Next Article in Journal
Enhancing the Startup Rate of Microbial Methanogenic Systems through the Synergy of β-lactam Antibiotics and Electrolytic Cells
Previous Article in Journal
Taxonomic Diversity and Functional Traits of Soil Bacterial Communities under Radioactive Contamination: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

How Does Airway Surface Liquid Composition Vary in Different Pulmonary Diseases, and How Can We Use This Knowledge to Model Microbial Infections?

by
Dean Walsh
*,
Jennifer Bevan
and
Freya Harrison
School of Life Sciences, University of Warwick, Coventry CV4 7AL, UK
*
Author to whom correspondence should be addressed.
Current address: Apollo, OPEN Health Communications LLP, London EC4M 7AN, UK.
Microorganisms 2024, 12(4), 732; https://doi.org/10.3390/microorganisms12040732
Submission received: 11 March 2024 / Revised: 26 March 2024 / Accepted: 28 March 2024 / Published: 3 April 2024
(This article belongs to the Special Issue Microbe–Host Interactions in Human Infections)

Abstract

:
Growth environment greatly alters many facets of pathogen physiology, including pathogenesis and antimicrobial tolerance. The importance of host-mimicking environments for attaining an accurate picture of pathogen behaviour is widely recognised. Whilst this recognition has translated into the extensive development of artificial cystic fibrosis (CF) sputum medium, attempts to mimic the growth environment in other respiratory disease states have been completely neglected. The composition of the airway surface liquid (ASL) in different pulmonary diseases is far less well characterised than CF sputum, making it very difficult for researchers to model these infection environments. In this review, we discuss the components of human ASL, how different lung pathologies affect ASL composition, and how different pathogens interact with these components. This will provide researchers interested in mimicking different respiratory environments with the information necessary to design a host-mimicking medium, allowing for better understanding of how to treat pathogens causing infection in these environments.

1. Introduction

Airway surface liquid (ASL) is a thin film of fluid lining the epithelium of the trachea, bronchi, and bronchioles [1]. ASL comprises two layers [1,2]: the periciliary layer (PCL) which ranges between 5 and 15 µm in depth [3] and the mucus layer (MCL) which can reach 70 µm in thickness [4]. The PCL is a watery layer that bathes the cilia, and may be the only layer of ASL present in healthy adults [5]. However, once the airways are irritated, submucosal and goblet cells in the tracheal and bronchial epithelium release MUC5B and MUC5AC, which form the bulk of the mucus gel-like layer that forms the MCL [1]. This layer is likely to be thicker in some regions than others [6]. In the bronchioles, goblet and submucosal cells are instead replaced by Club cells, which produce MUC5B [7]. ASL and mucociliary clearance are essential in maintaining a very low microbial load and healthy airway functioning [6]. The low viscosity of the PCL allows for ciliary beating, facilitating the movement of the mucus blanket as it is propelled by the tips of the cilia. Cilia have been shown to not only move the mucus sheet of ASL, but also the periciliary sol [1,2,8]. The mucus entraps inhaled particles and microorganisms. The MCL and immobilised particles are directed via cilia to the mouth, where the resulting phlegm is then swallowed or expectorated [8]. Cough clearance also clears mucus independently of cilia activity [9]. An increased height and volume of the PCL is essential for an efficient cough clearance function; conversely, an excessively viscous mucus gel reduces airway clearance via coughing alone [10].
ASL contains numerous microbial growth factors including sugars, protein sources, electrolytes, and lipids [11]. It can take hours to clear airborne pathogens entrapped in mucus; during this time, pathogens can use these components to proliferate. To combat this, antimicrobial peptides including lysozyme, lactoferrin, and defensins are deployed to inhibit pathogen growth [11,12]. Additionally, peptide, amino acid, and sugar transporters can sequester nutrients present in the airway, inhibiting pathogen growth by limiting nutrient availability [12,13]. Despite these innate airway defences, infection does occur, particularly when pre-existing disease or comorbidities compromise these defences. Changes in ASL composition have been shown to predispose individuals to infection [14]. For example, Buonfiglio and colleagues have shown that the increased iron present in the ASL of smokers promoted the growth and biofilm formation of Staphylococcus aureus and Pseudomonas aeruginosa [14]. Likewise, excess mucin in the ASL mucus of mechanically ventilated patients can impair neutrophil function and consequently increase bacterial growth [15].
Numerous host-mimicking growth media have been developed to mimic the nutrition of different host environments, allowing for the monitoring of how these growth environments alter the phenotypes of pathogens [16]. These include artificial urine [17], artificial saliva [18], and synthetic wound fluid [19] for the modelling of urinary tract infections [20], periodontitis [21], and chronic wound infections [22], respectively. For modelling the pulmonary environment, studies investigating both chronic and acute infection of the cystic fibrosis (CF) lung dominate, with a large array of CF sputum-mimicking media described [16,23,24,25,26,27,28,29,30,31]. However, CF sputum is not representative of the ASL and sputum composition in other chronic pulmonary diseases. The concentrations of various macromolecules, sputum viscoelasticity, biomarker profiles, and sputum hydration differ relative to samples from patients with asthma exacerbations, chronic obstructive pulmonary disease (COPD), bronchitis, and bronchiectasis [32,33,34,35,36]. To our knowledge, host-mimicking media simulating these non-CF pulmonary disease environments are yet to be developed, although Ruhleul and colleagues have recently developed healthy lung and healthy sinus-mimicking media [26]. This makes it difficult to mimic the environment of other respiratory infections, such as ventilator-associated pneumonia (VAP). Neglecting the modelling of these infections is concerning.
In VAP, biofilms form around the cuff of the intubated endotracheal tube (ETT) and pathogens use ASL and saliva for growth [37,38]. VAP is defined as a pneumonia occurring after >48 h of mechanical ventilation; incidence varies between countries and hospitals but it is a very common infection in intensive care units with a high mortality rate (25–45%, [39,40]). The global economic impact of biofilms in ventilated patients is an estimated USD920 million per year [41]. VAP has been of particular concern following the coronavirus disease 2019 (COVID-19) pandemic, because the incidence of VAP in COVID-19 patients is much higher (50–80%) [42] than in non-COVID-19 patients (5–40%) [43,44].
Due to the lack of in vitro options for simulating the VAP environment, researchers have turned to murine VAP models [45] or ventilated pig models [46] to answer questions regarding the development of VAP. These experiments can be undesirable due to ethical considerations and the notable differences between animal and human ASL [47,48,49,50,51]. PCL composition has been broadly defined (96% water, 1% salts, 1% proteins, 1% lipids, 1% mucus), and the MCL is composed of a mixture of debris and polypeptides tethered by mucins [2,52]. However, the lack of specific information on components and concentrations hampers the development of in vitro VAP models and artificial ASL growth medium. Here, we compile and summarise the findings of studies looking at the composition of both healthy human ASL and the ASL of patients with various lung diseases, aiding researchers in the development of artificial ASL media that can mimic the chemical profile of different pulmonary diseases. This will allow for the development of infection models simulating different pulmonary diseases, providing a better understanding of how to prevent and treat infections in patients with specific pulmonary comorbidities.

2. Mucin

Mucins are large, highly glycosylated (≥50% carbohydrate wt/wt) proteins that form the main macromolecular components of mucus, therefore playing an essential role in the innate defence of the airways [53]. Mucus is a protective lining of healthy airways, whilst sputum is defined as a mucus produced in inflamed airways that also includes cells, inflammatory mediators, bacteria, and DNA derived from inflammatory cell necrosis [54,55]. Within sputum, the secreted mucins MUC5AC and MUC5B dominate [56]. In tracheobronchial surface epithelium, MUC5AC is secreted by goblet cells, whereas MUC5B is produced by submucosal cells [57]. Mucin glycosylation is essential to the rheological and viscoelastic properties of mucus, with mucins in the viscous mucus secretions of people with CF having more sialylated and sulphated O-glycans relative to the mucins in secretions of non-diseased individuals [58]. Changes in mucus viscosity are seen in other chronic bronchial diseases, including COPD and asthma [59].
The sputum of patients experiencing CF exacerbations contained ~90% more MUC5AC and ~30% more MUC5B than CF stable patients. Mucus taken from the endotracheal tubes (ETT) of patients with no lung disease actually produced proportions of mucin similar to those found in the sputum of patients experiencing CF exacerbations (12% more MUC5AC in CF exacerbation vs. ETT, 4% less MUC5B in CF exacerbation vs. ETT) [55]. As in previous studies [55,60], Henderson and colleagues also used immunological techniques to show that the sputum of CF patients contained significantly lower concentrations of MUC5B than non-CF sputum. Further investigation showed that human neutrophil elastase, a potent protease found in the inflamed airways of patients with CF [61], VAP [62], COPD, severe asthma, and bronchiectasis [63], could cleave the exposed protein region of MUC5AC and MUC5B, accounting for the lower mucin concentrations in CF sputum when immunological techniques were used [61].
The subglottic mucus of mechanically ventilated ICU patients had significantly higher mucus viscosity and concentrations of MUC5B relative to controls. There were no notable differences in MUC5AC concentrations between the two groups. Mucin concentrations showed a significant, weakly positive correlation with the duration of ventilation and patient age [15]. The dominance of MUC5B is also seen in COPD patients, whilst MUC5AC is more prevalent in the airways of smokers [64]. Mucin hypersecretion aids innate airway defences in engulfing inhaled particles, but prolonged mucin hypersecretion, as in critically ill ventilated patients, leads to a viscous accumulation of mucus and the failure of mucociliary clearance. This is due to a combination of factors including the compression of the PCL by a more concentrated mucus layer, inhibiting the lubrication of mucus and causing the adherence of mucus to the epithelium [61]. This is further compounded by the failure of the mucociliary escalator resulting in the redirection of secretions [65]. The stagnation and accumulation of mucus was shown to induce mucin-mediated neutrophil dysfunction and promote bacterial growth [15], particularly as pathogens such as P. aeruginosa [66] and Streptococcus pneumoniae [67] exploit mucin as a nutrient source.
The pathogen P. aeruginosa employs dedicated lectins and adhesins to exploit the altered glycosylation of mucins, whose deviated glycans are used as receptors to enable bacterial adhesion [68]. MUC5AC attenuates P. aeruginosa virulence by dispersing biofilms, downregulating type III and VI secretion, suppressing quorum sensing [69] and phenazine biosynthesis [70]. However, it has been shown that in certain lung pathologies such as CF, the antimicrobial properties of mucins are greatly diminished. This has been attributed to altered mucin production, mucin degradation, the altered diffusion of microbial products due to mucus layer dehydration, and changes in mucin glycosylation, therefore altering the signalling potential of mucin [55,60,71,72,73,74]. These alterations to mucin structure and function may explain why P. aeruginosa can exhibit abundant quorum-sensing signalling and biofilm formation in CF sputum [75], despite unaltered mucins suppressing quorum sensing and dispersing P. aeruginosa biofilms [69]. Pyocyanin, a toxin produced by P. aeruginosa, is known to stimulate the hypersecretion of mucin from goblet cells [76]. S. aureus utilises serine protease A (SplA) to degrade mucin, therefore facilitating lung invasion in vivo [77]. In the opportunistic fungal pathogen Candida albicans, mucin suppresses filamentation, preventing killing by P. aeruginosa, which attacks hyphal C. albicans [78]. Both Haemophilus influenzae and Streptococcus pneumoniae can bind to mucin, facilitating the penetration of the mucosal layer, the colonisation of host epithelium cells, and biofilm formation in the airways [79,80]. The negatively charged capsular polysaccharide of S. pneumoniae repels the sialic acid-rich mucopolysaccharides of mucin, preventing entrapment [79].
Existing CF sputum-mimicking media have included mucin at 5 mg/mL [26,27,31,81], 10 mg/mL [24], or 20 mg/mL [30]. For media attempting to model ventilated airways, it may be appropriate to exceed these mucin concentrations, as Powell et al. found that the average total mucin content of ETT aspirates was ~25 mg/mL [15]. Ruhleul and colleagues also produced a healthy lung medium alongside their CF lung medium. The healthy medium contained 1.2 mg/mL mucin [26], based on the findings of Henderson et al. [61]. Porcine gastric mucin (PGM) was used in most studies [24,26,27,31], although one study did use bovine submaxillary mucin (BSM) [81]. PGM predominantly consists of MUC5AC, although MUC6 and MUC5B are also present [82,83], whereas BSM consists of MUC5B and MUC19 [84]. Due to this, BSM may be more appropriate when trying to model conditions in which excess MUC5B is produced, such as VAP and COPD [15,56].
One disadvantage of commercially available mucins is that they can no longer form hydrogels, limiting their rheological and antimicrobial relevance compared to native mucin [85]. Furthermore, commercially available PGM contains various contaminants, including lipids, peptides, amino acids, and metals. This can complicate experiments: a previous study investigating the mucin-enhanced virulence of Acinetobacter baumannii in the mouse lung found that mucin-enhanced virulence was at least partially due to iron present in the supplemented PGM [86]. Another study found that iron derived from PGM supplementation raised iron to a more clinically relevant concentration relative to the CF lung [87]. Although BSM is purified in native conditions [88] whilst PGM is not [89], the higher expense and reduced quantity available to purchase of this mucin still makes PGM the most practical mucin source for studies that require high volumes of mucin-containing medium. Despite these limitations, the use of commercially available mucins is likely to continue, as purifying native mucins is labour intensive and produces a low yield [90].

3. DNA

DNA present in airway surface liquid and sputum is typically derived from inflammatory responses. Neutrophils are essential for the protection of the airways and produce neutrophil extracellular traps (NETs) in response to invading pathogens. These NETs are composed of DNA complexed with a wide array of antimicrobial proteins [91,92]. Although a key component of innate airway defence, NETs can cause severe lung injury [92,93,94,95]. This is further compounded during mechanical ventilation, which, in addition to being injurious to the lungs in itself, increases NET production [95,96]. The combination of NETs with nucleic acids from dead neutrophils, bacteria, and to a lesser degree, airway epithelia, can lead to the accumulation of a tenacious mucus that obstructs the airways [97,98]. DNA can associate with mucins in the sputum of CF patients [99]; the presence of DNA in mucus makes it much more viscous, reducing mucociliary clearance [60]. Likewise, extracellular DNA has been shown to regulate the expression of numerous virulence and metabolic genes in P. aeruginosa and is an essential structural polymer in the biofilm matrix [100]. Matrix DNA is known to aid in the survival of bacteria within biofilms by promoting antibiotic tolerance through the acidification of the environment and shielding P. aeruginosa from aminoglycosides [101]. S. aureus isolates chronically infecting individuals with CF express higher levels of nuclease activity, allowing S. aureus to persist in the neutrophil-rich environment by degrading NETs [102].
When developing media to mimic healthy sinuses and healthy lungs, Ruhleul et al. incorporated 0.96 mg/mL of DNA into the final recipe, based on quantities found in the sputum of healthy patients by Henke and colleagues [55]. Numerous formulations of CF-mimicking media contained 4 mg/mL of DNA [26,28,29,31]. Some recipes decreased the DNA concentration to 1.4 mg/mL [24,25] or even 0.6 mg/mL (refs) to better reflect the results shown by Brandt et al. in which the DNA was quantified from a large number of CF sputum samples (ranging from 0 to 9.5 mg/mL) [103]. Deciding on an appropriate DNA concentration for future ASL-mimicking media is difficult, due to the large variance across multiple studies, even those using similar methodologies (Table 1) [15,55,92,103,104].

