rapidly eliminated in unchanged form, mainly in exhaled air. A fraction is eliminated via the urine, and a small proportion via the transdermal route.
The levels of N2O used for recreational purposes are difficult to quantify since they differ depending on the consumer and the mode of consumption (Figure 1). A canister holds 8 g of N2O, while a cylinder holds 640 g. Several sources claim that casual N2O users only consume a few dozen canisters per occasion (equal to about 100 g of N2O). However, recreational use of N2O is on the rise. Many people now use N2O in cylinders rather than canisters, with some people consuming numerous cylinders per day (Figure 1). Some claim to have consumed up to 25 cylinders per day (equal to 16 kg of N2O) (Figure 2). As a result, recreational exposures can be substantially higher than medical doses.
Uptake
The absorption of a gas is the result of the balance between the concentration inhaled and concentration in the alveoli, blood and tissues. Because N2O has relatively low solubility in blood (and other tissues), it equilibrates very quickly through tissue saturation. Due to its poor solubility in blood and tissues, the increase in alveolar concentration upon exposure is very quick, resulting in rapid saturation of the body. After 5 min of treatment, the alveolar fraction of N2O equals 90 % of the fraction inhaled [51]. The rate of increase in alveolar concentration is influenced by inhaled concentration. The greater the inhaled concentration, the faster the alveolar concentration rises [52]. Uptake is altered in cases of pulmonary disease such as chronic obstructive pulmonary disease (COPD) [53]. This has been demonstrated in patients undergoing surgery and receiving N2O. They were classified into three groups according to the severity of their respiratory impairment, on the basis of the Tiffeneau ratio. When comparing patients with COPD to healthy patients, concentration measurements in inspired and exhaled air revealed significative differences, especially with regard to equilibration time [53].
Distribution
In order to determine the quantity of N2O in blood, for consumer screening for example, it is essential to study the distribution of the gas in the body. It is important to understand the distribution kinetics in the various tissues to determine the best timing for N2O detection into the blood.
Partition coefficients
Because studying human tissue is highly challenging, extrapolations are frequently made from findings made during pre-clinical studies on animals or in vitro models. Tissue/gas distribution coefficients of N2O and cyclopropane were determined in rabbits and compared to the distribution coefficients of these agents in homogenates of identical in vitro tissues [54]. Rabbits anesthetized by parenteral injection of sodium pentobarbitone were used and received N2O mixed with oxygen (70/30). In order to maintain a constant arterial partial pressure, exhaled gas samples were evaluated by GC during the experiment, and the inhaled concentration was adjusted accordingly. Blood, brain, heart, kidney, liver and skeletal muscle were extracted. Equilibrating homogenates of the same tissues (unused in vivo samples where any remained) with a known partial pressure of N2O yielded the in vitro coefficients. A discrepancy of 6 % was detected between than in vitro, higher, and in vivo distribution coefficients. For the six animals used, a three-way cross classification analysis of variance supported the fact that this difference is statistically significant, but the authors pointed out that this difference, although real, still allows us to have confidence in the in vitro distribution coefficients in order to extrapolate to humans.
Kety et al. conducted in vitro and in vivo experiments in dogs to evaluate the solubility of N2O in whole blood and brain as well as the partition coefficient between the two. Human and dogs appear to have equivalent values [55]. In human whole blood, they found a solubility of 0.412. This solubility explains how quickly the alveolar concentration rises to the inspired concentration. The solubility in human entire brain is 0.437, giving a brain/blood partition coefficient of 1.06 (Figure 2).
As a result of lower tissue/blood partition coefficients, the equilibration of N20 in most tissues occurs rapidly.
Blood distribution
N2O does not bind to any protein and circulates freely in the blood. Anesthetists, surgeons and nurses were subjected to an occupational exposure study to determine their blood levels of N2O. For all processes, the N2O flow rate was standardized at 6 L/min. Except for surgeons, from whom collections were taken at the end of the operation, blood samples (blood gas syringes) were collected every hour. Simultaneously, an air sample was taken from the subject’s breathing zone for 60 s. A correlation was showed between blood and breathing zone atmosphere concentrations (n=116) but for some samples (32 %, n=37), blood concentration could not be accurately matched to atmospheric concentration [56]. Nonetheless, this study holds additional significance because they checked the N2O blood levels of subjects before exposure, which was undetectable in any of the groups. It leads to the conclusion that the N2O blood level following exposure and its correlation to the atmospheric concentration are solely due to current exposure. A quantitative study comparing variance in blood and atmospheric concentrations revealed more stable blood concentrations. This shows that this is the result of blood N2O saturation rather than a single exposure [56].
In subjects receiving N2O directly for dental analgesia, N2O was detected in the blood after 10 min of inhalation at concentrations 2 times lower than those measured in the nasal mask and up to 5 times lower than the concentration set on the machine [57]. After 5 min of 100 % oxygen inhalation, venous concentration divided by 2.
In 2010, after several patients’ death from hospital poisoning with N2O, a team of researchers showed that N2O was still present in their blood one month later [35]. These patients had mistakenly received N2O instead of oxygen for periods ranging from 25 to 125 min. During the autopsy (one month after death), various samples were taken, including blood, and N2O concentration measured by HS-GC-MS. Large quantities of N2O were still measurable 1 month after death (11.29–2,152.04 mg/L). However, measurements taken so long after death are questionable since the tissues could have released N2O stored in the blood due to post-mortem autolysis and falsely elevated the measurements.
Involvement for laboratory medicine
Data on N2O blood distribution are lacking in direct consumers of recreational dose(s). Recreationally absorbed doses are not equivalent to those absorbed by professionals exposed in hospitals. Chronic consumption of large doses has not been studied. The absence of detection in professionals just prior to experimental exposure suggests that, with regular exposure to low environmental doses, N2O is rapidly distributed and eliminated. Measuring it in the context of a diagnosis would therefore appear to be uninformative, due to the possibility of false negatives related to rapid disappearance from the blood. Again, these data are related to occupational exposure and the findings could be different for recreational consumption (Figure 2).
Metabolism
Xenobiotics are generally metabolized, mainly in the liver, to produce metabolites. If the substance is rapidly metabolized, it is challenging to detect it. On the other hand, if the substance is completely metabolized, the products could be of interest for exposure study.
Metabolic transformations
The CNESST (Commission on standards, equity, health and safety at work) sheet on N2O states that it is poorly metabolized in the liver, as the liver extracts only a tiny fraction of quantity received (Figure 2).
Because of its redox characteristics, N2O can be reduced by a variety of substances found in the body. This was demonstrated in 1952 when bacteria were exposed to an anaerobic, N2O-containing environment [58]. Pressure changes have been reported as a result of the transition of N2O into N2. The enzymes involved are adaptable and bacteria that have evolved to nitrates and nitrites have also adapted to N2O. As a result, N2O is a common intermediate in denitrification. In 1985, another experiment studied the denitrification by bacteria. The catalyzed equation is: N2O + 2e− + 2H+ → N2 + H2O [59].
Because bacteria may change N2O, a small amount of N2O can be digested by the microbiota. Rats (male and female) were sacrificed to recover a portion of gut and large intestine. The intestinal contents and wall were separated, and two of them received antibiotics (tetracycline, neomycin and bacitracin) for three days prior to sacrifice. Human intestinal contents were also studied at the same time by collecting faeces samples from one of the researchers. To avoid contamination from the intestinal bolus, the wall was cleaned. After diluting, samples were incubated with radiolabeled pure N2O or a mixture of N2O/O2 at various doses. The intestinal wall was incubated for 3 h and the contents for 16 h at 37 °C. The radiolabeled N2 generated by the reaction was analysed by mass spectrometry and quantified using an internal N2 standard. These findings demonstrated that human and rat intestinal contents metabolize N2O, as labelled N2 concentrations were measured in incubated samples [60]. At concentration of 10 and 20 %, and to a lesser extent at 5 %, oxygen inhibits this process. The intestinal wall samples had a very low quantity of N2, supporting the fact that metabolism is carried out by the intestinal microbiota [60].
Elimination
Data on N2O elimination may permit determining best matrix for consumption detection. Active substances could be eliminated from the body unchanged, inactivated, or in both forms, and in variable proportions. Regarding its metabolism, N2O is eliminated unchanged, as a tiny fraction seems to be metabolized. Elimination studies have focused primarily on exhaled air, urine and to a lesser extent skin.
Pulmonary elimination
As described in the absorption/distribution section, elimination of gases depends on solubility and partition coefficient. The blood/gas and tissue/gas coefficients reflect how readily some tissues can store xenobiotics. On the other hand, N2O is eliminated extremely rapidly. When N2O administration is stopped, the significant difference in concentration between the alveoli and the blood results in a rapid decrease in N2O concentration. Salanitre et al. demonstrated in 1962 that the uptake and elimination curves are comparable [61]. The aim of their study was to validate or refute the equivalence between N2O uptake and excretion. N2O was administered to five subjects at subanesthetic doses (10 % N2O/90 % O2), and the amount absorbed and eliminated was measured. A good correlation between N2O uptake and excretion was found, although for some patient curves not (but it could be attributable to the method) [61]. The excretion curves data were compared to a previous study with anesthetized patients. The only difference was an upward displacement of the curve in the anesthetized group, which might be explained by anatomic and physiological differences between the two groups, as well as a higher saturation in N2O in the anesthetized group.