4. pH and Ion Concentrations

Airway pH regulates ion transporters, influencing the movement of water and salt between cells and the ASL. Consequently, the effect of pH on the equilibrium between the secretion of chloride ions (Cl) and the absorption of sodium ions (Na+) dictates ASL volume and hydration [2,122]. The pH of ASL in healthy individuals is recorded as ranging from 6.78 to 7.1 [106,107]. People with CF, who have mutations in their cystic fibrosis transmembrane conductance regulator (CFTR) gene encoding an anion channel, have a reduced ASL pH [122,123], and their airway mucus becomes much thicker and more viscous. Many studies have measured the pH of CF respiratory mucus and collectively found values ranging between 6.0 and 7.4 [16,52,106,107,108,109,124]. Likewise, other disease states feature ASL acidification including pneumonia [111], COPD [112], and chronic lung disease [110]. Reducing pH below 7.0 reduces the activity of ASL antimicrobials, reduces the frequency of ciliary beating, and increases mucus viscosity [123,125]. ASL acidification also impairs S. aureus. clearance from human primary bronchial epithelial cells obtained from lobectomies [126] and P. aeruginosa clearance from a murine lung infection model [127]. Fungal pathogenesis is highly dependent upon pH, with many fungi, including Candida species and Aspergillus species able to utilise acidification and alkalinisation mechanisms to facilitate colonisation. However, this is still largely unexplored in the lung environment [128]. Ion concentrations have also been quantified, with elevated concentrations of both Na+ and Cl in CF tracheal aspirates [63,107,109,129].
Due to the variance in the reported pH values of CF sputum [16,52,106,107,108,109,124], CF sputum-mimicking media have used various pH values including 6.9 [26,28,29], 7 [31], 6.5 [24,25,30], and 6.8 [16,26,81]. NaCl concentrations in most of these media have remained largely unchanged since they were first incorporated into the artificial sputum medium (ASM) recipe of Ghani and Soothill [24,25,28,29,30,31], though later CF-mimicking media, such as the synthetic CF mucus media (SCFM) developed in the Whiteley laboratory, were used ion concentrations determined by those detected in sputum by Palmer and colleagues [16,81]. Likewise, Ruhleul and colleagues [26] used the works of Goldman et al. to determine the NaCl concentrations in their healthy sinus, healthy lung, CF sinus, and CF lung media [130].

5. Sugars and Other Carbon Sources

Glucose is present in the ASL of healthy individuals at ~0.4 mM, approximately 12X lower than its blood concentration [131]. Numerous factors can raise ASL glucose concentrations, including viral infection and diabetes mellitus [114]. Increases in breath glucose are seen in hyperglycaemic diabetes patients and CF patients, with hyperglycaemic CF patients having an even higher airway glucose concentration [113]. High blood glucose has been found to be the single most important factor determining whether patients develop exacerbations of COPD [132], severe COVID-19, or acute respiratory distress syndrome (ARDS); whether ICU patients require ventilation; or whether COVID-19 infection proves fatal [133,134,135,136]. Higher blood glucose concentrations have been shown to result in higher endotracheal glucose concentrations [114]. One study looking at 58 critically ill ventilated patients found that the glucose concentration of bronchial aspirates ranged between 2.7 and 4.4 mM [115].
The low glucose concentration in healthy ASLs ensures invading airborne pathogens are deprived of available carbon. This is maintained by tight junctions between airway epithelial cells, limiting the paracellular movement of glucose [137]. Furthermore, human airway epithelia are able to generate a transepithelial glucose concentration gradient, resulting in an ASL with a lower glucose concentration than blood [13]. Using a human airway epithelia culture model, Pezzulo and colleagues were able to show that concentrations of glucose similar to that of healthy ASL limited the growth of P. aeruginosa to the point that human airway epithelia were able to kill off the invading inoculum. However, when a hyperglycaemic murine lung infection model was used, increased susceptibility to P. aeruginosa was observed, caused by increased ASL glucose concentrations [13]. Likewise, when airway epithelial cells were exposed to pro-inflammatory mediators, ASL glucose concentrations increased due to the enhanced permeability of tight junctions [131]. This is further illustrated by the finding that treating hyperglycaemic mice with metformin, a drug that reduces tight junction permeability, reduced both airway glucose and airway bacterial load [138]. The increased availability of glucose in the ASL of inflamed and/or hyperglycaemic airways could explain the poorer outcomes in diabetic patients with community-acquired pneumonia [139]. Ventilated patients with higher ASL glucose concentrations were significantly more likely to be infected with pathogenic bacteria, particularly methicillin-resistant S. aureus (MRSA) [115].
In previously discussed CF sputum-mimicking medium, glucose has been used at 3 mM [16,81] following the quantification of glucose in CF sputum samples [16]. Ruhleul and colleagues used 1.2 mM glucose for both their CF sinus and lung media [26], following the finding that the nasal glucose concentration of CF patients averaged 1.2 ± 0.9 mM [140]. When designing media to mimic healthy airways, glucose concentrations of ~0.4 mM would be most appropriate, as this has been shown to be the resting glucose concentration of ASL in healthy individuals [131]. When designing media modelling of infected, inflamed, or ventilated airways, using 2–4.4 mM glucose may be more appropriate based on the findings of previous studies [113,115].
Little research has investigated other carbon sources in ASLs. One study found that 15 mM fructose induced a higher growth of S. aureus on human epithelial cells relative to 20 mM glucose; however, no studies have yet quantified fructose in human ASL [141]. Mucin degradation can provide an additional source of sugars and amino acids to act as carbon sources. Mucins are comprised of numerous monosaccharides, including N-acetylglucosamine (GlcNAc; 32% of mucin dry weight), galactose (29% dry weight), sialic acid, fucose, and N-acetylgalactose (GalNAc) [58]. The increased glycosylation and sulfation of mucins has been observed in both CF and VAP [58,142], whilst a positive correlation has been observed between airway-infection severity and the increased sialylation of mucins [143]. The increase in mucin modification in diseased states consequently means that higher sugar concentrations are available for bacteria to use as carbon sources following mucin degradation [58,144,145,146].
S. aureus isolates that can better utilise free sialic acid, such as isolates deficient in the Agr quorum-sensing system, are better adapted to the lung and are associated with chronic lung infection. Furthermore, sialic acid utilisation upregulates the production of the siderophore staphyloferrin, increasing iron acquisition [147]. Whilst S. aureus is incapable of liberating sialic acid from mucin itself, sialic acid can be cross-fed to S. aureus by residents of the airway microbiota, such as Streptococci or anaerobes [148]. Various respiratory pathogens use free sialic acid to undergo sialylation for the purpose of immune evasion. Sialylated P. aeruginosa are able to impede their trafficking to lysosomes when phagocytosed, persisting and replicating intracellularly in macrophages [149]. Furthermore, the binding of sialylated P. aeruginosa to neutrophils can suppress neutrophil activity, reducing ROS levels, NET formation, and elastase release [150]. Likewise, nontypeable Haemophilus influenzae (NTHi) sialylates lipooligosaccharides, protecting it from IgM and complement-mediated killing [151]. The presence of both sialic acid and GlcNAc has been shown to increase the biofilm formation of H. influenzae [152]. The sputum environment also induces GlcNAc catabolism by P. aeruginosa, upregulating the production of the phenazine antimicrobial pyocyanin, potentially mediating interspecies competition in the lung environment [146]. The inclusion of these additional carbon sources in host-mimicking media may be important to better represent how respiratory pathogens adapt to the respiratory environment. (Particularly if the medium includes mucin and the organisms of interest are incapable of liberating these carbon sources from mucin). For these reasons, some sputum-mimicking media have incorporated GlcNAc [26,81], galactose, and sialic acid [26].

6. Amino Acids

ASL contains numerous protein sources, including mucins and antimicrobial peptides, which are essential in innate lung defence [12]. To prevent excessive accumulation and subsequent airway obstruction by proteins and mucus, enzymatic degradation within the airways is employed [153]. This provides a wealth of peptides and amino acid sources within the ASL, which are in turn sequestered by various transporters in the lung epithelium. This cycle of protein degradation and peptide transport ensures that there is a constant, if limited, presence of amino acids within healthy ASLs [12]. The acquisition of branched-chain amino acids has been shown to aid in the colonisation of the nasopharynx and lungs by S. pneumoniae [154]. Amino acids are abundant in CF sputum, with P. aeruginosa preferring the assimilation of amino acids over sugars [155]. This leads to the emergence of P. aeruginosa mutants that are auxotrophic for various amino acids [156,157]. The catabolism of histidine has also been shown to been shown to be essential for A. baumannii virulence in a murine lung model [158]. Little research has been carried out to identify and quantify specific amino acids present in non-CF ASL. It has been noted that the amino acid of healthy lungs is lower than that of CF patients [116]. Likewise, total amino acid content was higher in CF patients suffering exacerbations than in clinically stable patients [117]. Higher levels of free amino acids are also observed in patients suffering from ARDS [159,160]. The increase in available amino acids in ASL during respiratory disease is hypothesised to be due to reduced amino acid transporter activity [12].
Amino acids have been fundamental for the development of effective ASM [16,24,25,26,27,29,30,31,81]. The earliest formulations of artificial sputum medium did not include amino acids [28], and this greatly impaired the biofilm formation of respiratory pathogens [161]. After adapting the findings by Ghani and Soothill [28], Sriramulu et al. began adding selected amino acids to ASM at equal ratios at 250 mg/L, and this was further implemented in subsequent ASM [24,27,29,31]. Conversely, some studies added casamino acids at 5 g/L [25,27] or 7.225× essential amino acids and 14.45× non-essential amino acids [30] instead of adding individual amino acids. Palmer et al. quantified free amino acids in CF sputum supernatants and found relatively consistent ratios of animo acids in samples from different patients, despite differences in overall concentrations. Following this, studies began to incorporate individual amino acids at a ratio reflecting this [16,26,81]. In Ruhleul et al.’s healthy sinus and healthy lung media, amino acid concentrations were determined by the findings of Schwab et al., who determined the amino acid content of non-CF airway secretions [26,162]. Whilst research into the amino acid concentrations in sputum has greatly benefitted the development of CF sputum-mimicking media, and to a lesser degree, media simulating healthy ASL, there is extremely little data on the amino acid concentrations of airways in other disease states, such as the ASL of ventilated patients. It would be logical to assume that the amino acid concentrations in ventilated ASL are higher than healthy ASL, as both people with CF and ARDS have higher sputum amino acid contents than healthy people [159,160,163].

7. Lipids

Until recently, lipids in the airways were seen only as an energy source or as essential structural components for membranes. Now, lipids are acknowledged to play key roles in mediating the airway immune response through various signalling mechanisms [164,165,166]. The lipid content of healthy ASL has broadly been quantified as approximately 1% of total ASL, and approximately a quarter of all solid content found within healthy ASL [2,52,161]. Likewise, lipids constituted approximately 30% of the dry material of sputum from patients with CF or asthma. Phosphatidlycholine has been identified as the main lipid present in sputum. However, sputum does include other lipid components including cholesterol, triglycerides, ceramides, sphingomyelin, and more [167]. Phosphatidylcholine is thought to act as a biosynthetic intermediate in the formation of phosphocholine-substituted structures on the surface of H. influenzae and S. pneumoniae [168]. Phosphocholine-modified lipoteichoic acids and teichoic acids of S. pneumoniae increase the cell adherence and invasion of the lung [169], whilst phosophocholine-modified lipopolysaccharides are associated with an increased persistence of H. influenzae on the airway mucosal surface [168,170]. Phosphatidylcholine utilisation via P. aeruginosa in mouse lung infection models increases the pathogen’s fitness, competitiveness, and aids P. aeruginosa adaptation to the lung environment [171]. Phosphatidylcholine and cholesterol in pulmonary surfactant have been shown to drive type-3 fimbria-mediated biofilm formation in Klebsiella. pneumoniae [172]. Lecithin, provided by egg yolk emulsion, has been the predominant lipid source in artificial CF sputum media [24,25,27,28,29,30,31], and has been universally used at 0.5% (v/v) in these formulations. Dioleoyl phosphatidylcholine has also been used as a lipid source at a concentration of 100 µg/mL [81]. The use of egg yolk lecithin would be more appropriate to provide an array of different lipids; however, this would need to be balanced with the likely diversity in composition between different suppliers and batches [173].

8. Antimicrobial Peptides and Enzymes

Antimicrobial peptides are released from epithelial cells, submucosal cells, resident and recruited macrophages, and are transported from plasma [174]. The most numerous of these antimicrobials in the airways are lysozyme and lactoferrin [118]. Lysozyme degrades peptidoglycan present in bacterial cell walls, causing lysis [175]. Lactoferrin functions by binding to iron, depriving bacteria of essential nutrition [176]. Lactoferrin can also permeabilise Gram-negative bacteria through interactions with lipopolysaccharide on the cell surface [177]. Concentrations of these antimicrobials can change depending on lung health. For instance, CF has been shown to increase concentrations of lactoferrin in bronchoalveolar lavage fluid (BALF) [61]. One study found no significant difference in lactoferrin concentration between smokers and non-smokers. However, because the BALF of smokers contained higher levels of iron relative to lactoferrin, these conditions were better able to stimulate bacterial growth and biofilm formation [14]. Sagel et al. also detected significant increases in lactoferrin and lysozyme in the BALF of culture-positive CF patients vs. culture-negative CF patients [118].
Other antimicrobial enzymes associated with neutrophils and NETs include myeloperoxidase, which is complexed with the DNA of NETs. Myeloperoxidase catalyses the formation of hypochlorous acid, which in turn kills invading pathogens [178]. Myeloperoxidase concentrations have been shown to increase in ARDS patients, VAP patients, and patients with both conditions (significantly so in the latter two groups) relative to patients with neither condition [92]. Other proteolytic enzymes released from neutrophils include human neutrophil elastase (HNE) and matrix metalloproteases (MMPs). VAP patients have been reported to have significantly elevated levels of HNE, MMP-8, and MMP-9 compared to non-VAP patients [179]. Another study also found significantly elevated levels of HNE in mechanically ventilated ICU patients compared to newly intubated controls [15].

9. Metals

Numerous metals are of significant importance in the airways. Calcium (Ca), magnesium (Mg), manganese (Mn), copper (Cu), and zinc (Zn) are important for inflammation [180]. Both Cu and Zn are essential for the functioning of anti-inflammatory superoxide dismutases [181]. Whilst these biometals act as essential co-factors for numerous host enzymes, they can also impact disease severity [182]. Iron (Fe) can cause significant lung damage by generating reactive oxygen species and oxidative stress, and increase bacterial virulence [183]. An iron-enriched lung environment impairs neutrophil function and inhibits the antimicrobial effects of lactoferrin and transferrin through binding saturation, all whilst providing pathogens like P. aeruginosa with the nutrition to readily replicate [182,184]. Likewise, excess Mg has been shown to induce neutrophil dysfunction, preventing both phagocytosis and oxidative bursts [185]. These metal ions are suspected to originate from vascular leakage, defects in ion channels, release from necrotic host cells or the lysis of invading microorganisms [183].
Increased metal concentrations have also been detected in numerous inflammatory pulmonary diseases (Table 2) [14,183,186]. Gray et al. quantified Zn, Fe, Cu, and Mn in the sputum of healthy individuals and patients with asthma, COPD, CF, and non-CF bronchiectasis [186]. Concentrations of Cu and Fe were higher in all disease groups compared to healthy controls, whist Zn was higher in all groups except asthmatic patients, and Mn was higher in all but COPD patients. Sputum Zn concentrations also had a strong positive correlation with the presence of lung inflammatory markers including calprotectin, IL-8, and myeloperoxidase [186]. Smith and colleagues found increases in Ca, Mg, Zn, Fe, and Cu concentrations in both CF and bronchiectasis patients relative to healthy controls. They also found that Zn, Fe, and Mg concentrations positively correlated with IL-8 [183], agreeing with the findings of Gray et al. [186]. Fe concentrations were also found to be approximately 4X higher in the BALF of smokers compared to non-smokers [14], whilst smoking status has been found to have little effect on the Zn and Fe concentrations of COPD patients [186].
Metals have not been incorporated into many synthetic CF media [24,25,27,28,29,30,31]. SCFM formulations do include FeSO4, CaCl2, and MgCl2 to provide a pool of metal ions [16,81], with Fe concentrations derived from the two studies by Stites et al. [187,188]. Ruhleul and colleagues used the findings of Smith et al. [183] for the basis of the metal concentrations found in their healthy sinus and lung media, as well as in their CF sinus and lung media [26]. Quinn et al. instead incorporated ferritin into their growth medium to better reflect the iron sources present in cystic fibrosis sputum [30].

10. Polyamines

Consecutive enzymatic reactions convert L-ornithine into the polyamines putrescine, spermidine, and then spermine [120]. Maintaining normal polyamine concentrations has been associated with preserving numerous essential cellular functions, including cell proliferation and differentiation, ion channel function, and protection against oxidative stress [189]. Polyamines are also synthesised by bacteria for the purposes of altering surface charges to aid antimicrobial and oxidative stress tolerance, protection against the phagolysosome, biofilm formation, and iron and free radical scavenging [172,190]. Polyamines can even be used as a sole carbon source by P. aeruginosa [191].
Polyamine concentrations in the lung have been shown to vary depending on disease status. For instance, significant increases in putrescine are observed in CF exacerbation patients relative to stable patients [192]. Furthermore, stable CF patients had significantly higher spermine concentrations relative to healthy controls. In addition, patients experiencing pulmonary exacerbation had significantly higher concentrations of both putrescine and spermidine relative to both healthy and CF-stable patients [120]. Metabolomic analyses also found 2–5 log fold increases in polyamines including spermine and spermidine in the endotracheal aspirates of VAP patients compared with samples taken pre-intubation [193]. Significantly higher levels of spermidine have also been observed in smokers and patients with COPD [194]. The elevated levels of polyamines in diseased airways can be exploited to P. aeruginosa, which can scavenge free polyamines from the environment [195] and utilise them to facilitate antimicrobial tolerance [195,196]. Polyamines have not yet been widely used in synthetic ASL formulations, though Ruhleul et al. have incorporated putrescine, spermidine, and spermine into formulations of healthy sinus, healthy lung, CF sinus, and CF lung media [26].

11. Serum Albumin

Albumin enters sputum through the vascular leakage that accompanies inflammatory pulmonary disease [119]. Albumin in ASL contributes to mucus plugging due to albumin acting as an alternative substrate for neutrophil proteases, and consequently inhibiting mucin degradation. This was evidenced by significantly higher proportions of albumin-degradation products found in the mucus of patients experiencing exacerbations of asthma. Furthermore, albumin can increase the viscoelasticity of airway mucus, impeding mucus clearance [197]. Some bacteria, including S. pneumoniae [198] can bind to albumin, exploiting albumin-derived fatty acids as a source of nutrition [199]. Yeast form C. albicans can also aggregate around albumin. This facilitates biofilm formation, decreases the penetration of antifungals through C. albicans-albumin plaques, and increases protection against phagocytic attack [200]. Albumin also neutralises the cytolytic toxin candidalysin produced by C. albicans, potentially reducing the cell damage induced by this pathogen [201]. In P. aeruginosa, albumin induces the expression of iron-controlled genes [202] and quenches the homoserine lactone quorum-sensing signal, attenuating the killing of S. aureus. via P. aeruginosa exoproducts [203]. Albumin levels are significantly elevated in both exacerbating and stable CF patients compared to patients with COPD or healthy individuals [119]. Another study also showed that sputum-producing CF patients produced nearly double the albumin found in the sputum of healthy controls [121]. Bovine serum albumin (BSA) has been incorporated into some previous sputum-mimicking media at 10 mg/mL [24,25]. Ruhleul et al. used BSA concentrations of 0.5 mg/mL for healthy sinus media, 1.5 mg/mL for healthy lung media, and 7 mg/mL for both CF sinus and CF lung media [26]. These concentrations were derived from a study that mimicked the environments for sinus colonisation and pneumonia using S. pneumoniae [144].

12. Airway Surface Liquid Component Interactions

Understanding how the components of an ASL-mimicking medium interact with each other is essential for interpretating results acquired using these host-mimicking models. PGM has been shown to inhibit the activity of both salivary lysozyme and hen egg-white lysozyme. In total. 2 mg/mL PGM inhibits approximately 40% of the activity of hen egg-white lysozyme, and 20% of the activity of unstimulated whole saliva lysozyme [204]. Mucin concentrations may exceed 2 mg/mL in many media seeking to mimic conditions such as CF exacerbations and VAP, so in these particular conditions, lysozyme inhibition via PGM is likely to be higher. An ASL-mimicking medium would ideally be pH 6–7, which is within the optimal pH range of lysozyme (pH 6–9) [205]. However, pH further impacts the degree of lysozyme inhibition via PGM, with lysozyme activity decreasing by 38.4% at pH 6, and 36.7% at pH 7, in the presence of PGM. Furthermore, many airway-mimicking models would be incubated at 37 °C, but hen egg-white lysozyme activity is inhibited by 45%, and unstimulated whole saliva lysozyme activity is inhibited by 20% after being incubated at 37 °C for 30 min with PGM [204]. Unlike PGM, BSM did not inhibit lysozyme activity [204], perhaps due to its increased purity relative to PGM [87]. Therefore, BSM may be the most appropriate mucin choice for researchers seeking to factor in the effect of lysozyme into their experiments. GlcNAc, which has previously been added to certain airway-mimicking media [26,81], may also lead to a decrease in lysozyme activity, because it is a competitive inhibitor of the enzyme [205,206]. Lysozyme is also known to bind to DNA, forming highly charged lysozyme-DNA complexes that alter nucleic function and make lysozyme a potent antiviral [207,208]. This antiviral activity is distinct from the catalytic mechanism responsible for the antibacterial activity of lysozyme [208]. It is currently unclear whether the presence of high levels of extracellular DNA in airway-mimicking media will alter lysozyme activity.
Albumin is able to bind and transport many endogenous and exogenous compounds. Due to this, serum albumin is also capable of binding to many other components that would be found in an ASL-mimicking medium, including fatty acids [209], DNA [210], polyamines [211], and metals [212]. Electrostatic interactions between PGM and BSA have been shown to produce mucin–albumin complexes. The formation of these complexes is highly pH-dependent, with mucin and BSA unable to bind at pH 3, but able to bind at pH 7.4, close to the pH used in airway-mimicking environments; BSA and mucin not only bind but cause changes to the secondary structure of mucin. The formation of mucin–BSA complexes alters the microrheology of the mucus layer in individuals experiencing cystic fibrosis, asthma, and ARDS, resulting in a mucus with much lower diffusivity [213]. The formation of mucin–BSA complexes is also facilitated by Ca2+ ions [213], which are provided by CaCl2 in numerous airway-mimicking media [16,26,81]. PGM has been found to bind lipids including free fatty acids, cholesterol, sphingomyelin, and phospholipids; these bound lipids shield mucin from attack by oxygen radicals [214]. Mucins, including commercially available PGM, have been shown to bind to metals, including iron [215,216] and zinc [215], with iron competitively inhibiting the binding of other metals to mucin [215]. Therefore, any PGM used in ASL-mimicking media may act as a sink for metal and lipid sources added to the medium. Any polyamines added to ASL-mimicking media could also sequester metal ions present in the medium [217]. Furthermore, the polycationic nature of polyamines can result in polyamine aggregation to extracellular DNA present in the growth medium [218], potentially limiting the uptake of polyamines by pathogens grown in ASL-mimicking environments.

13. Lung Microbiome

Much like the composition of ASL, the lung microbiome can dynamically change in response to disease, and is therefore an important factor to consider when designing models to mimic specific pulmonary diseases. The lung microbiome is not the core focus of this review, but more in-depth insights into the lung microbiome have been summarised comprehensively in other reviews [219,220]. Streptococcus, Veillonella, and Prevotella dominate the oral microbiome and are also among the most common genera in the lung microbiome of healthy individuals [220,221]. In the airways of individuals with asthma, COPD, and CF, these commensal species have a reduced abundance, and this is accompanied by an increase in pathogenic taxa including Pseudomonas, Staphylococcus, and Haemophilus [222,223,224,225,226,227]. Patients with VAP display a higher bacterial load, with reduced diversity in tracheal aspirates compared to controls. In VAP patients, Pseudomonas. and Corynebacterium were more abundant, whilst the presence of Prevotella and Streptococcus has been associated with patients that do not go on to develop VAP [228]. Numerous pathogens reside within the microbiome of protracted bacterial bronchitis patients, including H. influenzae, S. aureus, and S. pneumoniae Enrichments of commensal species of the genera Prevotella, Neisseria, and Streptococcus are also observed [229]. The impact of metabolites from the lung microbiome on lung disease is still largely unexplored, and will in time undoubtedly reveal important metabolites that should be included in ASL-mimicking media. However, the limited study so far into this area does reveal how microbiome and metabolomic data can be used for the development of models simulating different outcomes of a given disease. For instance, in the case of COPD, Streptococcus, Neisseria, and Veillonella in conjunction with metabolites such as polyamines, glycerophospholipids, and glycosphingolipids are associated with COPD and poorer lung function. Conversely, the presence of Prevotella with tyrosine and sialic acid are associated with either fewer COPD symptoms or no COPD [230]. As research linking the lung microbiome and lung metabolome with pulmonary disease progresses, the development of host-mimicking models will improve. Media should be assessed for their ability to support a microbiome typical of the condition they seek to model. This will facilitate not only the improved modelling of different pulmonary diseases, but will also allow for the simulation of how pathogens react in nutritional environments associated with different disease outcomes.

14. Future Developments and Conclusions

The composition of human ASL is complex and changes depending on lung health (summarised in Figure 1). Significant research into the CF lung has informed the development of CF sputum-mimicking media. Pathogens grown in these media display gene-expression profiles [16,26,81], the induction of mucoid phenotypes, and biofilm growth resembling that of pathogens growth in CF sputum [24,28,31]. In contrast, there has been only a single attempt to model the ASL of healthy individuals [26], and the development of media mimicking other inflammatory pulmonary conditions is completely lacking. The main reasons for this are, first, the relative lack of research investigating ASL composition in conditions other than CF, and, second, the fact that existing research often focuses on just one particular ASL component, such as amino acids or mucin [142,159,160]. This is partly explained by the relative ease of obtaining large volumes of sputum or BALF from people with CF, who historically have expectorated large quantities of sputum and undergone BAL as part of their medical treatment. Collecting samples from the thin layer of ASL present in healthy people is much harder; although the studies we have cited in this review used a range of methods to attempt this, including a variety of immunologic, chromatographic, and spectroscopic methods. Here, we seek to provide researchers with a comprehensive review of research investigating ASL and sputum composition across a spectrum of inflammatory pulmonary diseases, including ARDS, COPD, and VAP. Whilst CF and CF-mimicking media still feature heavily throughout, we hope this review will act as a reference point for researchers to see how information obtained through numerous studies can transition to the development of lung-mimicking media.

Author Contributions

D.W.: Investigation, Analysis, Writing—Original Draft; J.B.: Investigation, Analysis (Pilot Data); F.H.: Conceptualization, Writing—Review and Editing, Funding Acquisition, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the University of Warwick and Monash University, via the Monash-Warwick Alliance in Antimicrobial Resistance. J.B. contributed to this work during her MRes in Bioscience, funded by the Midlands Integrative Biosciences Training Partnership.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

ASLAirway surface liquid
PCLPericiliary layer
MCLMucus layer
CFCystic fibrosis
COPDChronic obstructive pulmonary disease
VAPVentilator-associated pneumonia
ETTEndotracheal tube
COVID-19Coronavirus disease 2019
ICUIntensive care unit
PGMPorcine gastric mucin
BSMBovine submaxillary mucin
NETNeutrophil extracellular trap
CFTRCystic fibrosis transmembrane conductance regulator
ASMArtificial sputum medium
SCFMSynthetic CF mucus media
ARDSAcute respiratory distress syndrome
MRSAMethicillin-resistant S. aureus
GlcNAcN-acetylglucosamine
GalNAcN-acetylgalactose
NTHiNontypeable Haemophilus influenzae
BALFBronchoalveolar lavage fluid
HNEHuman neutrophil elastase
MMPMatrix metalloproteases
BSABovine serum albumin