The concept of diffusion hypoxia or diffusion anoxia should not be underestimated. The diffusion hypoxia is caused by the fact that N2O returns to the alveoli faster than nitrogen leaves them leading to a dilution of the oxygen concentration in the alveoli. To simulate gas exchange in the pulmonary alveolus, an in vitro experiment was performed. Because the solubility of N2O in blood is similar to that of water, water could represent blood. Air represents the pulmonary alveolus. A flask containing 50 % water and 50 % air is saturated with N2O. N2O is 35 times more soluble in water than nitrogen, and 20 times more soluble than oxygen. At equilibrium, N2O partial pressures in water and air are equivalent. When the flask is brought into contact with air, the equilibrium of partial pressures is disturbed. The partial pressure in air is lower than that in water. As a result, N2O diffuses from the water into the air, diluting the oxygen and nitrogen in the air [62]. This explains the hypoxia that may result from N2O inhalation during the reoxygenation period.
In 1990, an experimental investigation was conducted to evaluate biological exposure to N2O in occupational exposed subjects (surgeons, anesthetists and nurses). Exhaled air and urine were obtained from 20 participants, working in five operating rooms, as well as samples of the ambient air in the operating room. A correlation (r=0.911) was found between concentrations in air and exhaled air (r=0.911) [63]. These subjects were not directly exposed to N2O. Einarsson et al., investigated the lung clearance rate of N2O in surgical patients in 1993 [64]. Thirty-six subjects were divided into six groups based on the duration of their N2O exposure (30, 60 and 120 min) and the type of ventilation after this exposure (normo- or hypoventilation). End-tidal N2O concentrations were evaluated using infrared spectroscopy and inhaled concentrations of 30 % oxygen and 70 % N2O. During normal ventilation, the concentration of N2O dropped rapidly after the end of the exposure, from 66–70 % to 6–9 % at 5 min and 2–4 % at 30 min [64] (Table 1) (Figure 2). It decreased a bit more slowly during hypoventilation (16–23 % at 5 min and 5–9 % at 30 min).
Nitrous oxide determination in biological fluids after exposure.
Study | Number of subjects | Exposure dose | Exposure time | Expired air | Blood (time) | Urine (time) |
---|---|---|---|---|---|---|
Einnarson et al. [64], 1993 | 18 | 68–70 % | 0.5, 1 or 2 h | 66–70 % (0 h) | NA | NA |
6–9 % (5 min) | ||||||
2–4 % (30 min) | ||||||
Brugnone et al. [66], 1996 | Urine=76/blood=80 | 49 ± 45 μg/L | NA | NA | 23 ± 18 μg/L (0 h) | 26 ± 19 μg/L (0 h) |
Urine=23/blood=27 | NA | NA | 0.9 μg/L (18 h) | 1.5 μg/L (18 h) | ||
Zaffina et al. [68], 2019 | 21 | 24.17 ppm | Minimum 4 h | NA | NA | 9.04 μg/L (0 h) |
Henderson et al. [113], 2003 | 46 | 313 ± 358 ppm | 4 h | NA | NA | 114 ± 191 μg/L (0 h) |
Imbriani et al. [114], 1995 | 835 | 75.5 ± 3.9 ppm | 4 h | NA | NA | 45.18 ± 3.44 μg/L (0 h) |
Accorsi et al. [115], 2003 | 121 | 6.5 ppm | NA | NA | NA | 9 μg/L (0 h) |
Krapez et al. [56], 1980 | 34 | 57 ppm | NA | NA | 83.6 μg/L (0 h) | NA |
Furthermore, there was no effect of exposure duration on the rate of removal, indicating that the organism had reached maximum saturation after 30 min. There is no evidence of difference in the decline in N2O rate according to age [65]. After 30 min of N2O inhalation, the levels and rates of exhalation are not significantly different between elderly males (63–69 yo) and young adult males (22–30 yo). The rate of elimination therefore appears to be similar in all subjects, independent of age.
Urinary excretion
Numerous studies have evaluated urinary N2O content following exposure, particularly in occupationally-exposed patients, because urine is an easy-to-recover matrix in medical staff. In 1995, urine and blood samples of patients working in nine hospital operating rooms were collected at the end of a professional exposure session. Personal passive samplers were used to monitor atmospheric N2O in the operating room (n=80) and then compared to urine and blood levels in relevant participants (Table 1). A control group of urine (n=25) and blood (n=27) morning samples of blood donors were used to assess exposure due to atmospheric N2O levels. This data revealed a significant correlation between both urinary and blood levels with environmental N2O concentrations [66]. Finally, when compared to blood donors, the occupationally exposed group had significantly higher values. Particularly high and unexpected values were found in some patients (n=10). In this study, those extreme values were associated to participants who had a bacterial urine infection, even though they were asymptomatic. Another study found that urinary tract infections had an effect on N2O levels in urine [67]. To study the kinetics of N2O production, a healthy donor’s urine was inoculated with Enterobacter aerogenes. Six vials were infected and incubated at 37 °C. The level of N2O was measured at 0, 2, 5, 8, 24 and 48 h after inoculation. Gram negative bacilli produce high amounts of N2O (78.5–464.7 μg/L) compared to others bacteria (0.1–4.2 μg/L) [67]. The production is time dependent and follows exponential kinetics (it is low for the first 5 h, increases exponentially up to 24–36 h, and the equilibration is reaching at 48 h) [67]. N2O concentrations measured in urine infected with Gram-positive cocci and yeasts were not significantly different from the control sample. To study the correlation between N2O concentration and the different bacterial strain, 16 urine samples from patients were collected. After acidification and storage of urine at 4 °C, N2O concentration was measured within 24 h by headspace GC. There was a strong correlation between the bacteria involved and the level of N2O in urine (high for Gram negative bacteria, lower for Gram positive cocci and almost non-existent for sterile samples). Because N2O is most likely produced in the bladder, four urine samples were analysed immediately after collection and varying quantities of N2O were detected, which could be linked to the time of infection onset [67]. This means that the amount of N2O measured in the urine may be greater the longer the infection has been present. The levels produced by bacteria in cases of urinary tract infection are well above those found in cases of occupational exposure to N2O. The average urinary N2O concentration in subjects working in the operating theatre was 26 μg/L [66]. As a comparison, in 2018, N2O level was measured in urine of a cohort of dentists and dental assistants exposed to N2O. Concentrations ranged from 2.54 to 25.4 μg/L [68]. It is therefore difficult to differentiate between exposure and urinary tract infection when N2O is detected in urine (Figure 2).
In 1983, volunteers were exposed to N2O (850 ± 25 ppm) in a 20 m3 chamber [69]. The exposure was intermittent, lasting 5–15 min and totaling 20–60 min. The last exposure was followed by a 30 min in N2O-free air. Participants were equipped with in a gastight container. At the end of the experiment, urine was discharged into a syringe and was equilibrated with an equivalent volume of N2O-free air. The syringe was shaken in a water bath at 38 °C for 20 min to obtain equilibration. N2O concentration was measured by infrared spectrophotometer and GC. A strong correlation was discovered between the bag and urine concentrations, suggesting that 90 % of urinary N2O is caused by environmental exposure [69]. As a result, the urine matrix may be a useful indicator of N2O environmental exposure. Another cohort confirmed these results in 1985. The subjects were outfitted with pump-bag sampling to monitor the N2O content, which they wore all day long (98 days of work in total). N2O was analysed by IR or GC. They collected urine before starting anesthetic work and at the end (Table 1). They also investigated the effects of ventilation, close scavenging and types of anesthetics (mask, intubation, epidural) in several departments (pediatric, gynecological, general, orthopaedical and thoracic surgery). There was a positive correlation between N2O in bags and urine headspace N2O, this correlation being influenced by intra-individual factors and handling of urine specimens [70]. Duplicate samples were stored in a refrigerator or at room temperature and analysed up 24 h after (correlation coefficient between pump bag concentration and urine headspace gas measurement=0.94) [70]. The investigators concluded that urine specimens can be used in a practical way.
Similar studies on occupational exposure have been carried out, in several countries [68, 71]. These studies showed that a quantifiable concentration of N2O was found in human urine after exposure, but the kinetics were not studied. There is no data on the time lapse after which it can be recovered. However, a correlation was found by Trevisan et al. between the concentration in exhaled air just after a 6-h exposure and the urinary concentration 2–3 min after the end of the exposure [63]. As previously mentioned, the concentration drops in exhaled air 5 min after exposure [64]. It may be hypothesized that urinary concentration falls in the same way after exposure ceases. This is consistent with an occupational exposure study which showed that urinary N2O concentrations decrease highly after exposure ends. Urine and blood samples were collected the next morning from 25 exposed participants, approximately 18 h after the exposure ended. The mean N2O level found was 1.52 μg/L compared to 40 μg/L at the end of operating session (Table 1). Moreover, this concentration so long after exposure is no different from the concentration of 1 μg/L found in control subjects (blood donors) [66].
However, these data are not derived from patients who were directly exposed, so they should be interpreted with caution. Furthermore, urine matrix is frequently difficult to obtain, and exposed patients may not want to provide. Moreover, N2O doses differ from occupational to abuse. Hence, urine may be a complicated matrix for laboratory medicine and it may be more interesting to focus on blood or exhaled air.
Transdermal elimination
Cullen et al. used an in vitro test to demonstrate how N2O diffused through the skin [72]. Healthy skin samples were recovered from amputated limbs. Each specimen (n=7) was immersed in a water bath in a diffusion chamber (37 °C or/and 24°). A steady flow of humidified N2O (85.1 %) was applied at a constant rate of 1 L per minute. The gas spread through the skin to the other side of the nitrogen-filled glass box. Gas samples were analysed by GC. The N2O rate diffusion through skin at 37 °C was 10.38 mL/min/m2 with a mean difference of 2.47 mL/min/m2 between 37 and 24 °C.
In 1968, the percutaneous loss in surgical patients anesthetized with N2O was measured. Patients were subjected to a flow of 80 % N2O, with 20 % O2. A gas sample was collected at the start and then every 10 min and analysed by GC. The N2O percutaneous loss was measured at 3.6 mL/m2/min or 214 mL/m2/h in an average sized adult after 100 min of anesthesia (plateau between 60 and 100 min) [73]. This is much lower than the rates found in vitro [72]. This may be due to the fact that the amount of N2O reaching the skin in patients does not represent the total intake, and that it diffuses and is distributed elsewhere in the body. To describe the influence of temperature on this process, ice and heating patches were applied and the desired temperature was maintained for 30 min. There was a linear relationship between temperature (20–40 °C) and skin losses. The increase of skin temperature opening capillaries nearer the skin surface which decrease the path length for diffusion who increasing percutaneous loss. The conclusion was drawn that diffusion is the most important limiting factor for percutaneous loss [73].
Maternal milk
There is no information available about excretion in breast milk. According to Hale et al., N2O is unlikely to be expelled in breast milk or ingested by newborns (albeit no experiments have been performed) [74].
Involvement for laboratory medicine
In the context of driving under the influence of N2O and because N2O elimination is primarily pulmonary, it may be useful to detect it with a technique based on a breathalyzer test kit. Elimination is however too rapid, considering that 5 min after the last inhalation, it has nearly entirely disappeared. Regarding the use of urine samples, in addition to the fact that patients are not always compliant for this type of sampling, data is lacking from patients who are direct consumers. Urine measurement seems possible even a few days after inhalation, but in addition to the above limitations, predicting pathological cut-off is complicated and there is a lack of studies in subjects chronically consuming large quantities of N2O. Also, patients with urinary tract infections may have false positive results.
As direct detection of N2O is complicated to implement routinely due to the specific methods and difficult to interpret due to the lack of kinetics data, it is necessary to look at the effects of N2O on metabolism in order to find indirect markers.
Effects monitoring: impact of N2O on metabolic pathways
The most commonly metabolic target of N2O in the literature is vitamin B12 (or cobalamin) [75]. Vitamin B12’s structure includes a cobalt ion. This ion is the target of N2O, which changes its oxidation state [76]. N2O is a powerful oxidizing agent. In vitro studies evaluated reactions with various transition-metal complexes including cobalamins. The component of the gas mixture (N2O, N2 and H2) was measured by mass spectrometry after N2O was mixed with the transition-metal complex. N2O reacted quickly with transition-metal complexes acting as a two-electron oxidizing agent with the cobalt ion of cobalamins [77]. N2O was reduced into N2 by the following reaction 2Co(I) + N2O → 2Co(II) + N2 [78]. When N2O was added to a vitamin B12 solution, the amount of N2 produced by the reaction was too small to be quantified. In a second experiment, the cobalamins were reduced using borohydride to get solely the Cob(I) forms before being exposed to N2O. As a result, the amount of N2 responding was high. N2O targets cob(I)forms, whose oxidation disrupts the physiological metabolic cycle [77].
Cobalamin(I) normally interacts with 5-methyl-tetrahydrofolate (THF) to regenerate the methylcobalamin which is essential for methionine synthase (MS) function [79] (Figure 3). In this case, a metabolic adaptation prevents the cycle from functioning following oxidation of the reduced forms, because the cobalamin(II) form has to go through a reactivation cycle before it can be methylated back into the cobalamin(III) form [80]. Because this reactivation reaction combines methylation, utilizing S-adenosyl-methionine (SAM) as the methyl donor, with the reduction of cob(II) to cob(I), it is known as reductive methylation [81]. The chemical mechanics of this reaction are not fully understood [82, 83]. The reaction equation is as follows: 2 cob(II)alamin-[methionine synthase] + NADPH + 2 S-adenosyl-l-methionine=H+ + 2 methylcob(III)alamin-[methionine synthase] + NADP+ + 2 S-adenosyl-l-homocysteine (Figure 3). As a result, the concentration of various metabolites varies compared to physiological levels. There is a redistribution of the various folate derivatives detected. Following N2O exposure, the cytosolic quantity of THF decreases, benefiting the 5-methyl-THF form, which is twice as abundant as in an unexposed rat population [84]. A significant variation is also observed in the measurement of the SAM/SAH ratio, due to the significant rise in SAM levels caused by the increased activity of methionine synthase reductase (MTRR) [85] (Figure 3).

Consequences of oxidation of cobalamins. Cbl, cobalamin; Co, cobalt; THF, tetrahydrofolate; MTR, methionine synthase; MTRR, methionine synthase reductase; MAT1A, methionine adenosyltransferase 1A; AHCY, S-adenosyl-L-homocysteine hydrolase-MTR (=MS) uses methyl cobalamin(III) as a cofactor to supply a methyl group to homocysteine, thus forming methionine. This transmethylation forms a highly unstable, excited cobalamin(I) which, under physiological conditions, is mostly remethylated via folates. In the presence of N2O and/or oxidative stress, the cobalamin(I) form is highly oxidized to cobalamin(II), thus blocking MTR activity, as the latter can no longer receive a methyl group. MTRR comes into play here, enabling electron transfer and cobalamin remethylation, this time using the methyl group of SAM.
Cobalamins are not produced by human cells, they are obtained solely from food from animal origin after intestinal absorption. The CD320 receptor transports vitamin B12 to cells in transcobalamin II-bound form, allowing for endocytosis [86] (Figure 4). Cobalamin is an important enzymatic cofactor for methylmalonyl-coA mutase (MM-CoAM) and methionine synthase (MS) respectively adenosylcobalamin and methylcobalamin [81]. MS is involved in the one-carbon metabolism [87], and in multiple regulatory pathways, including cell signaling and epigenetics via methylation of DNA, RNA and various proteins.

Impact of nitrous oxide on the one carbon metabolism. Cbl, cobalamin; THF, tetrahydrofolate; SAM, S-adenosylmethionine; SAH, S-adenosylhomocysteine; PCC, propionyl-CoA carboxylase; MM-CoA mutase, methylmalonyl-CoA-mutase; MS, methionine synthase; MAT, methionine adenosyl-transferase; GNMT, glycine N-methyltransferase; AHCY, S-adenosyl-L-homocysteine hydrolase; CBS, cystathione β-synthase; BHMT, betaine-homocysteine S-methyl-transferase; CHDH, choline deshydrogenase; SHMT, serine hydroxymethyl-transférase; MTHFR, methylenetetrahydrofolate reductase.
N2O interferes with this metabolic cycle by inactivating the functions of vitamin B12 cofactors [88]. In this pathway, methionine is transformed into homocysteine to provide a methyl group for other compounds. MS converts homocysteine into methionine by transferring the 5-methyl-THF group via methylcobalamin (Figure 4). This folate-dependent re-methylation pathway is largely dependent on appropriate and sufficient intake of vitamin B9, but also of vitamins B6 and B2, cofactors of the enzymes involved in this cycle. A parallel, folate-independent re-methylation mechanism involving betaine also re-methylates homocysteine to produce methionine, via its conversion to dimethylglycine (Figure 4). Finally, homocysteine can be coupled with a serine molecule via cystathione β-synthase (CBS) to produce cystathione (transsulfuration pathway), which allows for the synthesis of glutathione, an anti-oxidant metabolite [89]. Adenosylcobalamin (another active form of vitamin B12) is engaged in mitochondrial metabolic pathways. It is the cofactor of MM-CoAM, that transforms methylmalonyl-CoA to succinyl-CoA (Figure 4). Metabolism of methylmalonyl-CoA to methylmalonic acid (MMA) is a minority metabolic pathway in case of MM-CoAM enzymatic blockage.
N2O effect on the activity of enzymes co-factored with cobalamins has been demonstrated in vitro and in animal models [90], [91], [92], [93], [94], [95], [96], [97]. In 1978, MS activity was evaluated using the method of Kamely et al. in rats under normal conditions and after exposure to 50 % N2O/50 % oxygen (30 min and 6 h groups). Simultaneously, urine was collected from the same rats to determine the concentration of urine MMA. MS activity was reduced by 30 min of exposure to N2O. No significant difference in the MMA was found between the groups [96]. However, it cannot be proven that urine MMA accurately reflects MM-CoAM activity. In 1980, methionine production was evaluated to understand the observed decline in MS activity. Methionine levels in the liver decreased after 30 min of exposure to 50 % N2O/50 % oxygen and were undetectable after 6 h. Half of the animals remained abnormal after 84 h, supporting the fact that recovery could be prolonged [98]. In rats exposed to N2O, MS activity in the brain become undetectable. A deoxyuridine suppression test [99] was also performed, which is a sensitive index for the presence of megaloblastic anemia due to vitamin B12 or folate deficiency. It was normal in animals exposed for 30 min but became abnormal after 60 min and gradually declined until 6 h [98]. MS activity is decreased not only in the liver, but also in other tissues [100]. Rats were housed in a chamber with 80/20 combination of N2O and oxygen and were euthanized after 18 h to obtain bone marrow, kidney, brain, and liver samples in which MS activity was determined. The percentages of inhibition of MS in the liver were the lowest (83.3 %), while the kidney was the most impacted (93.3 %) [100]. The loss of MS activity is closely linked to the decrease in concentration of its cofactor, methylcobalamin [91]. This correlation is also found for the recovery of enzymatic activity. The blocking of the MS is irreversible and has been demonstrated in cell cultures subjected to N2O. When exposure was stopped, some cultures were supplemented with cycloheximide (a protein synthesis blocking agent). Under these conditions, no recovery of activity was observed [91]. The blocking action of N2O on MS activity is therefore irreversible.