References

  1. Widdicombe, J.H. Regulation of the depth and composition of airway surface liquid. J. Anat. 2002, 201, 313–318. [Google Scholar] [CrossRef] [PubMed]
  2. Zajac, M.; Dreano, E.; Edwards, A.; Planelles, G.; Sermet-Gaudelus, I. Airway Surface Liquid pH Regulation in Airway Epithelium Current Understandings and Gaps in Knowledge. Int. J. Mol. Sci. 2021, 22, 3384. [Google Scholar] [CrossRef] [PubMed]
  3. Thiagarajah, J.R.; Song, Y.; Derichs, N.; Verkman, A.S. Airway surface liquid depth imaged by surface laser reflectance microscopy. J. Gen. Physiol. 2010, 136, 353–362. [Google Scholar] [CrossRef] [PubMed]
  4. Atanasova, K.R.; Reznikov, L.R. Strategies for measuring airway mucus and mucins. Respir. Res. 2019, 20, 261. [Google Scholar] [CrossRef] [PubMed]
  5. Bhaskar, K.R.; O’Sullivan, D.D.; Seltzer, J.; Rossing, T.H.; Drazen, J.M.; Reid, L.M. Density gradient study of bronchial mucus aspirates from healthy volunteers (smokers and nonsmokers) and from patients with tracheostomy. Exp. Lung Res. 1985, 9, 289–308. [Google Scholar] [CrossRef] [PubMed]
  6. Knowles, M.R.; Boucher, R.C. Mucus clearance as a primary innate defense mechanism for mammalian airways. J. Clin. Investig. 2002, 109, 571–577. [Google Scholar] [CrossRef] [PubMed]
  7. Rostami, M.R.; LeBlanc, M.G.; Strulovici-Barel, Y.; Zuo, W.; Mezey, J.G.; O’Beirne, S.L.; Kaner, R.J.; Leopold, P.L.; Crystal, R.G. Smoking shifts human small airway epithelium club cells toward a lesser differentiated population. NPJ. Genom. Med. 2021, 6, 73. [Google Scholar] [CrossRef] [PubMed]
  8. Wanner, A.; Salathe, M.; O’Riordan, T.G. Mucociliary clearance in the airways. Am. J. Respir. Crit. Care. Med. 1996, 154, 1868–1902. [Google Scholar] [CrossRef] [PubMed]
  9. Roomans, G.M.; Kozlova, I.; Nilsson, H.; Vanthanouvong, V.; Button, B.; Tarran, R. Measurements of airway surface liquid height and mucus transport by fluorescence microscopy, and of ion composition by X-ray microanalysis. J. Cyst. Fibros. 2004, 3 (Suppl. S2), 135–139. [Google Scholar] [CrossRef]
  10. King, M.; Zahm, J.M.; Pierrot, D.; Vaquez-Girod, S.; Puchelle, E. The role of mucus gel viscosity, spinnability, and adhesive properties in clearance by simulated cough. Biorheology 1989, 26, 737–745. [Google Scholar] [CrossRef]
  11. Amatngalim, G.D.; Hiemstra, P.S. Airway Epithelial Cell Function and Respiratory Host Defense in Chronic Obstructive Pulmonary Disease. Chin. Med. J. 2018, 131, 1099–1107. [Google Scholar] [CrossRef] [PubMed]
  12. Mager, S.; Sloan, J. Possible role of amino acids, peptides, and sugar transporter in protein removal and innate lung defense. Eur. J. Pharmacol. 2003, 479, 263–267. [Google Scholar] [CrossRef] [PubMed]
  13. Pezzulo, A.A.; Gutiérrez, J.; Duschner, K.S.; McConnell, K.S.; Taft, P.J.; Ernst, S.E.; Yahr, T.L.; Rahmouni, K.; Klesney-Tait, J.; Stoltz, D.A.; et al. Glucose Depletion in the Airway Surface Liquid Is Essential for Sterility of the Airways. PLoS ONE 2011, 6, e16166. [Google Scholar] [CrossRef] [PubMed]
  14. Vargas Buonfiglio, L.G.; Borcherding, J.A.; Frommelt, M.; Parker, G.J.; Duchman, B.; Vanegas Calderon, O.G.; Fernandez-Ruiz, R.; Noriega, J.E.; Stone, E.A.; Gerke, A.K.; et al. Airway surface liquid from smokers promotes bacterial growth and biofilm formation via iron-lactoferrin imbalance. Respir. Res. 2018, 19, 42. [Google Scholar] [CrossRef] [PubMed]
  15. Powell, J.; Garnett, J.P.; Mather, M.W.; Cooles, F.A.H.; Nelson, A.; Verdon, B.; Scott, J.; Jiwa, K.; Ruchaud-Sparagano, M.H.; Cummings, S.P.; et al. Excess Mucin Impairs Subglottic Epithelial Host Defense in Mechanically Ventilated Patients. Am. J. Respir. Crit. Care Med. 2018, 198, 340–349. [Google Scholar] [CrossRef] [PubMed]
  16. Palmer, K.L.; Aye, L.M.; Whiteley, M. Nutritional cues control Pseudomonas aeruginosa multicellular behavior in cystic fibrosis sputum. J. Bacteriol. 2007, 189, 8079–8087. [Google Scholar] [CrossRef] [PubMed]
  17. Stickler, D.J.; Morris, N.S.; Winters, C. Simple physical model to study formation and physiology of biofilms on urethral catheters. Methods. Enzymol. 1999, 310, 494–501. [Google Scholar] [CrossRef] [PubMed]
  18. Pratten, J.; Smith, A.W.; Wilson, M. Response of single species biofilms and microcosm dental plaques to pulsing with chlorhexidine. J. Antimicrob. Chemother. 1998, 42, 453–459. [Google Scholar] [CrossRef] [PubMed]
  19. Werthen, M.; Henriksson, L.; Jensen, P.O.; Sternberg, C.; Givskov, M.; Bjarnsholt, T. An in vitro model of bacterial infections in wounds and other soft tissues. APMIS 2010, 118, 156–164. [Google Scholar] [CrossRef]
  20. Maierl, M.; Jorger, M.; Rosker, P.; Reisner, A. In vitro Dynamic Model of a Catheterized Bladder and Biofilm Assay. Bio. Protoc. 2015, 5, e1381. [Google Scholar] [CrossRef] [PubMed]
  21. Millhouse, E.; Jose, A.; Sherry, L.; Lappin, D.F.; Patel, N.; Middleton, A.M.; Pratten, J.; Culshaw, S.; Ramage, G. Development of an in vitroperiodontal biofilm model for assessing antimicrobial and host modulatory effects of bioactive molecules. BMC. Oral Health 2014, 14, 80. [Google Scholar] [CrossRef] [PubMed]
  22. Furner-Pardoe, J.; Anonye, B.O.; Cain, R.; Moat, J.; Ortori, C.A.; Lee, C.; Barrett, D.A.; Corre, C.; Harrison, F. Anti-biofilm efficacy of a medieval treatment for bacterial infection requires the combination of multiple ingredients. Sci. Rep. 2020, 10, 12687. [Google Scholar] [CrossRef] [PubMed]
  23. Harrington, N.E.; Sweeney, E.; Harrison, F. Building a better biofilm-Formation of in vivo-like biofilm structures by Pseudomonas aeruginosa in a porcine model of cystic fibrosis lung infection. Biofilm 2020, 2, 100024. [Google Scholar] [CrossRef] [PubMed]
  24. Fung, C.; Naughton, S.; Turnbull, L.; Tingpej, P.; Rose, B.; Arthur, J.; Hu, H.; Harmer, C.; Harbour, C.; Hassett, D.J.; et al. Gene expression of Pseudomonas aeruginosa in a mucin-containing synthetic growth medium mimicking cystic fibrosis lung sputum. J. Med. Microbiol. 2010, 59, 1089–1100. [Google Scholar] [CrossRef] [PubMed]
  25. Hare, N.J.; Soe, C.Z.; Rose, B.; Harbour, C.; Codd, R.; Manos, J.; Cordwell, S.J. Proteomics of Pseudomonas aeruginosa Australian epidemic strain 1 (AES-1) cultured under conditions mimicking the cystic fibrosis lung reveals increased iron acquisition via the siderophore pyochelin. J. Proteome Res. 2012, 11, 776–795. [Google Scholar] [CrossRef] [PubMed]
  26. Ruhluel, D.; O’Brien, S.; Fothergill, J.L.; Neill, D.R. Development of liquid culture media mimicking the conditions of sinuses and lungs in cystic fibrosis and health. F1000Research 2022, 11, 1007. [Google Scholar] [CrossRef] [PubMed]
  27. Dinesh, S.D. Artificial Sputum Medium. Protoc. Exch. 2010. [Google Scholar] [CrossRef]
  28. Ghani, M.; Soothill, J.S. Ceftazidime, gentamicin, and rifampicin, in combination, kill biofilms of mucoid Pseudomonas aeruginosa. Can. J. Microbiol. 1997, 43, 999–1004. [Google Scholar] [CrossRef] [PubMed]
  29. Kirchner, S.; Fothergill, J.L.; Wright, E.A.; James, C.E.; Mowat, E.; Winstanley, C. Use of artificial sputum medium to test antibiotic efficacy against Pseudomonas aeruginosa in conditions more relevant to the cystic fibrosis lung. J. Vis. Exp. 2012, 64, e3857. [Google Scholar] [CrossRef]
  30. Quinn, R.A.; Whiteson, K.; Lim, Y.W.; Salamon, P.; Bailey, B.; Mienardi, S.; Sanchez, S.E.; Blake, D.; Conrad, D.; Rohwer, F. A Winogradsky-based culture system shows an association between microbial fermentation and cystic fibrosis exacerbation. ISME 2015, 9, 1024–1038. [Google Scholar] [CrossRef]
  31. Sriramulu, D.D.; Lunsdorf, H.; Lam, J.S.; Romling, U. Microcolony formation: A novel biofilm model of Pseudomonas aeruginosa for the cystic fibrosis lung. J. Med. Microbiol. 2005, 54, 667–676. [Google Scholar] [CrossRef] [PubMed]
  32. Serisier, D.J.; Carroll, M.P.; Shute, J.K.; Young, S.A. Macrorheology of cystic fibrosis, chronic obstructive pulmonary disease & normal sputum. Respir. Res. 2009, 10, 63. [Google Scholar] [CrossRef] [PubMed]
  33. Fahy, J.V.; Kim, K.W.; Liu, J.; Boushey, H.A. Prominent neutrophilic inflammation in sputum from subjects with asthma exacerbation. J. Allergy. Clin. Immunol. 1995, 95, 843–852. [Google Scholar] [CrossRef] [PubMed]
  34. Imam, J.S.; Duarte, A.G. Non-CF bronchiectasis: Orphan disease no longer. Respir. Med. 2020, 166, 105940. [Google Scholar] [CrossRef] [PubMed]
  35. Eickmeier, O.; Huebner, M.; Herrmann, E.; Zissler, U.; Rosewich, M.; Baer, P.C.; Buhl, R.; Schmitt-Grohé, S.; Zielen, S.; Schubert, R. Sputum biomarker profiles in cystic fibrosis (CF) and chronic obstructive pulmonary disease (COPD) and association between pulmonary function. Cytokine 2010, 50, 152–157. [Google Scholar] [CrossRef] [PubMed]
  36. Rubin, B.K. Mucus, phlegm, and sputum in cystic fibrosis. Respir. Care 2009, 54, 726–732, discussion 732. [Google Scholar] [CrossRef] [PubMed]
  37. Koenig, S.M.; Truwit, J.D. Ventilator-associated pneumonia: Diagnosis, treatment, and prevention. Clin. Microbiol. Rev. 2006, 19, 637–657. [Google Scholar] [CrossRef] [PubMed]
  38. Charles, P.M.V.; Kali, A.; Easow, J.M.; Joseph, N.M.; Ravishankar, M.; Srinivasan, S.; Kumar, S.; Umadevi, S. Ventilator-associated pneumonia. Australas. Med. J. 2014, 7, 334–344. [Google Scholar] [CrossRef] [PubMed]
  39. Mehta, A.; Bhagat, R. Preventing Ventilator-Associated Infections. Clin. Chest. Med. 2016, 37, 683–692. [Google Scholar] [CrossRef] [PubMed]
  40. Nora, D.; Povoa, P. Antibiotic consumption and ventilator-associated pneumonia rates, some parallelism but some discrepancies. Ann. Transl. Med. 2017, 5, 450. [Google Scholar] [CrossRef]
  41. Camara, M.; Green, W.; MacPhee, C.E.; Rakowska, P.D.; Raval, R.; Richardson, M.C.; Slater-Jefferies, J.; Steventon, K.; Webb, J.S. Economic significance of biofilms: A multidisciplinary and cross-sectoral challenge. NPJ Biofilms Microbiomes 2022, 8, 42. [Google Scholar] [CrossRef] [PubMed]
  42. Gragueb-Chatti, I.; Hyvernat, H.; Leone, M.; Agard, G.; Peres, N.; Guervilly, C.; Boucekine, M.; Hamidi, D.; Papazian, L.; Dellamonica, J.; et al. Incidence, Outcomes and Risk Factors of Recurrent Ventilator Associated Pneumonia in COVID-19 Patients: A Retrospective Multicenter Study. J. Clin. Med. 2022, 11, 7097. [Google Scholar] [CrossRef] [PubMed]
  43. Alves, D.; Grainha, T.; Pereira, M.O.; Lopes, S.P. Antimicrobial materials for endotracheal tubes: A review on the last two decades of technological progress. Acta Biomater. 2023, 158, 32–55. [Google Scholar] [CrossRef] [PubMed]
  44. Papazian, L.; Klompas, M.; Luyt, C.E. Ventilator-associated pneumonia in adults: A narrative review. Intensive Care Med. 2020, 46, 888–906. [Google Scholar] [CrossRef] [PubMed]
  45. Yamada, K.; Yamamoto, Y.; Yanagihara, K.; Araki, N.; Harada, Y.; Morinaga, Y.; Izumikawa, K.; Kakeya, H.; Hasegawa, H.; Kohno, S.; et al. In vivo efficacy and pharmacokinetics of biapenem in a murine model of ventilator-associated pneumonia with Pseudomonas aeruginosa. J. Infect. Chemother. 2012, 18, 472–478. [Google Scholar] [CrossRef] [PubMed]
  46. Luna, C.M.; Sibila, O.; Agusti, C.; Torres, A. Animal models of ventilator-associated pneumonia. Eur. Respir. J. 2009, 33, 182–188. [Google Scholar] [CrossRef] [PubMed]
  47. Benahmed, M.A.; Elbayed, K.; Daubeuf, F.; Santelmo, N.; Frossard, N.; Namer, I.J. NMR HRMAS spectroscopy of lung biopsy samples: Comparison study between human, pig, rat, and mouse metabolomics. Magn. Reson. Med. 2014, 71, 35–43. [Google Scholar] [CrossRef] [PubMed]
  48. Kozlova, I.; Vanthanouvong, V.; Almgren, B.; Hogman, M.; Roomans, G.M. Elemental composition of airway surface liquid in the pig determined by x-ray microanalysis. Am. J. Respir. Cell Mol. Biol. 2005, 32, 59–64. [Google Scholar] [CrossRef] [PubMed]
  49. Cowley, E.A.; Govindaraju, K.; Guilbault, C.; Radzioch, D.; Eidelman, D.H. Airway surface liquid composition in mice. Am. J. Physiol. Lung. Cell. Mol. Physiol. 2000, 278, L1213–L1220. [Google Scholar] [CrossRef]
  50. Robinson, N.P.; Kyle, H.; Webber, S.E.; Widdicombe, J.G. Electrolyte and other chemical concentrations in tracheal airway surface liquid and mucus. J. Appl. Physiol. 1989, 66, 2129–2135. [Google Scholar] [CrossRef]
  51. Boucher, R.C.; Stutts, M.J.; Bromberg, P.A.; Gatzy, J.T. Regional differences in airway surface liquid composition. J. Appl. Physiol. Respir. Environ. Exerc. Physiol. 1981, 50, 613–620. [Google Scholar] [CrossRef] [PubMed]
  52. Song, Y.; Thiagarajah, J.; Verkman, A.S. Sodium and chloride concentrations, pH, and depth of airway surface liquid in distal airways. J. Gen. Physiol. 2003, 122, 511–519. [Google Scholar] [CrossRef] [PubMed]
  53. Rose, M.C.; Voynow, J.A. Respiratory tract mucin genes and mucin glycoproteins in health and disease. Physiol. Rev. 2006, 86, 245–278. [Google Scholar] [CrossRef] [PubMed]
  54. Boat, T.F.; Cheng, P.W.; Iyer, R.N.; Carlson, D.M.; Polony, I. Human respiratory tract secretion: Mucous glycoproteins of nonpurulent tracheobronchial secretions, and sputum of patients with bronchitis and cystic fibrosis. Arch. Biochem. Biophys. 1977, 177, 97–104. [Google Scholar]
  55. Henke, M.O.; John, G.; Germann, M.; Lindemann, H.; Rubin, B.K. MUC5AC and MUC5B mucins increase in cystic fibrosis airway secretions during pulmonary exacerbation. Am. J. Respir. Crit. Care Med. 2007, 175, 816–821. [Google Scholar] [CrossRef] [PubMed]