Human patients anesthetized with N2O have significantly lower hepatic MS activity [101]. When exposed to 66 % N2O for 3 h, enzyme blockage reduces MS activity by 25 %. Plasma methionine (synthesis product of MS) correlates with patients’ clinical severity, assessed by the PND (peripheral neuropathy Disability) score [102]. However, plasma concentrations remain within the physiological range, but the greater the clinical severity, the lower the level. Neurological damage is only partly explained by enzyme damage, and other hypotheses such as oxidative stress need to be explored [103]. On the other hand, analysis of biological data from a cohort of N2O consumers showed an increase in plasma (plasma homocysteine increase was defined as >15 μmol/L) and MMA (plasma MMA increase was defined as >0.4 μmol/L) in these patients [104]. The increase in MMA is less significant because it has not been systematically found, but MMA levels correlate with the patient’s clinical severity, which explain why this biomarker represents the extent of metabolic alteration. Because N2O interferes with the action of cobalamin cofactors, studies have investigated total vitamin B12 dosage for the diagnosis of N2O intoxication and it was not found to be a good marker. The alteration in metabolism is qualitative rather than quantitative. N2O is responsible for a functional alteration of the cycle. Hence, a decrease in vitamin B12 levels is not systematically observed [104, 105] and basically represents a quantitative associated deficit. Furthermore, many N2O users may supplement with vitamin B12, either by self-medicating due to unsubstantiated advice available on the internet and social networks, or after consulting with a medical practitioner. Thus, vitamin B12 consumption is unrelated to N2O consumption.
Involvement for laboratory medicine
Consuming N2O has metabolic side effects that are starting to be well understood, at least for the cytoplasmic pathway. It is still not well understood how N2O enters the mitochondria, or at least how it affects this metabolic pathway. Vitamin B12, homocysteine and MMA assays are already available in medical laboratories. The techniques used are those traditionally employed, namely immunochemistry, liquid chromatography and mass spectrometry [106], [107], [108], [109], [110]. The combined use of these markers would appear to be relevant in assessing the metabolic damage attributed to N2O consumption [104]. Mass spectrometry panels for homocysteine, methionine and MMA may be used to provide a complete interpretation of the patient’s profile.
Summary and critical view
N2O has generally been investigated in subjects with occupational or environmental exposure, but direct consumption has been less studied. Monitoring N2O in the context of professional use is critical, but even more so in the context of recreational use, which is currently on the rise. Exposure monitoring is complicated to set up because it is therefore difficult to establish detection times and values to diagnose N2O intake and its timing. Despite the fact that the pulmonary route is the predominant means of elimination, measurement in exhaled air is unusable, because of its rapid rate of elimination. On the other hand, blood and urinary concentrations are too unpredictable, due to a lack of data on direct consumers. In view of all the data described, determination of N2O in biological fluids as an exposure biomarker does not appear to be simple and practical for routine use at this time.
On the other hand, effects biomarkers are biological parameters that can predict a clinically relevant endpoint or intermediate outcome difficult to observe clinically [111]. Some biomarkers could have an interest in the case of N2O consumption. Importantly, a biomarker used to diagnose N2O intoxication is not necessarily the same biomarker to determine clinical severity or to monitor patient. To identify one or more biomarkers, it is essential to understand the pathophysiology of intoxication [112]. For this reason, a detailed understanding of related metabolism and impact of N2O on enzymes and metabolites in the cytoplasmic and mitochondrial pathways is essential.
Currently, there are no biomarkers specific to N2O intoxication. Use of biomarkers reflecting metabolic alterations, such as homocysteine and MMA, seems more efficient and easier to interpret and are also available by routine analytical techniques in laboratory medicine (LC-MS, immunochemistry, enzymatic method).
Homocysteine appears to be the most sensitive biological marker of N2O consumption with high plasma levels observed in most consumers. However, there is no correlation between the quantity consumed and homocysteine levels. Homocysteine rises very rapidly after the first consumption, but also decreases within a few days. A normal or slightly elevated level can therefore be attributed to a significant delay between sampling and last consumption; it is therefore a marker of recent consumption. However, this marker lacks specificity for recent N2O consumption. Indeed, its increase may also reflect common pathologies (renal failure, hypothyroidism, vitamin B12, B6 or B9 deficiency) as well as hereditary metabolic diseases such as methylenetetrahydrolate reductase (MTHFR), methionine synthase (MS) or cystathionine beta synthase (CBS) deficiency. Its utility in the clinical setting is also limited by its needing to be transported on ice and separated within a short time following collection.
MMA assays are useful, as they are more specific. On the other hand, an increase in MMA is not systematically observed. This biomarker is implicated in a separate metabolic pathway, which is mitochondrial. The effect of N2O on methylcobalamin has been widely demonstrated in pre-clinical models, but not on adenosylcobalamin. However, in several cases reported in the literature, MMA increases were observed in the most severe cases of intoxication. A correlation has been demonstrated between elevated MMA levels and clinical severity [30], reflecting deeper metabolic alteration.
Today, none of these described markers are strictly related to N2O consumption or the duration of its consumption [30], and further studies are needed to define the precise kinetics of those markers. Optimal patient management therefore relies on monitoring a combination of markers. Plasma homocysteine and MMA combination reflects metabolic saturation by N2O. A concomitant assay of vitamins B12, B9 or even B6 is also necessary to detect any associated deficiencies and supplement the patient in case of associated deficiency. This is essential for clinical management, after definitive cessation of consumption, but also aids with the interpretation of the above-mentioned markers. Liver and kidney function tests, as well as a full blood count, should also be carried out to rule out other secondary causes of increases in these metabolic markers of interest.
Conclusions
Routine and direct measurement of N2O in biological fluids is complicated and difficult to interpret. Literature show that measurement techniques are reserved for highly specialized laboratories, and nitrous oxide pharmacokinetics data show that this molecule is difficult to quantify in the usual matrices used in medical laboratories. Indirect metabolic markers such as homocysteine and MMA are therefore of major interest, but there is a lack of data on the kinetics of these markers in the context of nitrous oxide use, and these markers still lack specificity, so there is a need to have a better understanding of consequences and related pathophysiology identify new markers such as oxidative stress markers or other metabolites.
Acknowledgments
We would like to thank Laura Plasse and Isabelle Kim for their support in compiling the bibliography.
-
Research ethics: Not applicable.
-
Informed consent: Not applicable.
-
Author contributions: The authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Competing interests: The authors state no conflict of interest.
-
Research funding: None declared.
-
Data availability: Not applicable.
References
1. Becker, DE, Rosenberg, M. Nitrous oxide and the inhalation anesthetics. Anesth Prog 2008;55:124–31. https://doi.org/10.2344/0003-3006-55.4.124.Search in Google Scholar PubMed PubMed Central
2. Winstock, AR. Global drug survey (GDS). Key findings report. Lisbon; 2021.Search in Google Scholar
3. European Monitoring Centre for Drugs and Drug Addiction. Recreational use of nitrous oxide: a growing concern for Europe. [Internet]. Luxembourg : Publications Office; 2022 [cited 2023 Jul 19]. Available from: https://data.europa.eu/doi/10.2810/2003.Search in Google Scholar
4. van Amsterdam, JG, Nabben, T, van den Brink, W. Increasing recreational nitrous oxide use: should we worry? A narrative review. J Psychopharmacol 2022;36:943–50. https://doi.org/10.1177/02698811221082442.Search in Google Scholar PubMed
5. Fidalgo, M, Prud’homme, T, Allio, A, Bronnec, M, Bulteau, S, Jolliet, P, et al.. Nitrous oxide: what do we know about its use disorder potential? Results of the French Monitoring Centre for Addiction network survey and literature review. Subst Abuse 2019;40:33–42. https://doi.org/10.1080/08897077.2019.1573210.Search in Google Scholar PubMed
6. Ng, J, O’Grady, G, Pettit, T, Frith, R. Nitrous oxide use in first-year students at Auckland University. Lancet 2003;361:1349–50. https://doi.org/10.1016/s0140-6736(03)13045-0.Search in Google Scholar PubMed
7. Mair, D, Paris, A, Zaloum, SA, White, LM, Dodd, KC, Englezou, C, et al.. Nitrous oxide-induced myeloneuropathy: a case series. J Neurol Neurosurg Psychiatry 2023;94:681–8. https://doi.org/10.1136/jnnp-2023-331131.Search in Google Scholar PubMed PubMed Central
8. Keddie, S, Adams, A, Kelso, ARC, Turner, B, Schmierer, K, Gnanapavan, S, et al.. No laughing matter: subacute degeneration of the spinal cord due to nitrous oxide inhalation. J Neurol 2018;265:1089–95. https://doi.org/10.1007/s00415-018-8801-3.Search in Google Scholar PubMed PubMed Central
9. Garakani, A, Jaffe, RJ, Savla, D, Welch, AK, Protin, CA, Bryson, EO, et al.. Neurologic, psychiatric, and other medical manifestations of nitrous oxide abuse: a systematic review of the case literature: toxic Effects of Nitrous Oxide Abuse. Am J Addict 2016;25:358–69. https://doi.org/10.1111/ajad.12372.Search in Google Scholar PubMed
10. Meier, E, Malviya, M, Kaur, S, Ibrahim, J, Corrigan, A, Moawad, A, et al.. Neurologic and thrombotic complications in the setting of chronic nitrous oxide abuse. Case Rep Med 2023;2023:5058771. https://doi.org/10.1155/2023/5058771.Search in Google Scholar PubMed PubMed Central
11. Caris, MG, Kuipers, RS, Kiestra, BE, Ruijter, BJ, Riezebos, RK, Coppens, M, et al.. Nitrous oxide abuse leading to extreme homocysteine levels and thrombosis in young adults: a case series. J Thromb Haemost JTH 2023;21:276–83. https://doi.org/10.1016/j.jtha.2022.10.002.Search in Google Scholar PubMed
12. Oulkadi, S, Peters, B, Vliegen, AS. Thromboembolic complications of recreational nitrous oxide (ab)use: a systematic review. J Thromb Thrombolysis 2022;54:686–95. https://doi.org/10.1007/s11239-022-02673-x.Search in Google Scholar PubMed
13. McCaddon, A, Miller, JW. Homocysteine—a retrospective and prospective appraisal. Front Nutr 2023;10:1179807. https://doi.org/10.3389/fnut.2023.1179807.Search in Google Scholar PubMed PubMed Central
14. Paris, A, Lake, L, Joseph, A, Workman, A, Walton, J, Hayton, T, et al.. Nitrous oxide-induced subacute combined degeneration of the cord: diagnosis and treatment. Pract Neurol 2023;23:222–8. https://doi.org/10.1136/pn-2022-003631.Search in Google Scholar PubMed PubMed Central
15. Gernez, E, Lee, GR, Niguet, JP, Zerimech, F, Bennis, A, Grzych, G. Nitrous oxide abuse: clinical outcomes, pharmacology, pharmacokinetics, toxicity and impact on metabolism. Toxics 2023;11:962. https://doi.org/10.3390/toxics11120962.Search in Google Scholar PubMed PubMed Central
16. Sanders, RD, Weimann, J, Maze, M, Warner, D, Warner, M. Biologic effects of nitrous oxide: a mechanistic and toxicologic review. Anesthesiology 2008;109:707–22. https://doi.org/10.1097/aln.0b013e3181870a17.Search in Google Scholar
17. Ohashi, Y, Guo, T, Orii, R, Maze, M, Fujinaga, M. Brain stem opioidergic and GABAergic neurons mediate the antinociceptive effect of nitrous oxide in fischer rats. Anesthesiology 2003;99:947–54. https://doi.org/10.1097/00000542-200310000-00030.Search in Google Scholar PubMed
18. Nagele, P, Duma, A, Kopec, M, Gebara, MA, Parsoei, A, Walker, M, et al.. Nitrous oxide for treatment-resistant major depression: a proof-of-concept trial. Biol Psychiatry 2015;78:10–8. https://doi.org/10.1016/j.biopsych.2014.11.016.Search in Google Scholar PubMed
19. Kalmoe, MC, Janski, AM, Zorumski, CF, Nagele, P, Palanca, BJ, Conway, CR. Ketamine and nitrous oxide: the evolution of NMDA receptor antagonists as antidepressant agents. J Neurol Sci 2020;412:116778. https://doi.org/10.1016/j.jns.2020.116778.Search in Google Scholar PubMed
20. Turner, DA, Chen, D, Galbally, IE, Leuning, R, Edis, RB, Li, Y, et al.. Spatial variability of nitrous oxide emissions from an Australian irrigated dairy pasture. Plant Soil 2008;309:77–88. https://doi.org/10.1007/s11104-008-9639-8.Search in Google Scholar
21. Pattey, E, Edwards, GC, Desjardins, RL, Pennock, DJ, Smith, W, Grant, B, et al.. Tools for quantifying N2O emissions from agroecosystems. Agric For Meteorol 2007;142:103–19. https://doi.org/10.1016/j.agrformet.2006.05.013.Search in Google Scholar
22. Brut, A, Legain, D, Durand, P, Laville, P. A relaxed eddy accumulator for surface flux measurements on ground-based platforms and aboard research vessels. J Atmospheric Ocean Technol 2004;21:411. https://doi.org/10.1175/1520-0426(2004)021<0411:areafs>2.0.co;2.10.1175/1520-0426(2004)021<0411:AREAFS>2.0.CO;2Search in Google Scholar
23. Denmead, OT, Harper, LA, Freney, JR, Griffith, DWT, Leuning, R, Sharpe, RR. A mass balance method for non-intrusive measurements of surface-air trace gas exchange. Atmos Environ 1998;32:3679–88. https://doi.org/10.1016/s1352-2310(98)00091-0.Search in Google Scholar
24. Griffiths, SRPR, de Haseth, JA. Fourier transform infrared spectrometry (2nd edn.). Anal Bioanal Chem 2008;391:2379–80. https://doi.org/10.1007/s00216-008-2144-3.Search in Google Scholar
25. Kelliher, FM, Reisinger, AR, Martin, RJ, Harvey, MJ, Price, SJ, Sherlock, RR. Measuring nitrous oxide emission rate from grazed pasture using Fourier-transform infrared spectroscopy in the nocturnal boundary layer. Agric For Meteorol 2002;111:29–38. https://doi.org/10.1016/s0168-1923(02)00007-2.Search in Google Scholar
26. Röckmann, T, Kaiser, J, Brenninkmeijer, CAM, Crowley, JN, Borchers, R, Brand, WA, et al.. Isotopic enrichment of nitrous oxide (15N14NO, 14N15NO, 14N14N18O) in the stratosphere and in the laboratory. J Geophys Res Atmospheres 2001;106:10403–10. https://doi.org/10.1029/2000jd900822.Search in Google Scholar
27. Ekeberg, D, Ogner, G, Fongen, M, Joner, EJ, Wickstrøm, T. Determination of CH4, CO2 and N2O in air samples and soil atmosphere by gas chromatography mass spectrometry, GC-MS. J Environ Monit 2004;6:621–3. https://doi.org/10.1039/b401315h.Search in Google Scholar PubMed
28. Klepper, LA. Nitric oxide emissions from soybean leaves during in vivo nitrate reductase assays 1. Plant Physiol 1987;85:96–9. https://doi.org/10.1104/pp.85.1.96.Search in Google Scholar PubMed PubMed Central
29. Drescher, SR, Brown, SD. Solid phase microextraction-gas chromatographic–mass spectrometric determination of nitrous oxide evolution to measure denitrification in estuarine soils and sediments. J Chromatogr A 2006;1133:300–4. https://doi.org/10.1016/j.chroma.2006.08.056.Search in Google Scholar PubMed
30. Isobe, K, Koba, K, Ueda, S, Senoo, K, Harayama, S, Suwa, Y. A simple and rapid GC/MS method for the simultaneous determination of gaseous metabolites. J Microbiol Methods 2011;84:46–51. https://doi.org/10.1016/j.mimet.2010.10.009.Search in Google Scholar PubMed
31. Ishimura, Y, Gao, YT, Panda, SP, Roman, LJ, Masters, BSS, Weintraub, ST. Detection of nitrous oxide in the neuronal nitric oxide synthase reaction by gas chromatography–mass spectrometry. Biochem Biophys Res Commun 2005;338:543–9. https://doi.org/10.1016/j.bbrc.2005.07.202.Search in Google Scholar PubMed
32. Saloojee, Y, Cole, P. Estimation of nitrous oxide in blood. Gas chromatographic analysis of trace or analgesic levels. Anaesthesia 1978;33:779–83. https://doi.org/10.1111/j.1365-2044.1978.tb08493.x.Search in Google Scholar PubMed
33. Causon, R. Validation of chromatographic methods in biomedical analysis viewpoint and discussion. J Chromatogr B Biomed Sci Appl 1997;689:175–80. https://doi.org/10.1016/s0378-4347(96)00297-6.Search in Google Scholar PubMed
34. Winek, CL, Wahba, WW, Rozin, L. Accidental death by nitrous oxide inhalation. Forensic Sci Int 1995;73:139–41. https://doi.org/10.1016/0379-0738(95)01729-3.Search in Google Scholar PubMed
35. Poli, D, Gagliano-Candela, R, Strisciullo, G, Colucci, AP, Strada, L, Laviola, D, et al.. Nitrous oxide determination in postmortem biological samples: a case of serial fatal poisoning in a public hospital. J Forensic Sci 2010;55:258–64. https://doi.org/10.1111/j.1556-4029.2009.01218.x.Search in Google Scholar PubMed
36. Giuliani, N, Beyer, J, Augsburger, M, Varlet, V. Validation of an analytical method for nitrous oxide (N2O) laughing gas by headspace gas chromatography coupled to mass spectrometry (HS-GC–MS): forensic application to a lethal intoxication. J Chromatogr B 2015;983–984:90–3. https://doi.org/10.1016/j.jchromb.2014.12.034.Search in Google Scholar PubMed