  56. Kirkham, S.; Sheehan, J.K.; Knight, D.; Richardson, P.S.; Thornton, D.J. Heterogeneity of airways mucus: Variations in the amounts and glycoforms of the major oligomeric mucins MUC5AC and MUC5B. Biochem. J. 2002, 361, 537–546. [Google Scholar] [CrossRef] [PubMed]
  57. Wickstrom, C.; Davies, J.R.; Eriksen, G.V.; Veerman, E.C.; Carlstedt, I. MUC5B is a major gel-forming, oligomeric mucin from human salivary gland, respiratory tract and endocervix: Identification of glycoforms and C-terminal cleavage. Biochem. J. 1998, 334, 685–693. [Google Scholar] [CrossRef] [PubMed]
  58. Xia, B.; Royall, J.A.; Damera, G.; Sachdev, G.P.; Cummings, R.D. Altered O-glycosylation and sulfation of airway mucins associated with cystic fibrosis. Glycobiology 2005, 15, 747–775. [Google Scholar] [CrossRef] [PubMed]
  59. Patarin, J.; Ghiringhelli, E.; Darsy, G.; Obamba, M.; Bochu, P.; Camara, B.; Quetant, S.; Cracowski, J.L.; Cracowski, C.; Robert de Saint Vincent, M. Rheological analysis of sputum from patients with chronic bronchial diseases. Sci. Rep. 2020, 10, 15685. [Google Scholar] [CrossRef]
  60. Henke, M.O.; Renner, A.; Huber, R.M.; Seeds, M.C.; Rubin, B.K. MUC5AC and MUC5B Mucins Are Decreased in Cystic Fibrosis Airway Secretions. Am. J. Respir. Cell Mol. Biol. 2004, 31, 86–91. [Google Scholar] [CrossRef]
  61. Henderson, A.G.; Ehre, C.; Button, B.; Abdullah, L.H.; Cai, L.H.; Leigh, M.W.; DeMaria, G.C.; Matsui, H.; Donaldson, S.H.; Davis, C.W.; et al. Cystic fibrosis airway secretions exhibit mucin hyperconcentration and increased osmotic pressure. J. Clin. Investig. 2014, 124, 3047–3060. [Google Scholar] [CrossRef] [PubMed]
  62. Hellyer, T.P.; Morris, A.C.; McAuley, D.F.; Walsh, T.S.; Anderson, N.H.; Singh, S.; Dark, P.; Roy, A.I.; Baudouin, S.V.; Wright, S.E.; et al. Diagnostic accuracy of pulmonary host inflammatory mediators in the exclusion of ventilator-acquired pneumonia. Thorax 2015, 70, 41–47. [Google Scholar] [CrossRef] [PubMed]
  63. Gernez, Y.; Tirouvanziam, R.; Chanez, P. Neutrophils in chronic inflammatory airway diseases: Can we target them and how? Eur. Respir. J. 2010, 35, 467–469. [Google Scholar] [CrossRef] [PubMed]
  64. Kirkham, S.; Kolsum, U.; Rousseau, K.; Singh, D.; Vestbo, J.; Thornton, D.J. MUC5B Is the Major Mucin in the Gel Phase of Sputum in Chronic Obstructive Pulmonary Disease. Am. J. Respir. Crit. Care Med. 2008, 178, 1033–1039. [Google Scholar] [CrossRef] [PubMed]
  65. Bassi, G.L.; Zanella, A.; Cressoni, M.; Stylianou, M.; Kolobow, T. Following tracheal intubation, mucus flow is reversed in the semirecumbent position: Possible role in the pathogenesis of ventilator-associated pneumonia. Crit. Care Med. 2008, 36, 518–525. [Google Scholar] [CrossRef] [PubMed]
  66. Hoffman, C.L.; Lalsiamthara, J.; Aballay, A. Host Mucin Is Exploited by Pseudomonas aeruginosa to Provide Monosaccharides Required for a Successful Infection. mBio 2020, 11, e00060-20. [Google Scholar] [CrossRef] [PubMed]
  67. Yesilkaya, H.; Manco, S.; Kadioglu, A.; Terra, V.S.; Andrew, P.W. The ability to utilize mucin affects the regulation of virulence gene expression in Streptococcus pneumoniae. FEMS Microbiol. Lett. 2008, 278, 231–235. [Google Scholar] [CrossRef] [PubMed]
  68. Venkatakrishnan, V.; Packer, N.H.; Thaysen-Andersen, M. Host mucin glycosylation plays a role in bacterial adhesion in lungs of individuals with cystic fibrosis. Expert Rev. Respir. Med. 2013, 7, 553–576. [Google Scholar] [CrossRef] [PubMed]
  69. Wheeler, K.M.; Carcamo-Oyarce, G.; Turner, B.S.; Dellos-Nolan, S.; Co, J.Y.; Lehoux, S.; Cummings, R.D.; Wozniak, D.J.; Ribbeck, K. Mucin glycans attenuate the virulence of Pseudomonas aeruginosa in infection. Nat. Microbiol. 2019, 4, 2146–2154. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, B.X.; Wheeler, K.M.; Cady, K.C.; Lehoux, S.; Cummings, R.D.; Laub, M.T.; Ribbeck, K. Mucin Glycans Signal through the Sensor Kinase RetS to Inhibit Virulence-Associated Traits in Pseudomonas aeruginosa. Curr. Biol. 2021, 31, 90–102. [Google Scholar] [CrossRef] [PubMed]
  71. Wagner, C.E.; Wheeler, K.M.; Ribbeck, K. Mucins and Their Role in Shaping the Functions of Mucus Barriers. Annu. Rev. Cell Dev. Biol. 2018, 34, 189–215. [Google Scholar] [CrossRef] [PubMed]
  72. Schulz, B.L.; Sloane, A.J.; Robinson, L.J.; Prasad, S.S.; Lindner, R.A.; Robinson, M.; Bye, P.T.; Nielson, D.W.; Harry, J.L.; Packer, N.H.; et al. Glycosylation of sputum mucins is altered in cystic fibrosis patients. Glycobiology 2007, 17, 698–712. [Google Scholar] [CrossRef] [PubMed]
  73. Lai, S.K.; Wang, Y.Y.; Wirtz, D.; Hanes, J. Micro- and macrorheology of mucus. Adv. Drug Deliv. Rev. 2009, 61, 86–100. [Google Scholar] [CrossRef] [PubMed]
  74. Flynn, J.M.; Niccum, D.; Dunitz, J.M.; Hunter, R.C. Evidence and Role for Bacterial Mucin Degradation in Cystic Fibrosis Airway Disease. PLoS Pathog. 2016, 12, e1005846. [Google Scholar] [CrossRef] [PubMed]
  75. Singh, P.K.; Schaefer, A.L.; Parsek, M.R.; Moninger, T.O.; Welsh, M.J.; Greenberg, E.P. Quorum-sensing signals indicate that cystic fibrosis lungs are infected with bacterial biofilms. Nature 2000, 407, 762–764. [Google Scholar] [CrossRef] [PubMed]
  76. Zanin, M.; Baviskar, P.; Webster, R.; Webby, R. The Interaction between Respiratory Pathogens and Mucus. Cell Host Microbe 2016, 19, 159–168. [Google Scholar] [CrossRef] [PubMed]
  77. Paharik, A.E.; Salgado-Pabon, W.; Meyerholz, D.K.; White, M.J.; Schlievert, P.M.; Horswill, A.R. The Spl Serine Proteases Modulate Staphylococcus aureus Protein Production and Virulence in a Rabbit Model of Pneumonia. mSphere 2016, 1, e00208-16. [Google Scholar] [CrossRef] [PubMed]
  78. Takagi, J.; Aoki, K.; Turner, B.S.; Lamont, S.; Lehoux, S.; Kavanaugh, N.; Gulati, M.; Valle Arevalo, A.; Lawrence, T.J.; Kim, C.Y.; et al. Mucin O-glycans are natural inhibitors of Candida albicans pathogenicity. Nat. Chem. Biol. 2022, 18, 762–773. [Google Scholar] [CrossRef] [PubMed]
  79. Weiser, J.N.; Ferreira, D.M.; Paton, J.C. Streptococcus pneumoniae: Transmission, colonization and invasion. Nat. Rev. Microbiol. 2018, 16, 355–367. [Google Scholar] [CrossRef] [PubMed]
  80. Ahearn, C.P.; Gallo, M.C.; Murphy, T.F. Insights on persistent airway infection by non-typeable Haemophilus influenzae in chronic obstructive pulmonary disease. Pathog. Dis. 2017, 75, ftx042. [Google Scholar] [CrossRef]
  81. Turner, K.H.; Wessel, A.K.; Palmer, G.C.; Murray, J.L.; Whiteley, M. Essential genome of Pseudomonas aeruginosa in cystic fibrosis sputum. Proc. Natl. Acad. Sci. USA 2015, 112, 4110–4115. [Google Scholar] [CrossRef] [PubMed]
  82. Padra, M.; Adamczyk, B.; Benktander, J.; Flahou, B.; Skoog, E.C.; Padra, J.T.; Smet, A.; Jin, C.; Ducatelle, R.; Samuelsson, T.; et al. Helicobacter suis binding to carbohydrates on human and porcine gastric mucins and glycolipids occurs via two modes. Virulence 2018, 9, 898–918. [Google Scholar] [CrossRef] [PubMed]
  83. Barmpatsalou, V.; Dubbelboer, I.R.; Rodler, A.; Jacobson, M.; Karlsson, E.; Pedersen, B.L.; Bergstrom, C.A.S. Physiological properties, composition and structural profiling of porcine gastrointestinal mucus. Eur. J. Pharm. Biopharm. 2021, 169, 156–167. [Google Scholar] [CrossRef] [PubMed]
  84. Rousseau, K.; Kirkham, S.; Johnson, L.; Fitzpatrick, B.; Howard, M.; Adams, E.J.; Rogers, D.F.; Knight, D.; Clegg, P.; Thornton, D.J. Proteomic analysis of polymeric salivary mucins: No evidence for MUC19 in human saliva. Biochem. J. 2008, 413, 545–552. [Google Scholar] [CrossRef] [PubMed]
  85. Lieleg, O.; Lieleg, C.; Bloom, J.; Buck, C.B.; Ribbeck, K. Mucin biopolymers as broad-spectrum antiviral agents. Biomacromolecules 2012, 13, 1724–1732. [Google Scholar] [CrossRef] [PubMed]
  86. Harris, G.; Holbein, B.E.; Zhou, H.; Xu, H.H.; Chen, W. Potential Mechanisms of Mucin-Enhanced Acinetobacter baumannii Virulence in the Mouse Model of Intraperitoneal Infection. Infect. Immun. 2019, 87, e00591-19. [Google Scholar] [CrossRef] [PubMed]
  87. Neve, R.L.; Carrillo, B.D.; Phelan, V.V. Impact of Artificial Sputum Medium Formulation on Pseudomonas aeruginosa Secondary Metabolite Production. J. Bacteriol. 2021, 203, e0025021. [Google Scholar] [CrossRef] [PubMed]
  88. Nisizawa, K.; Pigman, W. The composition and properties of the mucin clot from cattle submaxillary glands. Arch. Oral Biol. 1959, 1, 161–170. [Google Scholar] [CrossRef] [PubMed]
  89. Tettamanti, G.; Pigman, W. Purification and characterization of bovine and ovine submaxillary mucins. Arch. Biochem. Biophys. 1968, 124, 41–50. [Google Scholar] [CrossRef]
  90. Schömig, V.J.; Käsdorf, B.T.; Scholz, C.; Bidmon, K.; Lieleg, O.; Berensmeier, S. An optimized purification process for porcine gastric mucin with preservation of its native functional properties. RSC Adv. 2016, 6, 44932–44943. [Google Scholar] [CrossRef]
  91. Brinkmann, V.; Reichard, U.; Goosmann, C.; Fauler, B.; Uhlemann, Y.; Weiss, D.S.; Weinrauch, Y.; Zychlinsky, A. Neutrophil extracellular traps kill bacteria. Science 2004, 303, 1532–1535. [Google Scholar] [CrossRef] [PubMed]
  92. Mikacenic, C.; Moore, R.; Dmyterko, V.; West, T.E.; Altemeier, W.A.; Liles, W.C.; Lood, C. Neutrophil extracellular traps (NETs) are increased in the alveolar spaces of patients with ventilator-associated pneumonia. Crit. Care 2018, 22, 358. [Google Scholar] [CrossRef] [PubMed]
  93. Radermecker, C.; Detrembleur, N.; Guiot, J.; Cavalier, E.; Henket, M.; d’Emal, C.; Vanwinge, C.; Cataldo, D.; Oury, C.; Delvenne, P.; et al. Neutrophil extracellular traps infiltrate the lung airway, interstitial, and vascular compartments in severe COVID-19. J. Exp. Med. 2020, 217, e20201012. [Google Scholar] [CrossRef] [PubMed]
  94. Barnes, B.J.; Adrover, J.M.; Baxter-Stoltzfus, A.; Borczuk, A.; Cools-Lartigue, J.; Crawford, J.M.; Dassler-Plenker, J.; Guerci, P.; Huynh, C.; Knight, J.S.; et al. Targeting potential drivers of COVID-19: Neutrophil extracellular traps. J. Exp. Med. 2020, 217, e20200652. [Google Scholar] [CrossRef] [PubMed]
  95. Porto, B.N.; Stein, R.T. Neutrophil Extracellular Traps in Pulmonary Diseases: Too Much of a Good Thing? Front. Immunol. 2016, 7, 311. [Google Scholar] [CrossRef] [PubMed]
  96. Yildiz, C.; Palaniyar, N.; Otulakowski, G.; Khan, M.A.; Post, M.; Kuebler, W.M.; Tanswell, K.; Belcastro, R.; Masood, A.; Engelberts, D.; et al. Mechanical ventilation induces neutrophil extracellular trap formation. Anesthesiology 2015, 122, 864–875. [Google Scholar] [CrossRef] [PubMed]
  97. De Rose, V. Mechanisms and markers of airway inflammation in cystic fibrosis. Eur. Respir. J. 2002, 19, 333–340. [Google Scholar] [CrossRef] [PubMed]
  98. Rosenecker, J.; Naundorf, S.; Rudolph, C. Airway surface liquid contains endogenous DNase activity which can be activated by exogenous magnesium. Eur. J. Med. Res. 2009, 14, 304–308. [Google Scholar] [CrossRef] [PubMed]
  99. Lethem, M.I.; James, S.L.; Marriott, C.; Burke, J.F. The origin of DNA associated with mucus glycoproteins in cystic fibrosis sputum. Eur. Respir. J. 1990, 3, 19–23. [Google Scholar] [CrossRef]
  100. Lewenza, S.; Johnson, L.; Charron-Mazenod, L.; Hong, M.; Mulcahy-O’Grady, H. Extracellular DNA controls expression of Pseudomonas aeruginosa genes involved in nutrient utilization, metal homeostasis, acid pH tolerance and virulence. J. Med. Microbiol. 2020, 69, 895–905. [Google Scholar] [CrossRef]
  101. Wilton, M.; Charron-Mazenod, L.; Moore, R.; Lewenza, S. Extracellular DNA Acidifies Biofilms and Induces Aminoglycoside Resistance in Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 2016, 60, 544–553. [Google Scholar] [CrossRef] [PubMed]
  102. Herzog, S.; Dach, F.; de Buhr, N.; Niemann, S.; Schlagowski, J.; Chaves-Moreno, D.; Neumann, C.; Goretzko, J.; Schwierzeck, V.; Mellmann, A.; et al. High Nuclease Activity of Long Persisting Staphylococcus aureus Isolates Within the Airways of Cystic Fibrosis Patients Protects Against NET-Mediated Killing. Front. Immunol. 2019, 10, 2552. [Google Scholar] [CrossRef] [PubMed]
  103. Brandt, T.; Breitenstein, S.; von der Hardt, H.; Tummler, B. DNA concentration and length in sputum of patients with cystic fibrosis during inhalation with recombinant human DNase. Thorax 1995, 50, 880–882. [Google Scholar] [CrossRef] [PubMed]
  104. Kirchner, K.K.; Wagener, J.S.; Khan, T.Z.; Copenhaver, S.C.; Accurso, F.J. Increased DNA levels in bronchoalveolar lavage fluid obtained from infants with cystic fibrosis. Am. J. Respir. Crit. Care Med. 1996, 154, 1426–1429. [Google Scholar] [CrossRef] [PubMed]
  105. Marcos, V.; Zhou-Suckow, Z.; Onder Yildirim, A.; Bohla, A.; Hector, A.; Vitkov, L.; Krautgartner, W.D.; Stoiber, W.; Griese, M.; Eickelberg, O.; et al. Free DNA in cystic fibrosis airway fluids correlates with airflow obstruction. Mediat. Inflamm. 2015, 2015, 408935. [Google Scholar] [CrossRef] [PubMed]
  106. McShane, D.; Davies, J.C.; Davies, M.G.; Bush, A.; Geddes, D.M.; Alton, E.W. Airway surface pH in subjects with cystic fibrosis. Eur. Respir. J. 2003, 21, 37–42. [Google Scholar] [CrossRef] [PubMed]
  107. Jayaraman, S.; Song, Y.; Vetrivel, L.; Shankar, L.; Verkman, A.S. Noninvasive in vivo fluorescence measurement of airway-surface liquid depth, salt concentration, and pH. J. Clin. Investig. 2001, 107, 317–324. [Google Scholar] [CrossRef] [PubMed]