37. Kemble, DJ, Cervinski, MA. A single-column gas chromatography method for quantifying toxic alcohols. J Appl Lab Med 2020;5:300–10. https://doi.org/10.1093/jalm/jfz019.Search in Google Scholar PubMed
38. Heusler, H. Quantitative analysis of common anaesthetic agents. J Chromatogr 1985;340:273–319. https://doi.org/10.1016/0378-4347(85)80200-0.Search in Google Scholar PubMed
39. Dahlgren, BE, Olander, L, Ovrum, P. Pollution of delivery ward air by nitrous oxide--methoxyflurane. Am Ind Hyg Assoc J 1979;40:666–72. https://doi.org/10.1080/15298667991430136.Search in Google Scholar PubMed
40. Henderson, KA, Matthews, IP. Environmental monitoring of nitrous oxide during dental anaesthesia. Br Dent J 2000;188:617–9. https://doi.org/10.1038/sj.bdj.4800556.Search in Google Scholar PubMed
41. Byhahn, C, Heller, K, Lischke, V, Westphal, K. Surgeon’s occupational exposure to nitrous oxide and sevoflurane during pediatric surgery. World J Surg 2001;25:1109–12. https://doi.org/10.1007/bf03215855.Search in Google Scholar PubMed
42. Rapson, TD, Dacres, H. Analytical techniques for measuring nitrous oxide. TrAC, Trends Anal Chem 2014;54:65–74. https://doi.org/10.1016/j.trac.2013.11.004.Search in Google Scholar
43. Hensen, A, Skiba, U, Famulari, D. Low cost and state of the art methods to measure nitrous oxide emissions. Environ Res Lett 2013;8:025022. https://doi.org/10.1088/1748-9326/8/2/025022.Search in Google Scholar
44. Esler, MB, Griffith, DWT, Turatti, F, Wilson, SR, Rahn, T, Zhang, H. N2O concentration and flux measurements and complete isotopic analysis by FTIR spectroscopy. Chemosphere – Glob Change Sci 2000;2:445–54. https://doi.org/10.1016/s1465-9972(00)00033-7.Search in Google Scholar
45. Güllük, T, Wagner, HE, Slemr, F. A high-frequency modulated tunable diode laser absorption spectrometer for measurements of CO2, CH4, N2O, and CO in air samples of a few cm3. Rev Sci Instrum 1997;68:230–9. https://doi.org/10.1063/1.1147814.Search in Google Scholar
46. Otteni, J, Fournier, S, Collin, F. Protoxyde d’azote. In: Conférences d’actualisation 1997 39e congrès national d’anesthésie et de réanimation. Paris: Masson 1997.Search in Google Scholar
47. Admin, B. Enquête sur le protoxyde d’Azote et l’anesthésie générale - La SFAR [Internet]. Société Française d’Anesthésie et de Réanimation. 2018 [cited 2023 Aug 1]. Available from: https://sfar.org/enquete-sur-le-protoxyde-dazote-et-lanesthesie-generale/.Search in Google Scholar
48. American Academy of Pediatric Dentistry. Guideline on use of nitrous oxide for pediatric dental patients. Pediatr Dent 2013;35:E174–8.Search in Google Scholar
49. Hallonsten, AL, Koch, G, Schröder, U. Nitrous oxide-oxygen sedation in dental care. Community Dent Oral Epidemiol 1983;11:347–55. https://doi.org/10.1111/j.1600-0528.1983.tb01390.x.Search in Google Scholar PubMed
50. Ashley, P, Anand, P, Andersson, K. Best clinical practice guidance for conscious sedation of children undergoing dental treatment: an EAPD policy document. Eur Arch Paediatr Dent Off J Eur Acad Paediatr Dent 2021;22:989–1002. https://doi.org/10.1007/s40368-021-00660-z.Search in Google Scholar PubMed PubMed Central
51. Papper, EM, Kitz, R. Uptake and distribution of anesthetic agents; 1963. [cited 2023 Aug 1]. Available from: https://www.semanticscholar.org/paper/Uptake-and-distribution-of-anesthetic-agents-Papper-Kitz/8fe65d5e22e4f4c3d692a73ff4c3a5afd4c6e5b3.Search in Google Scholar
52. Stenqvist, O. Nitrous oxide kinetics. Acta Anaesthesiol Scand 1994;38:757–60. https://doi.org/10.1111/j.1399-6576.1994.tb03997.x.Search in Google Scholar PubMed
53. Vichitvejpaisal, P, Joshi, GP, Liu, J, White, PF. Effect of severity of pulmonary disease on nitrous oxide washin and washout characteristics. J Med Assoc Thail Chotmaihet Thangphaet 1997;80:378–83.Search in Google Scholar
54. Evans, DE, Flook, V, Mapleson, WW. A comparison of in-vivo and in-vitro partition coefficients for nitrous oxide and cyclopropane. Br J Anaesth 1970;42:1028–32. https://doi.org/10.1093/bja/42.12.1028.Search in Google Scholar PubMed
55. Kety, SS, Harmel, MH, Broomell, HT, Clara Belle, R. The solubility of nitrous oxide in blood and brain. J Biol Chem 1948;173:487–96. https://doi.org/10.1016/s0021-9258(18)57420-2.Search in Google Scholar
56. Krapez, JR, Saloojee, Y, Hinds, CJ, Hackett, GH, Cole, PV. Blood concentrations of nitrous oxide in theatre personnel. Br J Anaesth 1980;52:1143–8. https://doi.org/10.1093/bja/52.11.1143.Search in Google Scholar PubMed
57. Sher, AM, Braude, BM, Cleaton-Jones, PE, Moyes, DG, Mallett, J. Nitrous oxide sedation in dentistry: a comparison between Rotameter settings, pharyngeal concentrations and blood levels of nitrous oxide. Anaesthesia 1984;39:236–9. https://doi.org/10.1111/j.1365-2044.1984.tb07233.x.Search in Google Scholar PubMed
58. Allen, MB, Van Niel, CB. Experiments on bacterial denitrification. J Bacteriol 1952;64:397–412. https://doi.org/10.1128/jb.64.3.397-412.1952.Search in Google Scholar PubMed PubMed Central
59. Coyle, CL, Zumft, WG, Kroneck, PM, Körner, H, Jakob, W. Nitrous oxide reductase from denitrifying Pseudomonas perfectomarina. Purification and properties of a novel multicopper enzyme. Eur J Biochem 1985;153:459–67. https://doi.org/10.1111/j.1432-1033.1985.tb09324.x.Search in Google Scholar PubMed
60. Hong, K, Trudell, JR, O’Neil, JR, Cohen, EN. Metabolism of nitrous oxide by human and rat intestinal contents. Anesthesiology 1980;52:16–9. https://doi.org/10.1097/00000542-198001000-00004.Search in Google Scholar PubMed
61. Salanitre, E, Rackow, H, Greene, LT, Klonymus, D, Epstein, RM. Uptake and excretion of subanesthetic concentrations of nitrous oxide in man. Anesthesiology 1962;23:814–22. https://doi.org/10.1097/00000542-196211000-00011.Search in Google Scholar PubMed
62. Fink, BR. Diffusion anoxia. Anesthesiology 1955;16:511–9. https://doi.org/10.1097/00000542-195507000-00007.Search in Google Scholar PubMed
63. Trevisan, A, Gori, GP. Biological monitoring of nitrous oxide exposure in surgical areas. Am J Ind Med 1990;17:357–62. https://doi.org/10.1002/ajim.4700170308.Search in Google Scholar PubMed
64. Einarsson, S, Stenqvist, O, Bengtsson, A, Houltz, E, Bengtson, JP. Nitrous oxide elimination and diffusion hypoxia during normo- and hypoventilation. Br J Anaesth 1993;71:189–93. https://doi.org/10.1093/bja/71.2.189.Search in Google Scholar PubMed
65. O’Reilly, JE, Matheny, JL, Falace, DA. The effects of nitrous oxide administration in the healthy elderly: N2O elimination and alveolar CO2. Anesth Prog 1983;30:187–92.Search in Google Scholar
66. Brugnone, F, Perbellini, L, Cerpelloni, M, Soave, C, Cecco, A, Giuliari, C. Nitrous oxide in blood and urine of operating theatre personnel and the general population. Int Arch Occup Environ Health 1996;68:22–6. https://doi.org/10.1007/bf01831629.Search in Google Scholar PubMed
67. Apostoli, P, Gelmi, M, Alessio, L, Turano, A. Interferences of urinary tract infection in the measurement of urinary nitrous oxide. Occup Environ Med 1996;53:591–4. https://doi.org/10.1136/oem.53.9.591.Search in Google Scholar PubMed PubMed Central
68. Zaffina, S, Lembo, M, Gilardi, F, Bussu, A, Pattavina, F, Tucci, MG, et al.. Nitrous oxide occupational exposure in conscious sedation procedures in dental ambulatories: a pilot retrospective observational study in an Italian pediatric hospital. BMC Anesthesiol 2019;19:42. https://doi.org/10.1186/s12871-019-0714-x.Search in Google Scholar PubMed PubMed Central
69. Sonander, H, Stenqvist, O, Nilsson, K. Urinary N2O as a measure of biologic exposure to nitrous oxide anaesthetic contamination. Ann Occup Hyg 1983;27:73–9.Search in Google Scholar
70. Sonander, H, Stenqvist, O, Nilsson, K. Nitrous oxide exposure during routine anaesthetic work. Acta Anaesthesiol Scand 1985;29:203–8. https://doi.org/10.1111/j.1399-6576.1985.tb02186.x.Search in Google Scholar PubMed
71. Accorsi, A, Barbieri, A, Raffi, G, Violante, F. Biomonitoring of exposure to nitrous oxide, sevoflurane, isoflurane and halothane by automated GC/MS headspace urinalysis. Int Arch Occup Environ Health 2001;74:541–8. https://doi.org/10.1007/s004200100263.Search in Google Scholar PubMed
72. Cullen, BF, Eger, EI. Diffusion of nitrous oxide, cyclopropane, and halothane through human skin and amniotic membrane. Anesthesiology 1972;36:168–73. https://doi.org/10.1097/00000542-197202000-00019.Search in Google Scholar PubMed
73. Stoelting, RK, Eger, EI. Percutaneous loss of nitrous oxide, cyclopropane, ether and halothane in man. Anesthesiology 1969;30:278–82. https://doi.org/10.1097/00000542-196903000-00008.Search in Google Scholar PubMed
74. Hale, TW. Anesthetic medications in breastfeeding mothers. J Hum Lact Off J Int Lact Consult Assoc 1999;15:185–94. https://doi.org/10.1177/089033449901500302.Search in Google Scholar PubMed