  108. Song, Y.; Salinas, D.; Nielson, D.W.; Verkman, A.S. Hyperacidity of secreted fluid from submucosal glands in early cystic fibrosis. Am. J. Physiol.-Cell Physiol. 2006, 290, C741–C749. [Google Scholar] [CrossRef] [PubMed]
  109. Jayaraman, S.; Joo, N.S.; Reitz, B.; Wine, J.J.; Verkman, A.S. Submucosal gland secretions in airways from cystic fibrosis patients have normal [Na+] and pH but elevated viscosity. Proc. Natl. Acad. Sci. USA 2001, 98, 8119–8123. [Google Scholar] [CrossRef]
  110. Bodem, C.R.; Lampton, L.M.; Miller, D.P.; Tarka, E.F.; Everett, E.D. Endobronchial pH. Relevance of aminoglycoside activity in gram-negative bacillary pneumonia. Am. Rev. Respir. Dis. 1983, 127, 39–41. [Google Scholar] [CrossRef]
  111. Karnad, D.R.; Mhaisekar, D.G.; Moralwar, K.V. Respiratory mucus pH in tracheostomized intensive care unit patients: Effects of colonization and pneumonia. Crit. Care Med. 1990, 18, 699–701. [Google Scholar] [CrossRef] [PubMed]
  112. Lozo Vukovac, E.; Mise, K.; Gudelj, I.; Peric, I.; Duplancic, D.; Vukovic, I.; Vucinovic, Z.; Lozo, M. Bronchoalveolar pH and inflammatory biomarkers in patients with acute exacerbation of chronic obstructive pulmonary disease. J. Int. Med. Res. 2019, 47, 791–802. [Google Scholar] [CrossRef] [PubMed]
  113. Baker, E.H.; Clark, N.; Brennan, A.L.; Fisher, D.A.; Gyi, K.M.; Hodson, M.E.; Philips, B.J.; Baines, D.L.; Wood, D.M. Hyperglycemia and cystic fibrosis alter respiratory fluid glucose concentrations estimated by breath condensate analysis. J. Appl. Physiol. 2007, 102, 1969–1975. [Google Scholar] [CrossRef] [PubMed]
  114. Philips, B.J.; Meguer, J.X.; Redman, J.; Baker, E.H. Factors determining the appearance of glucose in upper and lower respiratory tract secretions. Intensive Care Med. 2003, 29, 2204–2210. [Google Scholar] [CrossRef] [PubMed]
  115. Philips, B.J.; Redman, J.; Brennan, A.; Wood, D.; Holliman, R.; Baines, D.; Baker, E.H. Glucose in bronchial aspirates increases the risk of respiratory MRSA in intubated patients. Thorax 2005, 60, 761–764. [Google Scholar] [CrossRef] [PubMed]
  116. Barth, A.L.; Pitt, T.L. The high amino-acid content of sputum from cystic fibrosis patients promotes growth of auxotrophic Pseudomonas aeruginosa. J. Med. Microbiol. 1996, 45, 110–119. [Google Scholar] [CrossRef] [PubMed]
  117. Thomas, S.R.; Ray, A.; Hodson, M.E.; Pitt, T.L. Increased sputum amino acid concentrations and auxotrophy of Pseudomonas aeruginosa in severe cystic fibrosis lung disease. Thorax 2000, 55, 795–797. [Google Scholar] [CrossRef] [PubMed]
  118. Sagel, S.D.; Sontag, M.K.; Accurso, F.J. Relationship Between Antimicrobial Proteins and AirwayInflammation and Infection in Cystic Fibrosis. Pediatr. Pulmonol. 2009, 44, 402–409. [Google Scholar] [CrossRef] [PubMed]
  119. Reid, D.W.; Lam, Q.T.; Schneider, H.; Walters, E.H. Airway iron and iron-regulatory cytokines in cystic fibrosis. Eur. Respir. J. 2004, 24, 286–291. [Google Scholar] [CrossRef] [PubMed]
  120. Grasemann, H.; Shehnaz, D.; Enomoto, M.; Leadley, M.; Belik, J.; Ratjen, F. L-ornithine derived polyamines in cystic fibrosis airways. PLoS ONE 2012, 7, e46618. [Google Scholar] [CrossRef]
  121. Sagel, S.D.; Kapsner, R.; Osberg, I.; Sontag, M.K.; Accurso, F.J. Airway inflammation in children with cystic fibrosis and healthy children assessed by sputum induction. Am. J. Respir. Crit. Care Med. 2001, 164, 1425–1431. [Google Scholar] [CrossRef] [PubMed]
  122. Olivença, D.V.; Fonseca, L.L.; Voit, E.O.; Pinto, F.R. Thickness of the airway surface liquid layer in the lung is affected in cystic fibrosis by compromised synergistic regulation of the ENaC ion channel. J. R. Soc. Interface 2019, 16, 20190187. [Google Scholar] [CrossRef] [PubMed]
  123. Berkebile, A.R.; McCray, P.B., Jr. Effects of airway surface liquid pH on host defense in cystic fibrosis. Int. J. Biochem. Cell Biol. 2014, 52, 124–129. [Google Scholar] [CrossRef] [PubMed]
  124. Yoon, S.S.; Coakley, R.; Lau, G.W.; Lymar, S.V.; Gaston, B.; Karabulut, A.C.; Hennigan, R.F.; Hwang, S.H.; Buettner, G.; Schurr, M.J.; et al. Anaerobic killing of mucoid Pseudomonas aeruginosa by acidified nitrite derivatives under cystic fibrosis airway conditions. J. Clin. Investig. 2006, 116, 436–446. [Google Scholar] [CrossRef] [PubMed]
  125. Clary-Meinesz, C.; Mouroux, J.; Cosson, J.; Huitorel, P.; Blaive, B. Influence of external pH on ciliary beat frequency in human bronchi and bronchioles. Eur. Respir. J. 1998, 11, 330–333. [Google Scholar] [CrossRef] [PubMed]
  126. Simonin, J.; Bille, E.; Crambert, G.; Noel, S.; Dreano, E.; Edwards, A.; Hatton, A.; Pranke, I.; Villeret, B.; Cottart, C.H.; et al. Airway surface liquid acidification initiates host defense abnormalities in Cystic Fibrosis. Sci. Rep. 2019, 9, 6516. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, Y.; Xie, Y.Z.; Shi, Y.H.; Yang, L.; Chen, X.Y.; Wang, L.W.; Qu, J.M.; Weng, D.; Wang, X.J.; Liu, H.P.; et al. Airway acidification impaired host defense against Pseudomonas aeruginosa infection by promoting type 1 interferon beta response. Emerg. Microbes Infect. 2022, 11, 2132–2146. [Google Scholar] [CrossRef] [PubMed]
  128. Palmieri, F.; Koutsokera, A.; Bernasconi, E.; Junier, P.; von Garnier, C.; Ubags, N. Recent Advances in Fungal Infections: From Lung Ecology to Therapeutic Strategies With a Focus on Aspergillus spp. Front. Med. 2022, 9, 832510. [Google Scholar] [CrossRef] [PubMed]
  129. Joris, L.; Dab, I.; Quinton, P.M. Elemental composition of human airway surface fluid in healthy and diseased airways. Am. Rev. Respir. Dis. 1993, 148, 1633–1637. [Google Scholar] [CrossRef]
  130. Goldman, M.J.; Anderson, G.M.; Stolzenberg, E.D.; Kari, U.P.; Zasloff, M.; Wilson, J.M. Human beta-defensin-1 is a salt-sensitive antibiotic in lung that is inactivated in cystic fibrosis. Cell 1997, 88, 553–560. [Google Scholar] [CrossRef]
  131. Garnett, J.P.; Nguyen, T.T.; Moffatt, J.D.; Pelham, E.R.; Kalsi, K.K.; Baker, E.H.; Baines, D.L. Pro-inflammatory mediators disrupt glucose homeostasis in airway surface liquid. J. Immunol. 2012, 189, 373–380. [Google Scholar] [CrossRef] [PubMed]
  132. Baker, E.H.; Janaway, C.H.; Philips, B.J.; Brennan, A.L.; Baines, D.L.; Wood, D.M.; Jones, P.W. Hyperglycaemia is associated with poor outcomes in patients admitted to hospital with acute exacerbations of chronic obstructive pulmonary disease. Thorax 2006, 61, 284–289. [Google Scholar] [CrossRef] [PubMed]
  133. Mirabella, S.; Gomez-Paz, S.; Lam, E.; Gonzalez-Mosquera, L.; Fogel, J.; Rubinstein, S. Glucose dysregulation and its association with COVID-19 mortality and hospital length of stay. Diabetes Metab. Syndr. 2022, 16, 102439. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, D.; Hu, B.; Hu, C.; Zhu, F.; Liu, X.; Zhang, J.; Wang, B.; Xiang, H.; Cheng, Z.; Xiong, Y.; et al. Clinical Characteristics of 138 Hospitalized Patients With 2019 Novel Coronavirus-Infected Pneumonia in Wuhan, China. JAMA 2020, 323, 1061–1069. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, W.; Shen, M.; Tao, Y.; Fairley, C.K.; Zhong, Q.; Li, Z.; Chen, H.; Ong, J.J.; Zhang, D.; Zhang, K.; et al. Elevated glucose level leads to rapid COVID-19 progression and high fatality. BMC Pulm. Med. 2021, 21, 64. [Google Scholar] [CrossRef] [PubMed]
  136. Wu, C.; Chen, X.; Cai, Y.; Xia, J.; Zhou, X.; Xu, S.; Huang, H.; Zhang, L.; Zhou, X.; Du, C.; et al. Risk Factors Associated With Acute Respiratory Distress Syndrome and Death in Patients With Coronavirus Disease 2019 Pneumonia in Wuhan, China. JAMA Intern. Med. 2020, 180, 934–943. [Google Scholar] [CrossRef] [PubMed]
  137. Baker, E.H.; Baines, D.L. Airway Glucose Homeostasis: A New Target in the Prevention and Treatment of Pulmonary Infection. Chest 2018, 153, 507–514. [Google Scholar] [CrossRef] [PubMed]
  138. Gill, S.K.; Hui, K.; Farne, H.; Garnett, J.P.; Baines, D.L.; Moore, L.S.; Holmes, A.H.; Filloux, A.; Tregoning, J.S. Increased airway glucose increases airway bacterial load in hyperglycaemia. Sci. Rep. 2016, 6, 27636. [Google Scholar] [CrossRef] [PubMed]
  139. McAlister, F.A.; Majumdar, S.R.; Blitz, S.; Rowe, B.H.; Romney, J.; Marrie, T.J. The relation between hyperglycemia and outcomes in 2471 patients admitted to the hospital with community-acquired pneumonia. Diabetes Care 2005, 28, 810–815. [Google Scholar] [CrossRef]
  140. Brennan, A.L.; Gyi, K.M.; Wood, D.M.; Johnson, J.; Holliman, R.; Baines, D.L.; Philips, B.J.; Geddes, D.M.; Hodson, M.E.; Baker, E.H. Airway glucose concentrations and effect on growth of respiratory pathogens in cystic fibrosis. J. Cyst. Fibros. 2007, 6, 101–109. [Google Scholar] [CrossRef]
  141. Garnett, J.P.; Braun, D.; McCarthy, A.J.; Farrant, M.R.; Baker, E.H.; Lindsay, J.A.; Baines, D.L. Fructose transport-deficient Staphylococcus aureus reveals important role of epithelial glucose transporters in limiting sugar-driven bacterial growth in airway surface liquid. Cell. Mol. Life. Sci. 2014, 71, 4665–4673. [Google Scholar] [CrossRef] [PubMed]
  142. Dennesen, P.; Veerman, E.; van Nieuw Amerongen, A.; Jacobs, J.; Kessels, A.; van der Keybus, P.; Ramsay, G.; van der Ven, A. High levels of sulfated mucins in bronchoalveolar lavage fluid of ICU patients with ventilator-associated pneumonia. Intensive Care Med. 2003, 29, 715–719. [Google Scholar] [CrossRef] [PubMed]
  143. Davril, M.; Degroote, S.; Humbert, P.; Galabert, C.; Dumur, V.; Lafitte, J.; Lamblin, G.; Roussel, P. The sialylation of bronchial mucins secreted by patients suffering from cystic fibrosis or from chronic bronchitis is related to the severity of airway infection. Glycobiology 1999, 9, 311–321. [Google Scholar] [CrossRef] [PubMed]
  144. Aprianto, R.; Slager, J.; Holsappel, S.; Veening, J.W. High-resolution analysis of the pneumococcal transcriptome under a wide range of infection-relevant conditions. Nucleic Acids Res. 2018, 46, 9990–10006. [Google Scholar] [CrossRef] [PubMed]
  145. Blanchette, K.A.; Shenoy, A.T.; Milner, J., 2nd; Gilley, R.P.; McClure, E.; Hinojosa, C.A.; Kumar, N.; Daugherty, S.C.; Tallon, L.J.; Ott, S.; et al. Neuraminidase A-Exposed Galactose Promotes Streptococcus pneumoniae Biofilm Formation during Colonization. Infect. Immun. 2016, 84, 2922–2932. [Google Scholar] [CrossRef] [PubMed]
  146. Korgaonkar, A.K.; Whiteley, M. Pseudomonas aeruginosa enhances production of an antimicrobial in response to N-acetylglucosamine and peptidoglycan. J. Bacteriol. 2011, 193, 909–917. [Google Scholar] [CrossRef] [PubMed]
  147. Ding, X.; Robbe-Masselot, C.; Fu, X.; Leonard, R.; Marsac, B.; Dauriat, C.J.G.; Lepissier, A.; Rytter, H.; Ramond, E.; Dupuis, M.; et al. Airway environment drives the selection of quorum sensing mutants and promote Staphylococcus aureus chronic lifestyle. Nat. Commun. 2023, 14, 8135. [Google Scholar] [CrossRef] [PubMed]
  148. Lucas, S.K.; Villarreal, A.R.; Ahmad, M.M.; Itabiyi, A.; Feddema, E.; Boyer, H.C.; Hunter, R.C. Anaerobic Microbiota Derived from the Upper Airways Impact Staphylococcus aureus Physiology. Infect. Immun. 2021, 89, e0015321. [Google Scholar] [CrossRef] [PubMed]
  149. Mukherjee, K.; Khatua, B.; Mandal, C. Sialic Acid-Siglec-E Interactions During Pseudomonas aeruginosa Infection of Macrophages Interferes With Phagosome Maturation by Altering Intracellular Calcium Concentrations. Front. Immunol. 2020, 11, 332. [Google Scholar] [CrossRef]
  150. Khatua, B.; Bhattacharya, K.; Mandal, C. Sialoglycoproteins adsorbed by Pseudomonas aeruginosa facilitate their survival by impeding neutrophil extracellular trap through siglec-9. J. Leukoc. Biol. 2012, 91, 641–655. [Google Scholar] [CrossRef]
  151. Jackson, M.D.; Wong, S.M.; Akerley, B.J. Sialic Acid Protects Nontypeable Haemophilus influenzae from Natural IgM and Promotes Survival in Murine Respiratory Tract. Infect. Immun. 2021, 89, e00676-20. [Google Scholar] [CrossRef] [PubMed]
  152. Wong, S.M.; Jackson, M.D.; Akerley, B.J. Suppression of Alternative Lipooligosaccharide Glycosyltransferase Activity by UDP-Galactose Epimerase Enhances Murine Lung Infection and Evasion of Serum IgM. Front. Cell. Infect. Microbiol. 2019, 9, 160. [Google Scholar] [CrossRef] [PubMed]
  153. Rose, M.C.; Nickola, T.J.; Voynow, J.A. Airway mucus obstruction: Mucin glycoproteins, MUC gene regulation and goblet cell hyperplasia. Am. J. Respir. Cell Mol. Biol. 2001, 25, 533–537. [Google Scholar] [CrossRef] [PubMed]
  154. Green, A.E.; Pottenger, S.; Monshi, M.S.; Barton, T.E.; Phelan, M.; Neill, D.R. Airway metabolic profiling during Streptococcus pneumoniae infection identifies branched chain amino acids as signatures of upper airway colonisation. PLoS Pathog. 2023, 19, e1011630. [Google Scholar] [CrossRef] [PubMed]
  155. Rojo, F. Carbon catabolite repression in Pseudomonas: Optimizing metabolic versatility and interactions with the environment. FEMS Microbiol. Rev. 2010, 34, 658–684. [Google Scholar] [CrossRef] [PubMed]
  156. La Rosa, R.; Johansen, H.K.; Molin, S. Adapting to the Airways: Metabolic Requirements of Pseudomonas aeruginosa during the Infection of Cystic Fibrosis Patients. Metabolites 2019, 9, 234. [Google Scholar] [CrossRef] [PubMed]
  157. La Rosa, R.; Johansen, H.K.; Molin, S. Convergent Metabolic Specialization through Distinct Evolutionary Paths in Pseudomonas aeruginosa. mBio 2018, 9, e00269-18. [Google Scholar] [CrossRef] [PubMed]
  158. Ren, X.; Palmer, L.D. Acinetobacter Metabolism in Infection and Antimicrobial Resistance. Infect. Immun. 2023, 91, e0043322. [Google Scholar] [CrossRef] [PubMed]