75. Chanarin, I. Cobalamins and nitrous oxide: a review. J Clin Pathol 1980;33:909–16. https://doi.org/10.1136/jcp.33.10.909.Search in Google Scholar PubMed PubMed Central
76. Van De Lagemaat, E, De Groot, L, Van Den Heuvel, E. Vitamin B12 in relation to oxidative stress: a systematic review. Nutrients 2019;11:482. https://doi.org/10.3390/nu11020482.Search in Google Scholar PubMed PubMed Central
77. Banks, RGS, Henderson, RJ, Pratt, JM. Reactions of gases in solution. Part III. Some reactions of nitrous oxide with transition-metal complexes. J Chem Soc Inorg Phys Theor 1968:2886–9. https://doi.org/10.1039/j19680002886.Search in Google Scholar
78. Blackburn, R, Kyaw, M, Swallow, AJ. Reaction of Cob(I)alamin with nitrous oxide and Cob(III)alamin. J Chem Soc Faraday Trans 1 Phys Chem Condens Phases 1977;73:250–5. https://doi.org/10.1039/f19777300250.Search in Google Scholar
79. McCaddon, A, Regland, B, Hudson, P, Davies, G. Functional vitamin B12 deficiency and Alzheimer disease. Neurology 2002;58:1395–9. https://doi.org/10.1212/wnl.58.9.1395.Search in Google Scholar PubMed
80. Wolthers, KR, Lou, X, Toogood, HS, Leys, D, Scrutton, NS. Mechanism of coenzyme binding to human methionine synthase reductase revealed through the crystal structure of the FNR-like module and isothermal titration calorimetry. Biochemistry 2007;46:11833–44. https://doi.org/10.1021/bi701209p.Search in Google Scholar PubMed
81. Matthews, RG. Cobalamin- and corrinoid-dependent enzymes. Met Ions Life Sci 2009;6:53–114.10.1039/9781847559333-00053Search in Google Scholar
82. Banerjee, R. The Yin-Yang of cobalamin biochemistry. Chem Biol 1997;4:175–86. https://doi.org/10.1016/s1074-5521(97)90286-6.Search in Google Scholar PubMed
83. Banerjee, RV, Harder, SR, Ragsdale, SW, Matthews, RG. Mechanism of reductive activation of cobalamin-dependent methionine synthase: an electron paramagnetic resonance spectroelectrochemical study. Biochemistry 1990;29:1129–35. https://doi.org/10.1021/bi00457a005.Search in Google Scholar PubMed
84. Horne, DW, Holloway, RS. Compartmentation of folate metabolism in rat pancreas: nitrous oxide inactivation of methionine synthase leads to accumulation of 5-methyltetrahydrofolate in cytosol. J Nutr 1997;127:1772–5. https://doi.org/10.1093/jn/127.9.1772.Search in Google Scholar PubMed
85. Molloy, AM, Orsi, B, Kennedy, DG, Kennedy, S, Weir, DG, Scott, JM. The relationship between the activity of methionine synthase and the ratio of S-adenosylmethionine to S-adenosylhomocysteine in the brain and other tissues of the pig. Biochem Pharmacol 1992;44:1349–55. https://doi.org/10.1016/0006-2952(92)90536-r.Search in Google Scholar PubMed
86. Froese, DS, Fowler, B, Baumgartner, MR. Vitamin B12, folate, and the methionine remethylation cycle-biochemistry, pathways, and regulation. J Inherit Metab Dis 2019;42:673–85. https://doi.org/10.1002/jimd.12009.Search in Google Scholar PubMed
87. Ducker, GS, Rabinowitz, JD. One-carbon metabolism in health and disease. Cell Metab 2017;25:27–42. https://doi.org/10.1016/j.cmet.2016.08.009.Search in Google Scholar PubMed PubMed Central
88. Sobczyńska-Malefora, A, Delvin, E, McCaddon, A, Ahmadi, KR, Harrington, DJ. Vitamin B12 status in health and disease: a critical review. Diagnosis of deficiency and insufficiency – clinical and laboratory pitfalls. Crit Rev Clin Lab Sci 2021;58:399–429. https://doi.org/10.1080/10408363.2021.1885339.Search in Google Scholar PubMed
89. Tavernarakis, N. One-carbon metabolism modulates ageing and neurodegeneration. Scholarly Community Encyclopedia, Neurodegeneration.Search in Google Scholar
90. Frasca, V, Riazzi, BS, Matthews, RG. In vitro inactivation of methionine synthase by nitrous oxide. J Biol Chem 1986;261:15823–6. https://doi.org/10.1016/s0021-9258(18)66636-0.Search in Google Scholar
91. Riedel, B, Fiskerstrand, T, Refsum, H, Ueland, PM. Co-ordinate variations in methylmalonyl-CoA mutase and methionine synthase, and the cobalamin cofactors in human glioma cells during nitrous oxide exposure and the subsequent recovery phase. Biochem J. 1999;341:133–8, https://doi.org/10.1042/bj3410133.Search in Google Scholar
92. Christensen, B, Rosenblatt, DS, Chu, RC, Ueland, PM. Effect of methionine and nitrous oxide on homocysteine export and remethylation in fibroblasts from cystathionine synthase-deficient, cb1G, and cb1E patients. Pediatr Res 1994;35:3–9. https://doi.org/10.1203/00006450-199401000-00002.Search in Google Scholar PubMed
93. Koblin, DD, Watson, JE, Deady, JE, Stokstad, EL, Eger, EI. Inactivation of methionine synthetase by nitrous oxide in mice. Anesthesiology 1981;54:318–24. https://doi.org/10.1097/00000542-198104000-00012.Search in Google Scholar PubMed
94. Koblin, DD, Tomerson, BW. Methionine synthase activities in mice following acute exposures to ethanol and nitrous oxide. Biochem Pharmacol 1989;38:1353–8. https://doi.org/10.1016/0006-2952(89)90343-2.Search in Google Scholar PubMed
95. Kondo, H, Osborne, ML, Kolhouse, JF, Binder, MJ, Podell, ER, Utley, CS, et al.. Nitrous oxide has multiple deleterious effects on cobalamin metabolism and causes decreases in activities of both mammalian cobalamin-dependent enzymes in rats. J Clin Invest 1981;67:1270–83. https://doi.org/10.1172/jci110155.Search in Google Scholar PubMed PubMed Central
96. Deacon, R, Perry, J, Lumb, M, Chanarin, I, Minty, B, Halsey, MJ, et al.. Selective inactivation of vitamin B12 in rats by nitrous oxide. Lancet 1978;312:1023–4. https://doi.org/10.1016/s0140-6736(78)92341-3.Search in Google Scholar PubMed
97. Koblin, DD, Everman, BW. Vitamin B12 and folate status in rats after chronic administration of ethanol and acute exposure to nitrous oxide. Alcohol Clin Exp Res 1991;15:543–8. https://doi.org/10.1111/j.1530-0277.1991.tb00557.x.Search in Google Scholar PubMed
98. Deacon, R, Lumb, M, Perry, J, Chanarin, I, Minty, B, Halsey, M, et al.. Inactivation of methionine synthase by nitrous oxide. Eur J Biochem 1980;104:419–23. https://doi.org/10.1111/j.1432-1033.1980.tb04443.x.Search in Google Scholar PubMed
99. Wickramasinghe, SN, Matthews, JH. Deoxyuridine suppression: biochemical basis and diagnostic applications. Blood Rev 1988;2:168–77. https://doi.org/10.1016/0268-960x(88)90022-7.Search in Google Scholar PubMed
100. Wilson, SD, Horne, DW. Effect of nitrous oxide inactivation of vitamin B12 on the levels of folate coenzymes in rat bone marrow, kidney, brain, and liver. Arch Biochem Biophys 1986;244:248–53. https://doi.org/10.1016/0003-9861(86)90114-1.Search in Google Scholar PubMed
101. Koblin, DD, Waskell, L, Watson, JE, Stokstad, ELR, Eger, EI. Nitrous oxide inactivates methionine synthetase in human liver. Anesth Analg 1982;61:75–8. https://doi.org/10.1213/00000539-198202000-00001.Search in Google Scholar
102. Gernez, E, Deheul, S, Tard, C, Joncquel, M, Douillard, C, Grzych, G. Plasma methionine and clinical severity in nitrous oxide consumption. Toxics 2022;11:12. https://doi.org/10.3390/toxics11010012.Search in Google Scholar PubMed PubMed Central
103. Singh, SK, Misra, UK, Kalita, J, Bora, HK, Murthy, RC. Nitrous oxide related behavioral and histopathological changes may be related to oxidative stress. Neurotoxicology 2015;48:44–9. https://doi.org/10.1016/j.neuro.2015.03.003.Search in Google Scholar PubMed
104. Grzych, G, Deheul, S, Gernez, E, Davion, JB, Dobbelaere, D, Carton, L, et al.. Comparison of biomarker for diagnosis of nitrous oxide abuse: challenge of cobalamin metabolic parameters, a retrospective study. J Neurol 2023;270:2237–45. https://doi.org/10.1007/s00415-023-11570-z.Search in Google Scholar PubMed
105. Einsiedler, M, Voulleminot, P, Demuth, S, Kalaaji, P, Bogdan, T, Gauer, L, et al.. A rise in cases of nitrous oxide abuse: neurological complications and biological findings. J Neurol 2022;269:577–82. https://doi.org/10.1007/s00415-021-10702-7.Search in Google Scholar PubMed PubMed Central
106. Chen, X, Ren, F, Xu, J, Yu, Z, Lin, X, Bai, Z, et al.. A rapid quantitative chemiluminescence immunoassay for vitamin B12 in human serum. Clin Lab 2020;66. https://doi.org/10.7754/clin.lab.2019.190604.Search in Google Scholar PubMed
107. Jiang, Y, Mistretta, B, Elsea, S, Sun, Q. Simultaneous determination of plasma total homocysteine and methionine by liquid chromatography-tandem mass spectrometry. Clin Chim Acta Int J Clin Chem 2017;464:93–7. https://doi.org/10.1016/j.cca.2016.11.017.Search in Google Scholar PubMed
108. Suen, KFK, Lee, GR, Finnegan, M, Halton, K, Borovickova, I, Trench, C, et al.. Total plasma homocysteine measurement: evaluation of the Abbott immunoassay, comparison with the JEOL ion exchange chromatography and investigation of its clinical utility. Pract Lab Med 2022;32:e00295. https://doi.org/10.1016/j.plabm.2022.e00295.Search in Google Scholar PubMed PubMed Central
109. Tan, Y, Sun, X, Tang, L, Zhang, N, Han, Q, Xu, M, et al.. Automated enzymatic assay for homocysteine. Clin Chem 2003;49:1029–30. https://doi.org/10.1373/49.6.1029.Search in Google Scholar PubMed
110. Lo, SY, Gordon, C, Sadilkova, K, Jack, RM, Dickerson, JA. Quantifying MMA by SLE LC-MS/MS: unexpected challenges in assay development. Clin Biochem 2016;49:967–72. https://doi.org/10.1016/j.clinbiochem.2016.05.010.Search in Google Scholar PubMed
111. Aronson, JK, Ferner, RE. Biomarkers—a general review. Curr Protoc Pharmacol 2017;76:9.23.1–9.23.17. https://doi.org/10.1002/cpph.19.Search in Google Scholar PubMed
112. Califf, RM. Biomarker definitions and their applications. Exp Biol Med 2018;243:213–21. https://doi.org/10.1177/1535370217750088.Search in Google Scholar PubMed PubMed Central
113. Henderson, KA. Occupational exposure of midwives to nitrous oxide on delivery suites. Occup Environ Med 2003;60:958–61. https://doi.org/10.1136/oem.60.12.958.Search in Google Scholar PubMed PubMed Central
114. Imbriani, M, Ghittori, S, Pezzagno, G, Capodaglio, E. Anesthetic in urine as biological index of exposure in operating-room personnel. J Toxicol Environ Health 1995;46:249–60. https://doi.org/10.1080/15287399509532032.Search in Google Scholar PubMed
115. Accorsi, A, Valenti, S, Barbieri, A, Raffi, G, Violante, F. Proposal for single and mixture biological exposure limits for sevoflurane and nitrous oxide at low occupational exposure levels. Int Arch Occup Environ Health 2003;76:129–36. https://doi.org/10.1007/s00420-002-0379-4.Search in Google Scholar PubMed
© 2024 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- Editorial
- External quality assurance (EQA): navigating between quality and sustainability
- Reviews
- Molecular allergology: a clinical laboratory tool for precision diagnosis, stratification and follow-up of allergic patients
- Nitrous oxide abuse direct measurement for diagnosis and follow-up: update on kinetics and impact on metabolic pathways
- Opinion Papers
- A vision to the future: value-based laboratory medicine
- Point-of-care testing, near-patient testing and patient self-testing: warning points
- Navigating the path of reproducibility in microRNA-based biomarker research with ring trials
- Point/Counterpoint
- Six Sigma – is it time to re-evaluate its value in laboratory medicine?
- The value of Sigma-metrics in laboratory medicine
- Genetics and Molecular Diagnostics
- Analytical validation of the amplification refractory mutation system polymerase chain reaction-capillary electrophoresis assay to diagnose spinal muscular atrophy
- Can we identify patients carrying targeted deleterious DPYD variants with plasma uracil and dihydrouracil? A GPCO-RNPGx retrospective analysis
- General Clinical Chemistry and Laboratory Medicine
- Comparison of ChatGPT, Gemini, and Le Chat with physician interpretations of medical laboratory questions from an online health forum
- External quality assessment performance in ten countries: an IFCC global laboratory quality project
- Multivariate anomaly detection models enhance identification of errors in routine clinical chemistry testing
- Enhanced patient-based real-time quality control using the graph-based anomaly detection
- Performance evaluation and user experience of BT-50 transportation unit with automated and scheduled quality control measurements
- Stability of steroid hormones in dried blood spots (DBS)
- Quantification of C1 inhibitor activity using a chromogenic automated assay: analytical and clinical performances
- Reference Values and Biological Variations
- Time-dependent characteristics of analytical measurands
- Cancer Diagnostics
- Expert-level detection of M-proteins in serum protein electrophoresis using machine learning
- An automated workflow based on data independent acquisition for practical and high-throughput personalized assay development and minimal residual disease monitoring in multiple myeloma patients
- Cardiovascular Diseases
- Analytical validation of the Mindray CL1200i analyzer high sensitivity cardiac troponin I assay: MERITnI study
- Diabetes
- Limitations of glycated albumin standardization when applied to the assessment of diabetes patients
- Patient result monitoring of HbA1c shows small seasonal variations and steady decrease over more than 10 years
- Letters to the Editor
- Inaccurate definition of Bence Jones proteinuria in the EFLM Urinalysis Guideline 2023
- Use of the term “Bence-Jones proteinuria” in the EFLM European Urinalysis Guideline 2023
- Is uracil enough for effective pre-emptive DPD testing?
- Reply to: “Is uracil enough for effective pre-emptive DPD testing?”
- Accurate predictory role of monocyte distribution width on short-term outcome in sepsis patients
- Reply to: “Accurate predictory role of monocyte distribution width on short-term outcome in sepsis patients”
- Spurious parathyroid hormone (PTH) elevation caused by macro-PTH
- Setting analytical performance specifications for copeptin-based testing
- Serum vitamin B12 levels during chemotherapy against diffuse large B-cell lymphoma: a case report and review of the literature
- Evolution of acquired haemoglobin H disease monitored by capillary electrophoresis: a case of a myelofibrotic patient with a novel ATRX mutation
Articles in the same Issue
- Frontmatter
- Editorial
- External quality assurance (EQA): navigating between quality and sustainability
- Reviews
- Molecular allergology: a clinical laboratory tool for precision diagnosis, stratification and follow-up of allergic patients
- Nitrous oxide abuse direct measurement for diagnosis and follow-up: update on kinetics and impact on metabolic pathways
- Opinion Papers
- A vision to the future: value-based laboratory medicine
- Point-of-care testing, near-patient testing and patient self-testing: warning points
- Navigating the path of reproducibility in microRNA-based biomarker research with ring trials
- Point/Counterpoint
- Six Sigma – is it time to re-evaluate its value in laboratory medicine?
- The value of Sigma-metrics in laboratory medicine
- Genetics and Molecular Diagnostics
- Analytical validation of the amplification refractory mutation system polymerase chain reaction-capillary electrophoresis assay to diagnose spinal muscular atrophy
- Can we identify patients carrying targeted deleterious DPYD variants with plasma uracil and dihydrouracil? A GPCO-RNPGx retrospective analysis
- General Clinical Chemistry and Laboratory Medicine
- Comparison of ChatGPT, Gemini, and Le Chat with physician interpretations of medical laboratory questions from an online health forum
- External quality assessment performance in ten countries: an IFCC global laboratory quality project
- Multivariate anomaly detection models enhance identification of errors in routine clinical chemistry testing
- Enhanced patient-based real-time quality control using the graph-based anomaly detection
- Performance evaluation and user experience of BT-50 transportation unit with automated and scheduled quality control measurements
- Stability of steroid hormones in dried blood spots (DBS)
- Quantification of C1 inhibitor activity using a chromogenic automated assay: analytical and clinical performances
- Reference Values and Biological Variations
- Time-dependent characteristics of analytical measurands
- Cancer Diagnostics
- Expert-level detection of M-proteins in serum protein electrophoresis using machine learning
- An automated workflow based on data independent acquisition for practical and high-throughput personalized assay development and minimal residual disease monitoring in multiple myeloma patients
- Cardiovascular Diseases
- Analytical validation of the Mindray CL1200i analyzer high sensitivity cardiac troponin I assay: MERITnI study
- Diabetes
- Limitations of glycated albumin standardization when applied to the assessment of diabetes patients
- Patient result monitoring of HbA1c shows small seasonal variations and steady decrease over more than 10 years
- Letters to the Editor
- Inaccurate definition of Bence Jones proteinuria in the EFLM Urinalysis Guideline 2023
- Use of the term “Bence-Jones proteinuria” in the EFLM European Urinalysis Guideline 2023
- Is uracil enough for effective pre-emptive DPD testing?
- Reply to: “Is uracil enough for effective pre-emptive DPD testing?”
- Accurate predictory role of monocyte distribution width on short-term outcome in sepsis patients
- Reply to: “Accurate predictory role of monocyte distribution width on short-term outcome in sepsis patients”
- Spurious parathyroid hormone (PTH) elevation caused by macro-PTH
- Setting analytical performance specifications for copeptin-based testing
- Serum vitamin B12 levels during chemotherapy against diffuse large B-cell lymphoma: a case report and review of the literature
- Evolution of acquired haemoglobin H disease monitored by capillary electrophoresis: a case of a myelofibrotic patient with a novel ATRX mutation