  159. Hallman, M.; Merritt, T.A.; Akino, T.; Bry, K. Surfactant protein A, phosphatidylcholine, and surfactant inhibitors in epithelial lining fluid. Correlation with surface activity, severity of respiratory distress syndrome, and outcome in small premature infants. Am. Rev. Respir. Dis. 1991, 144, 1376–1384. [Google Scholar] [CrossRef] [PubMed]
  160. Robinson, M.J.; Krasnodembskaya, A.D. Therapeutic targeting of metabolic alterations in acute respiratory distress syndrome. Eur. Respir. Rev. 2020, 29, 200114. [Google Scholar] [CrossRef]
  161. Aiyer, A.; Manos, J. The Use of Artificial Sputum Media to Enhance Investigation and Subsequent Treatment of Cystic Fibrosis Bacterial Infections. Microorganisms 2022, 10, 1269. [Google Scholar] [CrossRef] [PubMed]
  162. Schwab, U.; Abdullah, L.H.; Perlmutt, O.S.; Albert, D.; Davis, C.W.; Arnold, R.R.; Yankaskas, J.R.; Gilligan, P.; Neubauer, H.; Randell, S.H.; et al. Localization of Burkholderia cepacia complex bacteria in cystic fibrosis lungs and interactions with Pseudomonas aeruginosa in hypoxic mucus. Infect. Immun. 2014, 82, 4729–4745. [Google Scholar] [CrossRef] [PubMed]
  163. Liaqat, A.; Mason, M.; Foster, B.J.; Kulkarni, S.; Barlas, A.; Farooq, A.M.; Patak, P.; Liaqat, H.; Basso, R.G.; Zaman, M.S.; et al. Evidence-Based Mechanical Ventilatory Strategies in ARDS. J. Clin. Med. 2022, 11, 319. [Google Scholar] [CrossRef] [PubMed]
  164. Farooqui, A.A. Lipid mediators in the neural cell nucleus: Their metabolism, signaling, and association with neurological disorders. Neuroscientist 2009, 15, 392–407. [Google Scholar] [CrossRef] [PubMed]
  165. Zehethofer, N.; Bermbach, S.; Hagner, S.; Garn, H.; Muller, J.; Goldmann, T.; Lindner, B.; Schwudke, D.; Konig, P. Lipid Analysis of Airway Epithelial Cells for Studying Respiratory Diseases. Chromatographia 2015, 78, 403–413. [Google Scholar] [CrossRef] [PubMed]
  166. Simons, K.; Toomre, D. Lipid rafts and signal transduction. Nat. Rev. Mol. Cell Biol. 2000, 1, 31–39. [Google Scholar] [CrossRef] [PubMed]
  167. Sahu, S.; Lynn, W.S. Lipid composition of sputum from patients with asthma and patients with cystic fibrosis. Inflammation 1978, 3, 27–36. [Google Scholar] [CrossRef] [PubMed]
  168. Sohlenkamp, C.; Lopez-Lara, I.M.; Geiger, O. Biosynthesis of phosphatidylcholine in bacteria. Prog. Lipid Res. 2003, 42, 115–162. [Google Scholar] [CrossRef] [PubMed]
  169. Cundell, D.R.; Gerard, N.P.; Gerard, C.; Idanpaan-Heikkila, I.; Tuomanen, E.I. Streptococcus pneumoniae anchor to activated human cells by the receptor for platelet-activating factor. Nature 1995, 377, 435–438. [Google Scholar] [CrossRef]
  170. Weiser, J.N.; Pan, N.; McGowan, K.L.; Musher, D.; Martin, A.; Richards, J. Phosphorylcholine on the lipopolysaccharide of Haemophilus influenzae contributes to persistence in the respiratory tract and sensitivity to serum killing mediated by C-reactive protein. J. Exp. Med. 1998, 187, 631–640. [Google Scholar] [CrossRef]
  171. Sun, Z.; Kang, Y.; Norris, M.H.; Troyer, R.M.; Son, M.S.; Schweizer, H.P.; Dow, S.W.; Hoang, T.T. Blocking Phosphatidylcholine Utilization in Pseudomonas aeruginosa, via Mutagenesis of Fatty Acid, Glycerol and Choline Degradation Pathways, Confirms the Importance of This Nutrient Source In Vivo. PLoS ONE 2014, 9, e103778. [Google Scholar] [CrossRef]
  172. Willsey, G.G.; Ventrone, S.; Schutz, K.C.; Wallace, A.M.; Ribis, J.W.; Suratt, B.T.; Wargo, M.J. Pulmonary Surfactant Promotes Virulence Gene Expression and Biofilm Formation in Klebsiella pneumoniae. Infect. Immun. 2018, 86, e00135-18. [Google Scholar] [CrossRef] [PubMed]
  173. Wei, J.; Zhao, X.; Wang, S.; Zhang, M.; Yao, W.; Yuan, Y. Determination of related substances in egg yolk lecithin by HPLC-CAD and characterization of its profiling by HPLC-Q-TOF-MS. J. Pharm. Biomed. Anal. 2022, 221, 115079. [Google Scholar] [CrossRef] [PubMed]
  174. Ganz, T. Antimicrobial polypeptides in host defense of the respiratory tract. J. Clin. Investig. 2002, 109, 693–697. [Google Scholar] [CrossRef] [PubMed]
  175. Feingold, D.S.; Goldman, J.N.; Kuritz, H.M. Locus of the action of serum and the role of lysozyme in the serum bactericidal reaction. J. Bacteriol. 1968, 96, 2118–2126. [Google Scholar] [CrossRef] [PubMed]
  176. Arnold, R.R.; Cole, M.F.; McGhee, J.R. A bactericidal effect for human lactoferrin. Science 1977, 197, 263–265. [Google Scholar] [CrossRef] [PubMed]
  177. Drago-Serrano, M.E.; de la Garza-Amaya, M.; Luna, J.S.; Campos-Rodríguez, R. Lactoferrin-lipopolysaccharide (LPS) binding as key to antibacterial and antiendotoxic effects. Int. Immunopharmacol. 2012, 12, 1–9. [Google Scholar] [CrossRef] [PubMed]
  178. Aratani, Y. Myeloperoxidase: Its role for host defense, inflammation, and neutrophil function. Arch. Biochem. Biophys. 2018, 640, 47–52. [Google Scholar] [CrossRef] [PubMed]
  179. Wilkinson, T.S.; Morris, A.C.; Kefala, K.; O’Kane, C.M.; Moore, N.R.; Booth, N.A.; McAuley, D.F.; Dhaliwal, K.; Walsh, T.S.; Haslett, C.; et al. Ventilator-Associated Pneumonia Is Characterized by Excessive Release of Neutrophil Proteases in the Lung. Chest 2012, 142, 1425–1432. [Google Scholar] [CrossRef]
  180. Jomova, K.; Valko, M. Advances in metal-induced oxidative stress and human disease. Toxicology 2011, 283, 65–87. [Google Scholar] [CrossRef]
  181. Kinnula, V.L.; Crapo, J.D. Superoxide dismutases in the lung and human lung diseases. Am. J. Respir. Crit. Care Med. 2003, 167, 1600–1619. [Google Scholar] [CrossRef] [PubMed]
  182. Reid, D.W.; Carroll, V.; O’May, C.; Champion, A.; Kirov, S.M. Increased airway iron as a potential factor in the persistence of Pseudomonas aeruginosa infection in cystic fibrosis. Eur. Respir. J. 2007, 30, 286–292. [Google Scholar] [CrossRef] [PubMed]
  183. Smith, D.J.; Anderson, G.J.; Bell, S.C.; Reid, D.W. Elevated metal concentrations in the CF airway correlate with cellular injury and disease severity. J. Cyst. Fibros. 2014, 13, 289–295. [Google Scholar] [CrossRef]
  184. Sunder-Plassmann, G.; Patruta, S.I.; Horl, W.H. Pathobiology of the role of iron in infection. Am. J. Kidney Dis. 1999, 34, S25–S29. [Google Scholar] [CrossRef] [PubMed]
  185. Bussiere, F.I.; Gueux, E.; Rock, E.; Girardeau, J.P.; Tridon, A.; Mazur, A.; Rayssiguier, Y. Increased phagocytosis and production of reactive oxygen species by neutrophils during magnesium deficiency in rats and inhibition by high magnesium concentration. Br. J. Nutr. 2002, 87, 107–113. [Google Scholar] [CrossRef] [PubMed]
  186. Gray, R.D.; Duncan, A.; Noble, D.; Imrie, M.; O’Reilly, D.S.; Innes, J.A.; Porteous, D.J.; Greening, A.P.; Boyd, A.C. Sputum trace metals are biomarkers of inflammatory and suppurative lung disease. Chest 2010, 137, 635–641. [Google Scholar] [CrossRef]
  187. Stites, S.W.; Walters, B.; O’Brien-Ladner, A.R.; Bailey, K.; Wesselius, L.J. Increased iron and ferritin content of sputum from patients with cystic fibrosis or chronic bronchitis. Chest 1998, 114, 814–819. [Google Scholar] [CrossRef] [PubMed]
  188. Stites, S.W.; Plautz, M.W.; Bailey, K.; O’Brien-Ladner, A.R.; Wesselius, L.J. Increased concentrations of iron and isoferritins in the lower respiratory tract of patients with stable cystic fibrosis. Am. J. Respir. Crit. Care Med. 1999, 160, 796–801. [Google Scholar] [CrossRef] [PubMed]
  189. Pegg, A.E. Functions of Polyamines in Mammals. J. Biol. Chem. 2016, 291, 14904–14912. [Google Scholar] [CrossRef]
  190. Shah, P.; Swiatlo, E. A multifaceted role for polyamines in bacterial pathogens. Mol. Microbiol. 2008, 68, 4–16. [Google Scholar] [CrossRef]
  191. Chou, H.T.; Li, J.; Peng, Y.; Lu, C. Molecular characterization of PauR and its role in control of putrescine and cadaverine catabolism through the γ-glutamylation pathway in Pseudomonas aeruginosa PAO1. J. Bacteriol. 2013, 195, 3906–3913. [Google Scholar] [CrossRef] [PubMed]
  192. Twomey, K.B.; Alston, M.; An, S.Q.; O’Connell, O.J.; McCarthy, Y.; Swarbreck, D.; Febrer, M.; Dow, J.M.; Plant, B.J.; Ryan, R.P. Microbiota and metabolite profiling reveal specific alterations in bacterial community structure and environment in the cystic fibrosis airway during exacerbation. PLoS ONE 2013, 8, e82432. [Google Scholar] [CrossRef] [PubMed]
  193. Pathak, K.V.; McGilvrey, M.I.; Hu, C.K.; Garcia-Mansfield, K.; Lewandoski, K.; Eftekhari, Z.; Yuan, Y.C.; Zenhausern, F.; Menashi, E.; Pirrotte, P. Molecular Profiling of Innate Immune Response Mechanisms in Ventilator-associated Pneumonia. Mol. Cell. Proteom. 2020, 19, 1688–1705. [Google Scholar] [CrossRef] [PubMed]
  194. Holz, O.; DeLuca, D.S.; Roepcke, S.; Illig, T.; Weinberger, K.M.; Schudt, C.; Hohlfeld, J.M. Smokers with COPD Show a Shift in Energy and Nitrogen Metabolism at Rest and During Exercise. Int. J. Chron. Obstruct. Pulmon. Dis. 2020, 15, 1–13. [Google Scholar] [CrossRef] [PubMed]
  195. Hasan, C.M.; Pottenger, S.; Green, A.E.; Cox, A.A.; White, J.S.; Jones, T.; Winstanley, C.; Kadioglu, A.; Wright, M.H.; Neill, D.R.; et al. Pseudomonas aeruginosa utilises the host-derived polyamine spermidine to facilitate antimicrobial tolerance. JCI Insight 2022, 7, e158879. [Google Scholar] [CrossRef] [PubMed]
  196. Kwon, D.H.; Lu, C.D. Polyamine effects on antibiotic susceptibility in bacteria. Antimicrob. Agents Chemother. 2007, 51, 2070–2077. [Google Scholar] [CrossRef] [PubMed]
  197. Innes, A.L.; Carrington, S.D.; Thornton, D.J.; Kirkham, S.; Rousseau, K.; Dougherty, R.H.; Raymond, W.W.; Caughey, G.H.; Muller, S.J.; Fahy, J.V. Ex vivo sputum analysis reveals impairment of protease-dependent mucus degradation by plasma proteins in acute asthma. Am. J. Respir. Crit. Care Med. 2009, 180, 203–210. [Google Scholar] [CrossRef] [PubMed]
  198. Sankar, S.; Yamaguchi, M.; Kawabata, S.; Ponnuraj, K. Streptococcus pneumoniae Surface Adhesin PfbA Exhibits Host Specificity by Binding to Human Serum Albumin but Not Bovine, Rabbit and Porcine Serum Albumins. Protein J. 2020, 39, 1–9. [Google Scholar] [CrossRef] [PubMed]
  199. Egesten, A.; Frick, I.M.; Mörgelin, M.; Olin, A.I.; Björck, L. Binding of Albumin Promotes Bacterial Survival at the Epithelial Surface. J. Biol. Chem. 2011, 286, 2469–2476. [Google Scholar] [CrossRef]
  200. Teevan-Hanman, C.; O’Shea, P. Candida albicans exhibit two classes of cell surface binding sites for serum albumin defined by their affinity, abundance and prospective role in interkingdom signalling. PLoS ONE 2021, 16, e0254593. [Google Scholar] [CrossRef]
  201. Austermeier, S.; Pekmezovic, M.; Porschitz, P.; Lee, S.; Kichik, N.; Moyes, D.L.; Ho, J.; Kotowicz, N.K.; Naglik, J.R.; Hube, B.; et al. Albumin Neutralizes Hydrophobic Toxins and Modulates Candida albicans Pathogenicity. mBio 2021, 12, e0053121. [Google Scholar] [CrossRef] [PubMed]
  202. Kruczek, C.; Wachtel, M.; Alabady, M.S.; Payton, P.R.; Colmer-Hamood, J.A.; Hamood, A.N. Serum albumin alters the expression of iron-controlled genes in Pseudomonas aeruginosa. Microbiology 2012, 158, 353–367. [Google Scholar] [CrossRef] [PubMed]
  203. Smith, A.C.; Rice, A.; Sutton, B.; Gabrilska, R.; Wessel, A.K.; Whiteley, M.; Rumbaugh, K.P. Albumin Inhibits Pseudomonas aeruginosa Quorum Sensing and Alters Polymicrobial Interactions. Infect. Immun. 2017, 85, e00116-17. [Google Scholar] [CrossRef] [PubMed]
  204. Park, W.K.; Chung, J.W.; Kim, Y.K.; Chung, S.C.; Kho, H.S. Influences of animal mucins on lysozyme activity in solution and on hydroxyapatite surfaces. Arch. Oral Biol. 2006, 51, 861–869. [Google Scholar] [CrossRef] [PubMed]
  205. Davies, R.C.; Neuberger, A.; Wilson, B.M. The dependence of lysozyme activity on pH and ionic strength. Biochim. Biophys. Acta 1969, 178, 294–305. [Google Scholar] [CrossRef] [PubMed]
  206. Pahud, J.J.; Schellenberg, D.; Monti, J.C.; Scherz, J.C. Lysozyme, an abomasal enzyme in the ruminants. Ann. Rech. Vet. 1983, 14, 493–501. [Google Scholar] [PubMed]
  207. Chirkov, N.S.; Lin, S.-Y.; Michailov, A.V.; Miller, R.; Noskov, B.A. DNA Penetration into a Lysozyme Layer at the Surface of Aqueous Solutions. Int. J. Mol. Sci. 2022, 23, 12377. [Google Scholar] [CrossRef] [PubMed]
  208. Bergamo, A.; Sava, G. Lysozyme: A Natural Product with Multiple and Useful Antiviral Properties. Molecules 2024, 29, 652. [Google Scholar] [CrossRef] [PubMed]
  209. Curry, S.; Brick, P.; Franks, N.P. Fatty acid binding to human serum albumin: New insights from crystallographic studies. Biochim. Biophys. Acta 1999, 1441, 131–140. [Google Scholar] [CrossRef] [PubMed]
  210. Vita, G.M.; De Simone, G.; De Marinis, E.; Nervi, C.; Ascenzi, P.; di Masi, A. Serum albumin and nucleic acids biodistribution: From molecular aspects to biotechnological applications. IUBMB Life 2022, 74, 866–879. [Google Scholar] [CrossRef]
  211. Dubeau, S.; Bourassa, P.; Thomas, T.J.; Tajmir-Riahi, H.A. Biogenic and synthetic polyamines bind bovine serum albumin. Biomacromolecules 2010, 11, 1507–1515. [Google Scholar] [CrossRef] [PubMed]
  212. Bal, W.; Sokolowska, M.; Kurowska, E.; Faller, P. Binding of transition metal ions to albumin: Sites, affinities and rates. Biochim. Biophys. Acta 2013, 1830, 5444–5455. [Google Scholar] [CrossRef] [PubMed]
  213. Kakati, N.; Parashar, C.K.; Thakur, S.; Deshmukh, O.S.; Bandyopadhyay, D. Microrheology of Mucin-Albumin Assembly Using Diffusing Wave Spectroscopy. ACS Appl. Bio Mater. 2022, 5, 4118–4127. [Google Scholar] [CrossRef] [PubMed]
  214. Gong, D.H.; Turner, B.; Bhaskar, K.R.; Lamont, J.T. Lipid binding to gastric mucin: Protective effect against oxygen radicals. Am. J. Physiol. 1990, 259 4 Pt 1, G681–G686. [Google Scholar] [CrossRef] [PubMed]
  215. Conrad, M.E.; Umbreit, J.N.; Moore, E.G. A role for mucin in the absorption of inorganic iron and other metal cations. A study in rats. Gastroenterology 1991, 100, 129–136. [Google Scholar] [CrossRef] [PubMed]
  216. Ghio, A.J.; Soukup, J.M.; Dailey, L.A.; Roggli, V.L. Mucus increases cell iron uptake to impact the release of pro-inflammatory mediators after particle exposure. Sci. Rep. 2023, 13, 3925. [Google Scholar] [CrossRef] [PubMed]
  217. Casale, A.; De Stefano, C.; Manfredi, G.; Milea, D.; Sammartano, S. Sequestration of Alkyltin(IV) compounds in aqueous solution: Formation, stability, and empirical relationships for the binding of dimethyltin(IV) cation by N- and O-donor ligands. Bioinorg. Chem. Appl. 2009, 2009, 219818. [Google Scholar] [CrossRef] [PubMed]
  218. Pelta, J.; Livolant, F.; Sikorav, J.L. DNA aggregation induced by polyamines and cobalthexamine. J. Biol. Chem. 1996, 271, 5656–5662. [Google Scholar] [CrossRef]
  219. Natalini, J.G.; Singh, S.; Segal, L.N. The dynamic lung microbiome in health and disease. Nat. Rev. Microbiol. 2023, 21, 222–235. [Google Scholar] [CrossRef]
  220. Li, R.; Li, J.; Zhou, X. Lung microbiome: New insights into the pathogenesis of respiratory diseases. Sig. Transduct. Target. Ther. 2024, 9, 19. [Google Scholar] [CrossRef]
  221. Morrison, A.G.; Sarkar, S.; Umar, S.; Lee, S.T.M.; Thomas, S.M. The Contribution of the Human Oral Microbiome to Oral Disease: A Review. Microorganisms 2023, 11, 318. [Google Scholar] [CrossRef] [PubMed]
  222. Cardenas, P.A.; Cooper, P.J.; Cox, M.J.; Chico, M.; Arias, C.; Moffatt, M.F.; Cookson, W.O. Upper airways microbiota in antibiotic-naive wheezing and healthy infants from the tropics of rural Ecuador. PLoS ONE 2012, 7, e46803. [Google Scholar] [CrossRef] [PubMed]
  223. Huang, Y.J.; Marsland, B.J.; Bunyavanich, S.; O’Mahony, L.; Leung, D.Y.; Muraro, A.; Fleisher, T.A. The microbiome in allergic disease: Current understanding and future opportunities-2017 PRACTALL document of the American Academy of Allergy, Asthma & Immunology and the European Academy of Allergy and Clinical Immunology. J. Allergy Clin. Immunol. 2017, 139, 1099–1110. [Google Scholar] [CrossRef] [PubMed]
  224. Lyon, J. The Lung Microbiome: Key to Respiratory Ills? JAMA 2017, 317, 1713–1714. [Google Scholar] [CrossRef] [PubMed]
  225. Larsen, J.M. The immune response to Prevotella bacteria in chronic inflammatory disease. Immunology 2017, 151, 363–374. [Google Scholar] [CrossRef] [PubMed]
  226. Sze, M.A.; Dimitriu, P.A.; Hayashi, S.; Elliott, W.M.; McDonough, J.E.; Gosselink, J.V.; Cooper, J.; Sin, D.D.; Mohn, W.W.; Hogg, J.C. The lung tissue microbiome in chronic obstructive pulmonary disease. Am. J. Respir. Crit. Care Med. 2012, 185, 1073–1080. [Google Scholar] [CrossRef] [PubMed]
  227. Boutin, S.; Graeber, S.Y.; Stahl, M.; Dittrich, S.A.; Mall, M.A.; Dalpke, A.H. Chronic but not intermittent infection with Pseudomonas aeruginosa is associated with global changes of the lung microbiome in cystic fibrosis. Eur. Respir. J. 2017, 50, 1701086. [Google Scholar] [CrossRef] [PubMed]
  228. Woo, S.; Park, S.Y.; Kim, Y.; Jeon, J.P.; Lee, J.J.; Hong, J.Y. The Dynamics of Respiratory Microbiota during Mechanical Ventilation in Patients with Pneumonia. J. Clin. Med. 2020, 9, 638. [Google Scholar] [CrossRef] [PubMed]
  229. Cuthbertson, L.; Craven, V.; Bingle, L.; Cookson, W.; Everard, M.L.; Moffatt, M.F. The impact of persistent bacterial bronchitis on the pulmonary microbiome of children. PLoS ONE 2017, 12, e0190075. [Google Scholar] [CrossRef] [PubMed]
  230. Madapoosi, S.S.; Cruickshank-Quinn, C.; Opron, K.; Erb-Downward, J.R.; Begley, L.A.; Li, G.; Barjaktarevic, I.; Barr, R.G.; Comellas, A.P.; Couper, D.J.; et al. Lung Microbiota and Metabolites Collectively Associate with Clinical Outcomes in Milder Stage Chronic Obstructive Pulmonary Disease. Am. J. Respir. Crit. Care Med. 2022, 206, 427–439. [Google Scholar] [CrossRef]
Figure 1. Schematic of the constituent components of healthy and infected ASL (sputum). (A) Healthy ASL, composed of a periciliary layer which bathes ciliated epithelium. Above the periciliary layer is a mucus layer, predominantly composed of mucin. Other components in the ASL include lipids, low levels of glucose, amino acids, metals, and DNA. (B) ASL infected with invading pathogens. In response to this, more mucin is produced to entrap these pathogens. Antimicrobial peptides are also produced by epithelial and submucosal cells to kill entrapped bacteria and fungi. DNA is present in inflamed ASL in higher amounts due to DNA released by dead host cells and microorganisms, as well as DNA released by neutrophils during the production of neutrophil extracellular traps. Vascular leakage introduces serum albumin into the ASL and causes an increase in ASL glucose concentrations. Airway inflammation has also been shown to result in higher levels of amino acids and metals being present in ASL. ASL pH becomes acidified in the ASL of inflamed airways, this alters ion concentrations and causes ASL dehydration, resulting in a thinner periciliary layer. Created with BioRender.com.
Figure 1. Schematic of the constituent components of healthy and infected ASL (sputum). (A) Healthy ASL, composed of a periciliary layer which bathes ciliated epithelium. Above the periciliary layer is a mucus layer, predominantly composed of mucin. Other components in the ASL include lipids, low levels of glucose, amino acids, metals, and DNA. (B) ASL infected with invading pathogens. In response to this, more mucin is produced to entrap these pathogens. Antimicrobial peptides are also produced by epithelial and submucosal cells to kill entrapped bacteria and fungi. DNA is present in inflamed ASL in higher amounts due to DNA released by dead host cells and microorganisms, as well as DNA released by neutrophils during the production of neutrophil extracellular traps. Vascular leakage introduces serum albumin into the ASL and causes an increase in ASL glucose concentrations. Airway inflammation has also been shown to result in higher levels of amino acids and metals being present in ASL. ASL pH becomes acidified in the ASL of inflamed airways, this alters ion concentrations and causes ASL dehydration, resulting in a thinner periciliary layer. Created with BioRender.com.
Microorganisms 12 00732 g001
Table 1. Concentrations of different ASL components detected in different airways pathologies.
Table 1. Concentrations of different ASL components detected in different airways pathologies.
ComponentAverage ConcentrationsConditionsDetection Method
Mucin2.7 mg/mL [61], 6.5 mg/mL [61], ~5 mg/mL [15], ~20 mg/mL [15]. Healthy [61], CF exacerbation [61], newly intubated elective laryngoscopy patients [15], ICU patients mechanically ventilated for at least 4 days [15]. Size exclusion chromatography/differential refractive index [61], enzyme-linked immunosorbent assay of subglottic samples [15].
DNA0.96 mg/mL [55], 6.7mg/mL [55], 5.2 mg/mL [55], 20 ng/mL [15,92], 40 ng/mL [15], ~100 ng/mL [92], ~250 ng/mL [92], 0.7 µg/mL [104], 3.2 µg/mL [104], 5.4 µg/mL [104], 2 µg/mL [105], 10 µg/mL [105], 416% more DNA by area in CF sputum compared to asthma and chronic bronchitis sputum [60].Healthy [55,92,105], stable CF [55,105], CF exacerbation [55], non-ICU patients [15], ICU patients [15], VAP [92], ARDS [92], VAP and ARDS [92], non-CF patients [104], infants with CF [104], older CF patients [104].Microfluorimetry [55], fluorometric assays [15], colorimetric assays [92], Hoechst dye-binding assay [104],
Quant-iT PicoGreen assay [105],
confocal microscopy [60].
pH6.2–7 (nasal) [106], 7.1 (lower airways) [106], 6.78 [107], 7.18 [108], 6.57 [108], 6.97 [109], 6.58 [110], 6.62 [110], 6.72 [111], 6.61 [110], 6.89 [112].Non-CF [106,107,108,109,110], CF [106,108], pneumonia [110,111], chronic lung disease [110], acute exacerbation of COPD [112].Monocrystalline antimony catheter [106], in-gold combined pH-glass electrode [106], fluorescent indicators on freshly excised human bronchi [107,109], fluorescent indicators on nasal biopsies [108], pH electrode [110], pH strips [112].
Glucose0.4 mM [113], 1 mM [114], 4 mM [114], 1.2 mM [113], 2 mM [113], 3.5 mM [115].Healthy [113], viral infection [114], hyperglycaemic diabetes [114], CF [113], CF and diabetes [113], mechanically ventilated patients [115].High-performance anion-exchange chromatography with pulsed amperometric detection [113], glucose oxidase sticks [115].
Amino acids2.52 mg/mL; total [116], 5.7 mg/mL; total [116], 12.3 mM; total [117], 18.2 mM; total [117], 0.42 nmol/mg; alanine [47], 2.2 nmol/mg; asparagine [47], 0.42 nmol/mg; glutamine [47], 1.06 nmol/mg; glycine [47], 0.43 nmol/mg; lysine [47], 0.13 nmol/mg; valine [47].Healthy [116], CF [116,117], CF exacerbation [117], healthy tissue from lobectomies of lung cancer patients [47].Thin layer chromatography [116], high-performance liquid chromatography [117], nuclear magnetic resonance [47].
Lysozyme3.9 µg/mL [118], 9.1 µg/mL [118].Culture-negative CF patients [118], culture-positive CF patients [118].Lysozyme activity assay [118].
Lactoferrin5 µg/mL [61], 9 µg/mL [61], 3.0 µg/mL [118], 22.3 µg/mL [118].Non-CF [61], CF [61], culture-negative CF patients [118], culture-positive CF patients [118].Immunologic techniques [61], enzyme-linked immunosorbent assay [118].
Ferritin0.2 µg/mL [119], 2.4 µg/mL [119], 3.6 µg/mL [119], 0.6 µg/mL [119].Healthy [119], CF [119], CF exacerbation [119], COPD [119].Microparticle enzyme immunoassay [119].
Putrescine11.91 µmol/L [120], 6.18 µmol/L [120], 96.02 µmol/L [120], 20.59 µmol/L [120].Healthy [120], CF stable [120], CF exacerbation pre-antibiotic treatment [120], CF exacerbation post-antibiotic treatment [120].High-performance liquid chromatography [120].
Spermine0.22 µmol/L [120], 1.71 µmol/L [120], 7.32 µmol/L [120], 1.35 µmol/L [120].Healthy [120], CF stable [120], CF exacerbation pre-antibiotic treatment [120], CF exacerbation post-antibiotic treatment [120].High-performance liquid chromatography [120].
Spermidine0.88 µmol/L [120], 1.62 µmol/L [120], 0.78 µmol/L [120], 0.62 µmol/L [120].Healthy [120], CF stable [120], CF exacerbation pre-antibiotic treatment [120], CF exacerbation post-antibiotic treatment [120].High-performance liquid chromatography [120].
Serum albumin0.1 dg/L [119], 0.4 dg/L [119], 0.7 dg/L [119], 0.2 dg/mL [119]. 127.4 µg/mL [121], 244.4 µg/mL [121].Healthy [119,121], CF [119,121], CF exacerbation [119], COPD [119].Rate immunophelometry [119], competitive radioimmunoassay [121].
Table 2. Concentrations of metals (µg/L) in different pulmonary disease across different studies.
Table 2. Concentrations of metals (µg/L) in different pulmonary disease across different studies.
MetalHealthyAsthmaCOPDCFBronchiectasisSmoker
Zn15.35 (10.4–25.6) a
179 (103–597) b
40.45 (20.99) c
12.7 (7.2–41.4) a 25.4 (9.8–50.7) a135.3 (54.2–209.6) a
1285 (678–1811) b
111.3 (46.1–150.7) a
537 (401–838) b
48.16 (35.06) c
Fe13.5 (8.6–21.5) a
0 (0–37) b
6.38 (9.12) c
30 (6.9–35.3) a21.3 (3.1–35.6) a56.9 (24.3–115.3) a
797 (398–1292) b
54.2 (22.7–91.6) a
1075 (862–1324) b
23.37 (28.47) c
Mn0 (0–0.25) a
5 (2–9) b
0.12 (0.16) c
0.8 (0.2–1.7) a0 (0–0.7) a0.3 (0.1–0.8) a
6 (4–17) b
0.6 (0.2–1.3) a
6 (4–10) b
0.21 (0.24) c
Cu8.6 (3–16.4) a
106 (55.3–196) b
4.77 (4.9) c
15.2 (8.6–29.5) a15.2 (12.2–22) a19.5 (14.5–30.1) a
173 (128–257) b
15.7 (10.9–33.3) a
226 (130–314) b
4 (2.26) c
Ca45,000 (28,000–58,000) b
811.7 (181.61) c
102,000 (76,000–123,000) b124,000 (78,000–156,000) b856.8 (183.34) c
Mg4000 (2000–7000) b
389.06 (66.93) c
30,000 (19,000–44,000) b33,000 (27,000–39,000) b428.76 (110.42) c
a denotes metal concentrations quantified in [186]. Metals detected by inductively coupled plasma optical emission spectrometry. Data shown as medians (interquartile range). b denotes metal concentrations quantified in [183]. Metals detected by inductively coupled plasma mass-spectrometry. Data shown as medians (interquartile range). c denotes metal concentrations quantified in [14]. Metals detected by inductively coupled plasma mass-spectrometry. Data shown as means (standard deviation).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Walsh, D.; Bevan, J.; Harrison, F. How Does Airway Surface Liquid Composition Vary in Different Pulmonary Diseases, and How Can We Use This Knowledge to Model Microbial Infections? Microorganisms 2024, 12, 732. https://doi.org/10.3390/microorganisms12040732

AMA Style

Walsh D, Bevan J, Harrison F. How Does Airway Surface Liquid Composition Vary in Different Pulmonary Diseases, and How Can We Use This Knowledge to Model Microbial Infections? Microorganisms. 2024; 12(4):732. https://doi.org/10.3390/microorganisms12040732

Chicago/Turabian Style

Walsh, Dean, Jennifer Bevan, and Freya Harrison. 2024. "How Does Airway Surface Liquid Composition Vary in Different Pulmonary Diseases, and How Can We Use This Knowledge to Model Microbial Infections?" Microorganisms 12, no. 4: 732. https://doi.org/10.3390/microorganisms12040732

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop