3,971
Views
30
CrossRef citations to date
0
Altmetric
Review Articles

Assays and enumeration of bioaerosols-traditional approaches to modern practices

, , , , , , & show all
Pages 611-633 | Received 10 Jul 2019, Accepted 21 Jan 2020, Published online: 24 Feb 2020

Abstract

In recent years, much attention has been drawn to the exposure to bioaerosols in both occupational and indoor environments due to adverse effects on human health. Exposure to these agents may cause infectious diseases, allergic diseases, acute toxic effects, respiratory diseases, neurological effects, and cancer. Bioaerosols play a significant role in air quality studies, as well as in industrial and agricultural regulations. There is a diversity of sources for bioaerosol exposures, which include occupational activities such as waste disposal, sorting and composting, agricultural and food processing, livestock production and handling, healthcare, and fungal growth following flooding. Bioaerosol monitoring includes the measurement of viable (culturable and nonculturable) and nonviable microorganisms in both indoor and outdoor environments. The development of standardized methods for the detection and quantitation of bioaerosols has become an important issue. Technologies, including microbial plating, physical-chemical assays, and molecular techniques, have been adapted for the assessment of collected bioaerosol particles. In this review, we will discuss traditional and modern assays that have been applied in bioaerosol enumeration based on culturability, optical properties, immunoreactions, and nucleic acid amplification and sequencing of biological particles. The analysis of additional characteristics of bioaerosols including microbiome and antimicrobial resistance is also discussed. Finally, concluding thoughts are offered regarding the challenges and perspectives in the field.

Copyright © 2020 American Association for Aerosol Research

1. Brief history and summary of topic

In recent years, a growing number of studies have focused on the assessment of exposure to biological agents in occupational and indoor environments due to their various negative effects on human health (Qian et al. Citation2012; Van Leuken et al. Citation2016). Although there has been a recent expansion of access to data, models, and management guidelines via the Internet, there is still a need for updating specific studies of aerobiology (Kakde Citation2012).

In 1861, French scientist Louis Pasteur recorded the first measurements of airborne organisms (Eduard et al. Citation2012). A century later, the inhalation of spores from thermophilic microorganisms was identified in the development of farmers’ lung, while endotoxins from Gram-negative bacteria were connected to byssinosis in cotton workers (Eduard et al. Citation2012). From 1985 to 2003, the International Commission on Occupational Health (ICOH) Organic Dust Committee organized four meetings which made important contributions to recognizing organic dust-related diseases by identifying the differences between the potentially disabling disease, hypersensitivity pneumonitis, and the benign organic dust toxic syndrome (Rylander, Peterson, and Donham Citation1990).

The health effects of the bioaerosols and the nature of the disease (infections vs. allergies vs. toxic effects) play a significant role in determining the analytical method to be used. Fungi and bacteria have been found to produce diverse human health problems including irritations, infections, and allergies (Kim, Kabir, and Jahan Citation2018). Symptoms of fatigue, cognitive difficulties or memory loss as well as common diseases such as allergy, asthma, and hypersensitivity pneumonitis are caused by fungal exposure (Siersted and Gravesen Citation1993; Selman et al. Citation2010; Lee et al. Citation2006). There is a vast diversity of fungi, with more than an estimated one million species that produce airborne spores, conidia, hyphae, and other fragments that can affect human health (Lee et al. Citation2010). According to Green, Tovey, et al. (Citation2006), about 112 fungal genera, belonging to three taxonomic groups known as Ascomycota, Basidiomycota, and Deuteromycota, release allergens. In past studies, high concentrations of Aspergillus niger have been found in aerosol samples from homes where asthmatic children reside (Mousavi et al. Citation2016). Other fungal species, such as Aspergillus ochraceus, Aspergillus unguis, and Penicillium variabile, have also been found in dust associated with childhood asthma (Reponen et al. Citation2012). Disruption of fungi caused by daily activities usually causes the production of aerosolized fungal fragments, which are broken or fractured conidia and hyphae (Green, Tovey, et al. Citation2006). Bacterial cell wall components, including endotoxins in Gram-negative bacteria and the less studied peptidoglycans in Gram positive bacteria may induce respiratory symptoms (Douwes et al. Citation2003).

Viruses, just like any microorganism, can become airborne. They are among the smallest of bioaerosols, with diameters in the nanoregime (>20 nm) (Duan Citation2006). However, viruses are more likely attached to other airborne particles (e.g., Yang, Elankumaran, and Marr Citation2011). They can remain infectious outside their hosts for prolonged periods of time (Pirtle and Beran Citation1991), leading to infections by indirect contact with surfaces that became contaminated by deposited infectious droplets (Verreault, Moineau, and Duchaine Citation2008).

Allergens are different macromolecular structures comprising mostly of proteins of biological origin, including IgE binding enzymes from fungi and bacteria (Cullinan et al. Citation2000). Plant pollens are similarly IgE binding allergens that may cause allergic reactions in greenhouse workers (Sinha et al. Citation2014). Animal proteins, including allergens from dust mite, cat, mouse and rat can also trigger immune responses (Platts-Mills and Woodfolk Citation2011; Kuehn and Hilger Citation2015).

When studying bioaerosols at field sites it is useful to monitor environmental factors as well, to fully understand the source of bioaerosols (Mandal and Brandl Citation2011). Stetzenbach, Buttner, and Cruz (Citation2004) concluded that there is evidence for the association between bioaerosol levels and certain environmental factors, such as temperature and relative humidity. Seasonal variations significantly affect indoor microbial exposures, which are influenced by temperature, relative humidity, and air exchange rates (Talley, Coley, and Kursar Citation2002; Frankel et al. Citation2012; Dedesko and Siegel Citation2015). Everyday ordinary human activities such as talking, walking, sneezing, showering, and cleaning floors also contribute to the generation of bioaerosols (Chen and Hildemann Citation2009, Estrada-Perez et al. Citation2018).

There are various levels that bioaerosols can be characterized or quantified. Bioaerosols are categorized into two main groups: viable and nonviable (), where viable organisms are able to reproduce, whereas nonviable cannot reproduce. Viable bioaerosols can be further categorized as either culturable or nonculturable. Culturable bioaerosols, including fungi, bacteria, and viruses are microorganisms that are able to reproduce in a controlled environment (Jensen and Schafer Citation1998).

Table 1. Comparison of air sampler performances for sample analysis. Optimal performance is indicated by light gray color, acceptable by medium gray and low by dark gray.

Recently, Pearson et al. (Citation2015) identified a diversity of sources for bioaerosol exposures, including occupational activities such as waste sorting and composting, agricultural and food processing, livestock handling, healthcare and others. In addition, reports indicate that workers who are susceptible to such exposure exhibit symptoms of various respiratory diseases or allergies (Beck, Young, and Huffnagle Citation2012; Rohr et al. Citation2015; Sanchez-Monedero and Stentiford Citation2003).

2. Collection of environmental samples

The preparation of an environmental bioaerosol sample involves collecting air particles on a solid or in a liquid medium and further testing the collected particles for the presence of bioaerosols. Several particle collection methods can be used, with major methods being impaction, impingement, and filtration, while alternative methods include gravitational settling, electrostatic precipitation, and cyclone separation.

Impaction is widely used due to its low cost and is beneficial for direct collection on a solid growth medium with reduced post-sampling processing. However, the sampler can be overloaded if collecting a sample in an area of highly concentrated bioaerosols, making this method more applicable in low concentration areas for enumeration only (Vaughan Citation1989; King and McFarland Citation2012a; Ghosh, Lal, and Srivastava Citation2015). The method involves drawing in air and changing the flow direction causing particles with higher inertia to be impacted on the collecting surface. If the velocity of impact increases, some particles may richochet off the medium resulting in some microorganisms to be damaged and less viable (Ghosh, Lal, and Srivastava Citation2015). In a study by King and McFarland (Citation2012b), a high airflow bioaerosol sampler with a foam collection surface was used for the impaction of aerosolized E. coli and washed between tests to collect the sample. The viability and DNA integrity of the collected bioaerosols was reduced as compared to the collection by liquid impingement using a wetted wall cyclone collector.

The impingement method uses a liquid medium for collection, where the liquid used can vary depending on whether it is an isotonic or buffered solution (Jensen and Schafer Citation1998; Mandal and Brandl Citation2011). Air is drawn through a narrow inlet into the liquid and the collected samples are enumerated by microbial plating or other quantitative methods. Higher total recovery was found for airborne yeast collected by the AGI-30 impinger compared to Nuclepore and gelatin filtration methods where dehydration effects were significant (Lin Citation1999).

Filtration is a frequently used collection method due to its simplicity, low cost, and effectiveness (Cox and Wathes Citation1995). For bioaerosols, the most beneficial filter material is gelatin as it offers a better growth environment for plating and doesn’t inhibit molecular analysis (Ghosh, Lal, and Srivastava Citation2015; Wu, Shen, and Yao Citation2010). However, because of the filter’s small size and the limitation in flow rate they are generally used in personal samplers (Ghosh, Lal, and Srivastava Citation2015). When used as a personal sampler, porous media submerged in a liquid layer and plated after collection showed detection limits of 15 bacterial CFU/m3 and above 5 CFU/m3 for fungi at a sampling time of 8 h (Agranovski et al. Citation2002; Xu et al. Citation2011). Similarly to impaction samplers, filtration samplers used for enumeration in high concentration environments can become overloaded. Samples are removed from the filter and prepared for analysis through resuspension after collection (Hinds Citation1982). Losses can occur during the collection and extraction as some bacterial cells collected cannot survive this process, depending on the sampling time, temperature, and relative humidity or remain attached to the filter (Wang et al. Citation2001). However, vortexing the filters can help in the recovery of bacterial cells from environmental samples.

Gravitational settling is an alternative method of collection where an agar medium is exposed to the environment and airborne microorganisms are allowed to settle onto it due to gravity (Grinshpun, Buttner, and Willeke Citation2007). Post-sampling preparation is not needed; the sample is collected directly on the medium and can be enumerated with minimal interference. This method is only qualitative since the volume of the air is not known, and the sample would be biased to larger particles, which settle more quickly; thus, not giving an accurate idea of all the microorganisms in the environment (Pasquarella, Pitzurra, and Savino Citation2000).

Electrostatic precipitation based bioaerosol collectors when compared to the Biosampler (SKC Inc., Eighty Four, PA, USA), may produce more concentrated culturable samples for bacteria, but lower concentrations of fungal spores. The electric charge applied to the particles can cause some microorganisms to become nonviable (Mainelis et al. Citation2001; Han, Thomas, and Mainelis Citation2018; Bond and Russel Citation2000).

Due to the limitations of traditional approaches that can negatively impact microbial viability, cyclone-based bioaerosol samplers including the BioSampler®, NIOSH two-stage cyclone, Coriolis® and the wetted wall cyclone (WWC) were developed (Willeke, Lin, and Grinshpun Citation1998; Chen et al. Citation2004; Lindsley, Schmechel, and Chen Citation2006; Carvalho et al. Citation2008; McFarland et al. Citation2010). Cyclones capture the airborne microorganisms in a liquid by centrifugal forces applied to the particles as the airstream rotates around the cyclone at increasing velocities (Sigaev et al. Citation2006). With WWCs, real-time samples can be collected while maintaining the culturability and DNA integrity of the collected bioaerosols due to continuous liquid flow in the system (Hu and McFarland Citation2007; McFarland et al. Citation2010; Hubbard et al. Citation2011; King and McFarland Citation2012b). In addition to continuous wetted wall cyclones, there are batch wetted wall cyclones (BWWC), where the particles are collected in liquid batches that are replaced on pre-set time intervals. Similarly to the WWC, BWWC have near-continuous water input that forms a thin film of liquid on the cyclone wall, allowing the air particles to be deposited and transported to the external liquid collection system (McFarland et al. Citation2010; King et al. Citation2009). BWWC systems are advantageous because they support long sampling periods, which allow for more concentrated samples, and unlike the impingement method, the sampler accounts for evaporation (King et al. Citation2009; Yooseph et al. Citation2013).

3. Analysis of culturable bioaerosols

3.1. Fungi

Most culture-based methods are biased to rapidly growing and highly concentrated species; reducing detection of non-culturable and non-viable spores. There is no ideal medium for growth of all fungi (Keswani, Kashon, and Chen Citation2005); consequently, a wide variety of media formulations is used since physical and chemical factors have a large effect on the diagnostic characteristics of fungi (Sharma and Pandey Citation2010). Lignocellulose agar (LCA) was found to be the most suitable for heavy sporulation while potato dextrose agar (PDA) reproduced the most visible colony morphology from decaying vegetable fungi. In most studies, malt extract agar (MEA) and dichloran glycerol 18 (DG18) agar are used for fungal quantification by cultivation (Jürgensen and Madsen Citation2016). In earlier studies, DG18 agar was recommended for the identification and enumeration of fungi, including xerophilic fungi (Jensen and Schafer Citation1998). In traditional studies, MEA and rose bengal agar (RBA) are recommended for collection and enumeration of fungi because of the broad spectrum of fungal growth (Jensen and Schafer Citation1998). However, RBA may make the medium toxic to some fungi depending on whether it is exposed to natural or artificial light. To prevent the growth of bacteria, antibiotics should be added to the medium, such as streptomycin (Jensen and Schafer Citation1998). Certain conditions, such as the temperature of incubation, can produce morphological changes in dimorphic fungi. For example, Histoplasma capsulatum, a thermally dimorphic human pathogen, exists as a spore or mycelial form below 25 °C, but higher temperatures have shown to cause a transition to the yeast form (Jensen and Schafer Citation1998; Sil and Andrianopoulos Citation2015). In a study conducted by Keswani, Kashon, and Chen (Citation2005), Aspergillus fumigatus was cultured on MEA, and Stachybotrys chartarum was cultured on PDA at 25 °C for 10 days to induce sporulation. In a separate study conducted by Górny and Ławniczek-Wałczyk (Citation2012), three isolated fungal species, Aspergillus versicolor, Cladosporium cladosporioides, and Penicillium chrysogenum, commonly found indoors, were grown on MEA for 14 days at 22 °C. In most cases, the incubation temperature remains at a constant 25 °C with varying incubation time depending if sporulation is desired.

Exposure to airborne β(1→3)-glucans, that are glucose polymers originating from most fungi, some bacteria and plants (Stone and Clarke Citation1992) may induce inflammatory responses and has been studied extensively (Rylander Citation1996; Kozajda et al. Citation2017; Adhikari et al. Citation2009).

3.2. Bacteria

The most used analytical technique for monitoring airborne bacteria in both indoor and outdoor environments is the culture-based colony counting method (Jung and Lee Citation2016; Byeon et al. Citation2008; Dong and Yao Citation2010; Park et al. Citation2015). Bioaerosol sampling with a one-stage Andersen impactor was found to have higher recoveries for vegetative bacteria than those of MAS-100 and Burkard samplers (Li and Lin Citation1999). Another advantage of this direct sampling method is the lack of need for post-collection sample processing. However, culture-based colony counting generally requires several days for colony formation following sampling. Additionally, this technique only applies to culturable microbes that are able to divide at a sufficient rate to form colonies. It has been suggested that only 1–2% of all environmental bacteria are culturable (Sharma et al. Citation2005). As a result, this method may underestimate the number of cells due to the presence of viable but non-culturable (VBNC) cells which proliferate under certain conditions (Park et al. Citation2015). Correction factors have been developed to reduce counting errors due to colony masking (Chang et al. Citation1994). A statistical “positive hole correction” table was developed for the well-known six-stage Andersen sampler to evaluate highly loaded plates (Andersen Citation1958). A method was provided to estimate size distributions for both culturable particles (CP) and culturable organisms (CO) from a single sample collected with an Andersen bioaerosol impactor (King and McFarland Citation2012a).

Airborne bacteria experience high stress from the lack of optimal conditions, warm temperatures, and hydration. However, they can grow resistance to certain stresses, due to a stress response gene. This can result in them living longer within areas that have low relative humidity, in the range of 30–40% (Ng, Chan, and Lai Citation2017). The stress response gene increases the survivability of the bacteria in certain conditions; thus, collection and sample preparation is very important for viable bioaerosol samples (Hoisington et al. Citation2014).

It is essential to obtain pure bacterial colonies in most studies in order to perform enumeration, accurate genome sequencing, antibiotic susceptibility analysis, and bacterial gene manipulation and transformation. Nonselective media, such as tryptic soy agar (TSA), casein soy peptone agar (CSPA) and nutrient agar (NA), permit the growth of most microorganisms since no inhibitors are present (Jensen and Schafer Citation1998; Lagier et al. Citation2015). According to Jensen and Schafer (Citation1998), it is preferred to use more than one type of culture medium to collect aerosolized microorganisms; therefore, a nonselective culture medium should be used initially to cultivate all viable and culturable bacteria from a bioaerosol sample. After observation of bacterial growth on a nonselective culture medium, a selective medium is often used to select for specific microorganisms of interest. A selective medium usually contains indicators, specifically of pH, that allow for the recognition of the microorganism of interest due to its particular metabolic activities resulting in pH changes (Atlas Citation2005). A culture-based enumeration method is solely applicable to culturable bacteria, which usually underestimates the number of cells since many bacteria cells are classified as VBNC.

The time and temperature of incubation used for bacterial plate counts are significant variables that affect bacterial growth. The amount of time needed to grow a microcolony of bacteria can range from hours for a fast-growing bacterium to weeks for a drug resistant microorganism (Jensen and Schafer Citation1998). Bacteria found in ambient environmental conditions (e.g., outdoors) are usually incubated at a range of 25 to 30 °C, while those sourced from humans are incubated at 35 to 37 °C. Thermophilic bacteria such as Actinomycetes, are incubated at 50 to 56 °C (Jensen and Schafer Citation1998).

The study by Mirhoseini et al. (Citation2016) utilized two common biosamplers, a single-stage Andersen impactor and an all-glass impinger (AGI), to evaluate levels of airborne bacteria and fungi in indoor environments. Using the colony-forming unit method with TSA and MEA plates at different temperatures, the study showed that particle counting was not an adequate substitute for bioaerosol measurements but should be combined with bioaerosol sampling for rapid detection. The results of Chang and Wang (Citation2014) show that the BioSampler performed better than the 1-stage Andersen and AGI‐30 and using Tween mixture was superior to Phosphate Buffer Saline (PBS) and distilled water for the collection of Staphylococcus aureus bioaerosols.

3.3. Viruses

Cultivation methods for the quantitation of viral bioaerosols includes either human embryonic kidney (HEK) or human embryonic fibroblast tissue cultures (Couch et al. Citation1966). Although culture is often used to determine virus concentrations, most aerosol sampling methods affect viral infectivity, making culture inadequate for calculating the concentrations of airborne viruses (Verreault, Moineau, and Duchaine Citation2008). In laboratory studies bacteriophages of E. coli or other bacteria have been used as virus surrogates, enabling the quantitation by double agar overlay plaque assay (Dubovi and Akers Citation1970; Tseng and Li Citation2005; Tseng and Li Citation2006). The overlay assay has been used to determine the stability of spray dried bacteriophage aerosols that have been the focus of numerous studies due to their potential application for the treatment of pulmonary infections (Pabary et al. Citation2016; Leung et al. Citation2016; Chang et al. Citation2017).

3.4. Enumeration of culturable bioaerosols

Enumeration and calculation of the concentration of pathogenic bioaerosols are crucial to risk assessment (Garre et al. Citation2019). The exposure limit set by the World Health Organization (World Health Organization [WHO]) Citation1990, Citation2002) for total bacteria bioaerosols is 500 CFU/m3. Fungal and bacterial concentrations greater than 104 CFU/m3 are considered a health threat (Heida, Bartman, and van der Zee Citation1995).

A standard method for enumerating bacteria from a sample after collection is to inoculate a small portion on either Luria-Bertani or tryptic soy agar (Missiakas and Schneewind Citation2013; Novoa Rama et al. Citation2018). A study completed to compare different agar types for the drug resistant pathogen Staphylococcus aureus found that TSA had a higher recovery of S. aureus colonies, after being aerosolized, compared to mannitol salt agar, CHROMagar, Chapman stone medium, and Baird–Park agar (Chang and Wang Citation2015). Usually, serial dilutions are plated (Mitsuboshi et al. Citation2016) and incubated at ∼30 °C for 2 days (Gutarowska et al. Citation2015) or at 37 °C for 1 day (Chuang et al. Citation2013), depending on the rate of colony growth (Chuang et al. Citation2013). Optimal incubation time and temperature are culture-specific. One study found that after 3 h of incubation, the CFU count of Lactococcus lactis was smaller than at lower temperatures. However, after 3–12 h the colony count was larger at the lower temperature (Yang and Moon Citation2018). In another study, E. coli was incubated at 37 °C for only 16 h (Huang et al. Citation2018).

Colonies should be counted only for plates with 30–300 CFUs (Suaifan, Alhogail, and Zourob Citation2017). In the study of Kaur et al. (Citation2015), plates with <300 colonies were counted and considered for concentration calculations using a food microbiology standard to calculate the CFU/ml or CFU/m3. Colony-forming units (CFU) of M. tuberculosis measured directly in cough aerosols produced by patients with pulmonary tuberculosis (TB) was found to predict quantitatively IGRA (Interferon gamma release assay) readouts on secondary infection and disease in household contacts of positive TB cases (Acuña-Villaorduña et al. Citation2018). Contacts of high aerosol cases had greater IGRA readouts (median 4.6 IU (International Unit)/mL) when compared to those with low (0.8, 0.2–10) or no aerosol (0.1, 0–3.7; p = 0.08).

3.5. Identification of culturable bioaerosols

While researchers continue to investigate the best methods of identification, a general method for identifying culturable bioaerosols is first collecting air samples, followed by a culturing step to isolate single colonies, then taking another culturing step to grow the different single colonies, and finally identifying the colonies (Duquenne Citation2018). Researchers can identify bioaerosols by macroscopic and microscopic phenotyping (Dawson et al. Citation2015; Xu and Yao Citation2011). For culturable fungal aerosols, microscopy method can be used for the identification of fungal genera based on visual comparison of their images with published fungal morphologies (Bold, Alexopoulos, and Delevoryas Citation1980). Macroscopic phenotyping can give insight into the growth conditions of a microorganism, for example, agar nutrients needed, temperature, light and oxygen requirements (Allen and Waclaw Citation2019). Microscopic phenotyping can give information such as cell shape, color, size, and Gram stain association. Based on microscopy, bacteria can be grouped into Gram-positive cocci, Gram-positive spore bacillus, Gram-positive bacteria, Gram-negative non-enteric bacteria, and Gram-negative intestinal bacteria (Liang et al. Citation2013; Yang, Blair, et al. Citation2016). Gram-positive bacteria have thicker cell walls and Gram-negative cells are covered by an outer membrane outside of the thin cell wall that can be distinguished by staining (Coico Citation2005; Bayraktar et al. Citation2019). This information is important for classifying bacteria but also in treating infections (Kapoor, Saigal, and Elongavan Citation2017).

Several types of biochemical tests, such as catalase, oxygen tolerance, etc., are used to further classify microorganisms by uncovering how a microorganism oxidizes or metabolizes. In one study, a biochemical API 20 A test (www.biomerieux.com) was used for the identification of culturable microorganisms by analyzing the ability of anaerobic bacteria to break down organic substrates and detecting the appropriate metabolites synthesized by the reactions (Duquenne Citation2018). For this purpose, biochemical metabolism tests such as API 20 CH or the API 50 CH are used in clinical settings (Cyprowski et al. Citation2018). Another study used McIlvaine buffer to show the pH influence of the culture medium on the Gluconacetobacter physiology (Yassine et al. Citation2016).

4. Analysis of non-culturable bioaerosols

VBNC and nonviable and nonculturable bioaerosols are classified through different methods than used for culturable microorganisms. These techniques include traditional microbiology methods, microscopy, immunoassays, molecular techniques and mass spectrometry (). Microscopy provides visual images and limited biological information of the collected particles that have to be collected undamaged. Immunoassays based on the analysis of specific or bulk proteins provide quantification of the biological composition (Womiloju et al. Citation2003). Nucleic acid based molecular techniques have been used for the identification of organisms at the species level (Hua and Tong Citation1992). Williams, Ward, and McCartney (Citation2001) have applied conventional polymerase chain reaction (PCR) to analyze air samples for the presence of airborne mycobacteria and fungi commonly associated with adverse health effects. A PCR assay allows for the detection and identification of non-culturable airborne microorganisms (Peccia and Hernandez Citation2006) but does not allow for distinguishing between non-viable and viable microorganisms. To cover the gap between traditional microbial and molecular techniques for bioaerosol monitoring the culture-based analysis was combined with molecular analysis to increase the observed bacterial diversity (Hubad and Lapanje Citation2013). Currently, viability real-time PCR (RT-PCR) analysis is capable of accurate measurements of total microorganism concentrations in environmental samples with the ability to discriminate between live and dead cells by using propidium monoazide (PMA) (Nocker and Camper Citation2009). However, the toxicity of PMA at higher concentrations presents a limitation of the method (Taylor, Bentham, and Ross Citation2014). The advantage of RT-PCR is the ability of rapid sample quantification and species-specific identification (An, Mainelis, and White Citation2006). Microarrays, based on multiplexed 16S PCR reactions can identify a large number of genes if their sequence is known (Brodie et al. Citation2006). Whole genome sequencing and metagenomic analysis offer broad information about a bioaerosol sample (Boissy et al. Citation2014). The molecular assays offer the potential of real-time identification of single bioaerosol particles, however, the complexity of assays makes it difficult to detect single particles.

Table 2. Current practices for the assays and enumeration of bioaerosols.

Endotoxin assays, where the major cell wall component of gram-negative bacteria, the lipopolysaccharide (LPS) is detected based on a colorimetric method, can provide researchers with valid endotoxin exposure estimates (Zhang and Ghosh Citation2000). Although the Limulus Amebocyte Lysate (LAL) assay has been used most extensively to quantify endotoxins (Milton et al. Citation1990), other methods including the whole blood assay (Liebers et al. Citation2009) and the toll-like receptor (TLR4) assay (Peters, Fritz, and Bufe Citation2012) are also available (Duquenne, Marchand, and Duchaine Citation2013).

4.1. Microscopy

Bright-field microscopy is the oldest and best-known microscope technique, using a simple microscope with front or backlighting with basic slides. Standard bright-field microscopy uses dyes to enhance the contrast between the specimen and its surroundings and by observing characteristics of the cells, help identify them. This technique is limited by the optical resolution of the microscope (no less than 200 nm) and the contrast method applied.

Further research is being conducted to use machine learning and visual processing to identify microorganisms on a slide or other medium. The simple collection, mounting, and sensing requirements for bright-field microscopy mean that more effort can be spent on the computing requirements, rather than equipment setup. During an intentional release of a pathogen, topological similarities can be quickly examined for a monodisperse bioaerosol and identified with high accuracy (Wagner and Macher Citation2012). For mixed bioaerosols such as pollen and spores, image recognition is less precise but still useful. A holographic technique was demonstrated with up to 96% precision when identifying pollens and mold spores with the highest consistency for the largest particles (Wu et al. Citation2018).

Phase-contrast microscopy is a variant of bright-field microscopy that takes advantage of the variation in the refractive index between a microorganism and its surrounding medium to enhance contrast and provide easier viewing (Zernike Citation1942; Morris Citation1995). Phase-contrast microscopy can be used to observe live microorganisms, but is particularly useful for imaging low-contrast specimens. Fluorescence is a powerful tool that can be used in microscopy to examine particular structures or molecules in a microorganism or its surroundings. Fluorescence can be achieved either with fluorescent dyes or with cells that are naturally or artificially fluorescent. A variant of fluorescent microscopy is confocal microscopy, which allows very precise resolution and location of fluorescent molecules in a cell. Epifluorescence microscopy has been used to determine viral abundance in bioaerosols (Michaud et al. Citation2018).

Visible light microscopy as described in the previous sections has a resolution limit of approx. 200 nm. Any structural detail smaller than this scale cannot be observed (Morris Citation1995).

Electron microscopy is a technique that uses a beam of electrons to observe a target, and is capable of considerably higher magnifications, with a practical limit around 0.5 nm. Because an electron beam is dispersed by air, the highest pressure an electron microscope can operate at is about 2.7 kPa.

Scanning electron microscopy is currently the most useful electron microscopy method for identifying and counting bioaerosols. Because it observes the surface of an object and does not require thin-slicing like transmission microscopy, it is also more accessible. SEM analysis enables the morphological classification of bioaerosols (Vestlund et al. Citation2014), however, the assignment of a large number of collected particles can be challenging (Wittmaack et al. Citation2005).

Recently, cryogenic transmission electron microscopy (cryo-TEM) has been used for structural studies and identification of biological particles (Patterson et al. Citation2016).

There is no currently available electron microscopy method to observe live microorganisms, but collected particles can be examined for classification, counting, and physical characteristics. Due to the significantly higher magnification available from an electron microscope, even very small particles can be identified with high confidence (Straumfors et al. Citation2014).

4.2. Endotoxin and β-D-glucan assays

The chemical structure and toxicity of endotoxins differ across species of Gram-negative bacteria. Using endotoxin from E. coli is typically used as a reference to assess the combined activity of endotoxins with the Limulus amebocyte lysate (LAL) assay using an enzyme system from the horseshoe crab (Spaan et al. Citation2007). However, the LAL method is prone to interlaboratory variations and can be used only for water-soluble endotoxins (Chun et al. Citation2006). Monoclonal antibody-based methods have also been developed but are less sensitive than the LAL assay (Eduard et al. Citation2012). Endotoxins can also be estimated by gas chromatography-mass spectrometry using 3-hydroxy fatty acids as chemical markers (Spaan et al. Citation2008).

Two methods have been described for the measurement of (1→3)β-D-glucans, that originate from most fungi, based on the LAL assay (Aketagawa et al. Citation1993) and an enzyme immunoassay (Douwes et al. Citation1996).

4.3. Bioluminescence—ATP assay

One of the potential methods to reduce processing time in bioaerosol analysis is the bioluminescence-based technique that detects the presence of adenosine triphosphate (ATP), the basic energy molecule present in all types of living organisms. The method uses the firefly enzyme luciferase to catalyze a reaction between its substrate D-luciferin and ATP, causing luciferin to emit photons in the 500 nm range (Karl Citation1980). Since the intensity of produced light is directly proportional to the ATP content (which is proportional to biomass), it is possible to quantify microbial biomass by measuring the ATP content using bioluminescence. The ATP method was used successfully for the rapid characterization of bioaerosol sampling devices when collecting bacterial aerosols in various environments (Seshadri et al. Citation2009; Park et al. Citation2014).

4.4. Immunoassays

Immunoassays are based on the specific reaction between antigens and antibodies. Monoclonal or polyclonal antibodies can be employed to detect the collected bioaerosols by binding to antigens, which are usually the proteins or polysaccharides, on the cell surface of the organisms. Many immunoassays are commercially available, allowing the rapid detection and identification of the microbes regardless of whether they are culturable. Some of the more widely used formats are detailed below (Jensen and Schafer Citation1998):

4.4.1. Radioimmunoassay (RIA)

The assay is based on the competitive binding of the unlabeled antigen (in the collected bioaerosol) to the specific antibody in the presence of the quantitatively added radiolabeled antigen. Higher concentrations of unlabeled antigen in the sample will result in lower levels of radiolabeled antigen-antibody complexes. A calibration curve created with known concentrations of the unlabeled antigen enables the assessment of the amount of an unknown, unlabeled antigen present in a sample (Garvey, Cremer, and Sussdorf Citation1977; Douwes et al. Citation2003).

4.4.2. Fluorescent immunoassay (FIA)

Fluorescent immunoassays utilize fluorescent-labeled antibodies to detect bacterial antigens either by directly detecting the antigen (cell-bound) using fluorescent antibody or by indirectly using antibody and fluorescent anti-gamma globulin antibody. In indirect FIA, serum antibody is detected using antigen, serum, and fluorescent antibody. A fluorescent microscope is used to evaluate the samples and to count the number of fluorescent viruses and microorganisms (Garvey, Cremer, and Sussdorf Citation1977; Hemmillä Citation1985).

4.4.3. Enzyme immunoassay (EIA)

In enzyme immunoassays, the antibody is coupled with an enzyme, such as horseradish peroxidase (HRP) or alkaline phosphatase (AP), which, in the presence of a chromogen, produces a colored end-product that can be quantitated by spectrophotometry. Most commercially available EIAs are enzyme-linked immunosorbent assays (ELISAs) where the antibody is coated onto the surfaces of test tubes, or wells of a microtiter plate, enabling rapid detection and speciation of the collected bioaerosols (Douwes et al. Citation2003).

4.4.4. Enzyme-Based halogen immunoassay (HIA)

The halogen immunoassay relies on the detection of airborne particles by capturing them onto protein-binding membranes by volumetric air sampling (Tovey et al. Citation2000). Eluted antigens are either detected with monoclonal antibodies (mAbs) to identify specific antigens of interest (environmental monitoring) or immunostained with human serum IgE to identify the patient’s sensitization patterns (serological monitoring) (Green, Millecchia, et al. Citation2006).

4.5. Molecular methods

Molecular methods have the potential to quantify exposure to airborne microbes independently of culturability and with high specificity. DNA extraction includes lysis to separate microbial cell contents, followed by precipitation and purification (Gorny et al. Citation2018). Rittenour et al. (Citation2012) compared different types of DNA extraction kits for DNA extraction starting with a lysis buffer and using a supernatant mix with a binding buffer to wash the DNA precipitate. A vortex step and a heating step are recommended to reduce the amount of DNA that will bind to filters. These steps have led to a 10-fold increase in DNA yield, allowing for more analysis of air samples (Dommergue et al. Citation2019).

The molecular level identification methods include PCR, mass spectrometry and sequencing. PCR analysis detects and amplifies a target segment of DNA (Stewart and Dowhanick Citation1996). Mass spectrometry measures the protein mass of isolated cultures and compares the results to a database of species’ profiles. Mass spectrometric analysis of bioaerosols has been described in another review in this special issue (Huffman et al. Citation2019). Sequencing is the most expensive route of identification (Rodrigues et al. Citation2017). Sequencing technologies use universal 16S rRNA to encode the genes and match the bacterial DNA sequence to a known profile. After PCR amplification the pure product of the 16S gene is sequenced, and compared to bacterial DNA segments in a database to identify the bacteria (Barghouthi Citation2011). Analysis of the fungal community in a complex sample is performed by sequencing the internal transcribed spacer (ITS) in the fungal rRNA operon (McTaggart et al. Citation2019). Species-specific primers can be used to identify the collected microorganisms (Novoa Rama et al. Citation2018).

Nebulization and sampling of aerosols may cause stress on the microorganisms, hindering the direct representation of the biodiversity at the sample site. Finding ways to represent biodiversity and analyze microbes from larger air volumes will provide a more accurate insight into what humans, animals, and other living organisms are exposed to (Duquenne Citation2018; Adhikari et al. Citation2017). Some DNA extraction methods can create PCR inhibitors and it is crucial to uncover these inhibitors for future research (Ferguson et al. Citation2019).

In a study involving students with influenza-like symptoms, cough-generated aerosols were collected using a NIOSH two-stage bioaerosol cyclone sampler or an SKC BioSampler. Quantitative real-time reverse-transcription PCR (qPCR) analysis of the viral RNA in the samples indicates that the cough-aerosols contain influenza virus and viral RNA within particles in the respirable size range (Lindsley et al. Citation2010). The number of viral aerosol studies in swine facility surveillance has significantly increased in recent years (Anderson et al. Citation2017). A study by Lauterbach et al. (Citation2018) on the molecular detection and recovery of influenza A virus from both the air and surfaces collected at agricultural fairs provides evidence for potential intra- and inter-species viral transmission through these routes.

4.5.1. Polymerase chain reaction (PCR)

4.5.1.1. Conventional PCR

Conventional polymerase chain reaction (PCR) is a system performed for DNA amplification, allowing the detection of specific nucleic acid sequences of DNA found in both culturable and non-culturable microorganisms. According to Alvarez, Buttner, and Stetzenbach (Citation1995) solid-phase PCR amplification detects airborne bacterial cells when a traditional culture cannot. Therefore, conventional PCR can be used as an alternative method for total bacterial count assessment and has been applied to analyze air samples for endemic microorganisms such as airborne mycobacteria and fungi (Li, Gao, and Liu Citation2011). During PCR amplification the desired gene or part of a gene in the collected bioaerosol is detected by amplifying the relevant fragment in a DNA template using a thermostable Taq polymerase enzyme and specific DNA primers that will recognize and bind only to the target DNA sequence (Alvarez, Buttner, and Stetzenbach Citation1995).

DNA sample preparation used for PCR amplification should be optimized for the successful amplification of increasingly longer targets from genomic DNA (Cheng et al. Citation1995). Therefore, before the PCR process, the intactness of the genomic DNA should be maintained during the collection and isolation process to create reproducible amplification of fragments with sizes >1.3 kb (Deagle, Eveson, and Jarman Citation2006). A larger target size can increase the probability of producing an unusable DNA template strand by randomly introducing a single-stranded (ss) nick or double-stranded (ds) break within the target sequence (Zhou, Pape, and Schwartz Citation2008). The DNA sample preparation varies for each experiment. Genomic DNA is isolated from the sample’s cultured cells by various methods such as boiling in the presence of a chelating resin, using alkaline lysis, or phenol extraction (Cheng et al. Citation1995). A standard reaction mixture contains the DNA template along with the specified forward and reverse primers, deoxyribonucleotide triphosphate (dNTP) mix, buffer with Mg++ for the specific DNA polymerase, and DNA polymerase. The PCR process typically utilizes thermal cycling, which consists of an initial denaturing step and a series of 20 to 40 cycles of denaturing, annealing and extension steps (Svabenska Citation2012; Liang and Johnson Citation1988; Sambrook, Fritsch, and Maniatis Citation1989).

The product of the PCR process is usually analyzed with gel electrophoresis to verify if the single product of the correct size was formed. Recently nano-materials (metallic and non-metallic) have been introduced to enhance PCR amplification by effectively reducing inhibition and increasing detection efficiencies (Xu and Yao Citation2013). Although colony counting is not as efficient as PCR, a combination of colony counts and conventional PCR can provide more detailed information when monitoring and verifying the presence of bioaerosols (Li, Gao, and Liu Citation2011).

PCR techniques are highly sensitive to bioaerosol detection and identification, making it independent of culturing. The advantage of using the PCR method is that it can be applied to most biological matter with nucleic acids, such as fungi, viruses, bacteria, and various allergens (Peccia and Hernandez Citation2006).

A disadvantage of using PCR is that air samples commonly contain compounds that could inhibit the amplification assay, therefore environmental interferences should be assessed before utilizing PCR for field screening (Alvarez, Buttner, and Stetzenbach Citation1995). Additional steps may be required to remove or inactivate PCR inhibitors (Chen et al. Citation2010; Schrader et al. Citation2012). For example, when using an impactor for bioaerosol collection there is an increased likelihood for the presence of EDTA in the impactor substrate wash liquid (King and McFarland Citation2012b). High concentrations of EDTA inhibit PCR (Huggett et al. Citation2008), therefore a procedure should be used to reduce the EDTA content before the PCR process.

4.5.1.2. Real-time (quantitative) PCR

Real-time quantitative polymerase chain reaction (qPCR) is a widely used method for the detection, quantification, and typing of different microbial agents (He and Yao Citation2011; Kralik and Ricchi Citation2017). It is also used to compare gene transcription levels (Saint-Marcoux et al. Citation2015). By simple definition, qPCR is the process of amplifying DNA, using fluorescent (SYBR® Green) dye to detect the dsDNA PCR product as it accumulates during PCR and intercalates with the dye. However, these dyes detect the accumulation of both specific and nonspecific PCR products. To improve the specificity of the detection, fluorogenic-labeled probes (TaqMan) have been introduced that use the 5’ nuclease activity of Taq DNA polymerase (Holland et al. Citation1991; Georgakopoulos et al. Citation2009). A probe containing a fluorescent reporter dye on the 5’ end and a quencher dye on the 3’ end anneals downstream from one of the primer sites and is cleaved by the 5’ nuclease activity of Taq DNA polymerase as this primer is extended. The cleavage of the probe separates the reporter dye from the quencher dye, increasing the fluorescence emitted by the reporter dye. Due to its simpler process and lower cost, SYBR® Green assays are more applied (Libert et al. Citation2015). The steps of qPCR are the same as for conventional PCR, including DNA denaturation, primer annealing, and extension (Kralik and Ricchi Citation2017).

Although PCR is a common method for amplifying DNA, it is also applicable for RNA templates by using reverse transcriptase to synthesize complementary DNA (cDNA) (Saint-Marcoux et al. Citation2015). Reverse transcription is performed by adding a primer or random hexamer primers and transcriptase before amplification (Lauterbach et al. Citation2018). DNA is less likely to degrade than RNA samples (Bustin et al. Citation2009). A small volume of the prepared cDNA sample can be used for further amplification (Saint-Marcoux et al. Citation2015).

Controls, positive and negative, and quantification calibrators are needed in all qPCR reactions. Positive controls are taken from the nucleic acids of the original sample to monitor assay variation and needed for the calibration curves (Bustin et al. Citation2009). Quantification calibrators need to be diluted and may act as the negative controls (Bustin et al. Citation2009). Multiple cycles, 20 and above, are applied to amplify the DNA properly. Amplicons containing 80 to 150 nucleotides in length can be used to reduce primer mismatch leading to low amplification efficiency and amplification bias (Saint-Marcoux et al. Citation2015). After performing the amplifications and comparing against amplicons, a calibration curve is prepared for final quantification either by using dilutions of the target genomic nucleic acid or plasmid standards (Kralik and Ricchi Citation2017). An alternative method for increased quantification is using whole-cell PCR, where bacterial cells followed by a 15 min heat lysis are added to the reaction as a DNA template (Hospodsky, Yamamoto, and Peccia Citation2010). Fresh mid-log phase bacteria can be used for the calibration curve (King and McFarland Citation2012b).

The four factors to consider in assessing the performance of a PCR are the limit of detection, the limit of quantification, precision, and PCR efficiency. Limit of detection, or LOD, is the lowest amount of analyte that can be detected in a 95% confidence, but not the exact value (Kralik and Ricchi Citation2017; Bustin et al. Citation2009; Forootan et al. Citation2017). This value can be estimated from the analysis of replicate standard curve[s] (Forootan et al. Citation2017). No more than 5% of the positive control samples can fail for the PCR reaction to be considerable. Limit of quantification, or LOQ, similar to LOD, can be estimated from the standard curve (Forootan et al. Citation2017). LOQ is the smallest amount of analyte which can be measured and quantified with defined precision and accuracy, with a value proposed to be fixed under 25% (Pavšic, Žel, and Milavec Citation2016).

The precision of qPCR can be affected by temperature differences affecting the completion of annealing and denaturation, concentration differences introduced by pipetting errors, and stochastic variation (Bustin et al. Citation2009). The repeatability of the qPCR is dependent on intra- and inter-assay variation describing the variability of replicates in the same experiment and the variability of different experiments conducted on different days (Kralik and Ricchi Citation2017). The amplification efficiency is determined by the slope of the log-linear portion of the calibration curve (Bustin et al. Citation2009).

After completing the qPCR amplification cycles, a melting curve can be generated for the resulting amplicon presenting a negative first derivative plot to indicate the loss in fluorescence over a temperature range of 55–92 °C. The desired amplicon will appear as a peak with a specific melting temperature (Tm). Potential contamination due to nonspecific binding, or the presence of primer dimers will have lower Tm (An, Mainelis, and White Citation2006).

Normalization of results from qPCR corrects the biological and experimental variability (Saint-Marcoux et al. Citation2015). In addition, it allows for comparisons of data from different RNA concentrations (Bustin et al. Citation2009). Normalization could be done by comparing transcript levels of the gene of interest to those of a reference gene similar to that being tested (Bustin et al. Citation2009). Qualitative assays of the qPCR data should also be done to verify the validity of the results (Bustin et al. Citation2009).

4.5.1.3. Viability PCR

Because of the amplification capabilities of PCR reactions, it is possible to design qPCR assays for every microorganism (Kralik and Ricchi Citation2017). The main advantages of this method are its fast and high-throughput detection and quantification of target DNA in different sequences. In addition, qPCR avoids cross-contamination, has a wide dynamic range for the quantification and allows for multiplexing which is essential for detection and quantification (Kralik and Ricchi Citation2017). However, the biggest disadvantage of qPCR is the inability to tell the difference between live and dead cells (Kralik and Ricchi Citation2017), leading to the development of viability qPCR (Nocker and Camper Citation2009). The application of a simple pretreatment of the sample by using specific intercalating photo-reactive reagents enables the neutralization of the DNA of dead cells. As a result, only DNA from live cells will be detected by PCR. Propidium monoazide (PMA) has been used as an effective intercalating reagent (Nocker, Cheung, and Camper Citation2006). Although the efficiency of photoactivation has been greatly improved, the viability PCR method is still under development.

4.5.1.4. Digital PCR

Digital PCR (dPCR) carries out a single reaction within the sample, however, it is separated into a large number of partitions (droplets) and the reaction is carried out in each partition individually, allowing a more reliable collection and sensitive, digital measurement of nucleic acid concentration. By measuring the number of droplets that are fluorescing due to amplification, dPCR provides absolute quantification while qPCR relies on the standard curve. The method has been routinely used for clonal amplification of samples for next-generation sequencing. Damit (Citation2017) developed a droplet microfluidics-based bioaerosol detector and analyzed bioaerosols impinged directly into the droplet.

4.5.1.5. Convective PCR

PCR by thermal convection is an alternative to other PCR processes. Convective PCR is shown to be a faster, near-instantaneous method for amplification (Wheeler et al. Citation2011). In this method, the bottom surface of the sample is heated, then cooled by the surrounding air creating a temperature gradient so PCR can be completed by natural convection (Chou et al. Citation2011). The thermal gradient through natural convection allows the sample to go through all three steps of PCR. That temperature gradient must be maintained throughout the experiment for it to be successful (Chou et al. Citation2011). In convective PCR, the temperature during annealing/extension and denaturation must be monitored and can be optimized depending on which gene pair is being amplified. If the temperature is not properly controlled, the process can fail (Zhang and Xing Citation2009). Overall, PCR by natural convective heat transfer is a simpler, and faster method of amplification.

The sample for the convective PCR is prepared similarly to the conventional PCR (Muddu, Hassan, and Ugaz Citation2011, Zhang and Xing Citation2009). The thermocycler, which facilitates convective PCR, had set conditions similar to conventional or real-time PCR technologies (Chou et al. Citation2011) and may be connected to impingement based continuous bioaerosol collectors via microfluidics. The convective PCR product is separated by agarose electrophoresis (Muddu, Hassan, and Ugaz Citation2011; Chou et al. Citation2011). The recently developed capillary version is able to monitor the fluorescence in situ with a smartphone camera, enabling real-time amplification (Qiu et al. Citation2019). In some cases, additional analysis may be done to validate the amplification, such as melting curve analysis (Chou et al. Citation2011).

4.5.2. Microbiome analysis

More detailed identification of biodiversity and microbiome is possible with the advancement and combination of powerful tools. The most commonly used method to identify bacterial taxonomy is the 16S rRNA gene sequencing as this gene is present in most bacteria, its function has not changed over time, and it has a fairly long base pair of 1500 bp for analysis (Janda and Abbott Citation2007). This is advantageous in many ways, such as identifying the species of ambiguous isolates, but it also has its limitations. Some identified species strains are outdated and may not be accurate anymore, the sequencing resolution does not clearly distinguish between certain bacterial strains, and there are still some errors in public databases of sequences. Databases, including the open access GenBank (National Center for Biotechnology Information [NCBI]), the European Nucleotide Archive, the DNA Data Bank of Japan, and the UniProt, Protein Data Bank, Ensembl and InterPro are powerful tools that allow the search and alignment of nucleotide or protein sequences (Boratyn et al. Citation2013; Wheeler and Bhagwat Citation2007). Shotgun sequencing is another method of high-throughput sequencing that uses untargeted sequencing over a microbial genome to compile whole genomes. Although it has some limitations, such as high cost, lack of valid annotations, and genome biases, shotgun sequencing results in metagenomics that enables complex microorganisms to be identified and analyzed (Quince et al. Citation2017). Next-generation sequencing (NGS) platforms enable other types of microbiome analysis and identification, such as whole-genome sequencing, total RNA sequencing, and methylation sequencing (Illumina Citation2016). The NGS technologies enable the generation of millions of reads from aerosol samples to analyze the genomics and transcriptomics (for RNA analysis), and have been successfully used in several bioaerosol studies (Womack et al. Citation2015; Shin et al. Citation2015). Metagenomic approaches enable comprehensive determination of the diversity and metabolic potential of the collected airborne organisms.

Various computer programs and platforms are available for the analysis of genomic sequences. Typically, analyzing the 16S rRNA gene data for bacterial and archaeal communities starts by clustering sequences into operational taxonomic units (OTUs). A similar approach can be used for the processing of other marker genes, including 18S rRNA gene sequences for eukaryotic analyses or ITS sequences for fungal studies (Khot, Ko, and Fredricks Citation2009). An open-source software package called Mothur combines algorithms from previous tools to perform tasks such as aligning sequences, assigning sequences to molecular operational taxonomic units (OTUs), calculating α and β diversity in samples, and calculating pairwise distances with and without trimming sequences (Schloss et al. Citation2009, Edgar Citation2018).

Another software called the Quantitative Insights Into Microbial Ecology (QIIME) and QIIME2 can be used for sequencing 16S rRNA gene fragments and identifying their taxonomy classification (Lawley and Tannock Citation2016; Hall and Beiko Citation2018). However, it has been reported that amplicons from Illumina sequencing sometimes contain errors in quality filtering and constructing OTUs for clusters of sequencing reads that differ by less than a fixed dissimilarity threshold. This problem can be overcome by using another open-source R package called DADA2, which contains an algorithm that can correct these errors enabling amplicon sequence variants (ASVs) to be resolved to the level of single-nucleotide differences over the sequenced gene region (Callahan et al. Citation2016). Recent debates focus on phasing out the traditional OTUs of marker gene sequences and instead delineating microbial taxa using exact sequence variants (ESVs) (Callahan, McMurdie, and Holmes Citation2017; Glassman and Martiny Citation2018). This approach avoids clustering sequences and instead uses only unique, identical 16S rRNA sequences for downstream community analyses that could differ by only one base pair. Another error-prone process is the conversion of genome abundances to cell numbers that requires taxa-specific information about the number of genome copies in the cell (Bonk et al. Citation2018). However, this is a useful way to normalize NGS data (Dannemiller et al. Citation2014). A number of additional tools are available to ‘de-noise’ and identify ESVs including Deblur, oligotyping, and UNOISE2 (Amir et al. Citation2017). Recent studies have found that archaea and bacteria can have less or more than 10 genome copies per cell, making it hard to correct for ploidy (Soppa Citation2014).

5. Antimicrobial resistance

The use of antimicrobials in agriculture, human medicine, and veterinary medicine has become a driving force in the evolution of drug-resistant microorganisms (Blair et al. Citation2015). In the past antimicrobial agents have been overused to treat infections, this overuse has led to an increase of selected mutated bacteria that have resistant genes (Karakonstantis and Kalemaki Citation2019; Schwarz, Loeffler, and Kadlec Citation2017).

Over 100,000 people die every year in US hospitals due to untreatable infections. Transmissible airborne pathogens are a possible cause for patient readmittance and are difficult to control due to air movement in the facility. When confined to the hospital setting, community-associated methicillin-resistant S. aureus (CA-MRSA) has now emerged as an important pathogen in the medical community (Le et al. Citation2017). The main drivers of antibiotic resistance include the overuse of antibiotics, in humans and animals, and the advantage of over three billion years of existence. Microorganisms use different methods to resist antimicrobials such as through mutations or through acquiring outside DNA by horizontal gene transfer. When an antibiotic is used, the susceptible bacteria will die off creating a concentration of growing colonies of resistant species (Munita and Arias Citation2016). Microorganisms can use chemical or genotypic changes, a decrease in membrane penetration, an increase in efflux of drugs, and changes in target sites as mechanisms for drug resistance (Blair et al. Citation2015; Knopp and Andersson Citation2018). Streptococcus pneumoniae, a multidrug-resistant bacterium (Nguyen et al. Citation2019) is the leading cause of lower respiratory infection morbidity and mortality globally (Troeger et al. Citation2018). Multidrug-resistant strains that cause nosocomial infections can be highly concentrated, with Acinetobacter baumannii and Staphylococci; some causes of infections included mechanical ventilation and time (Li, Sun, et al. Citation2018; Kim, Kim, and Kim Citation2010).

Antibiotics, metals and pollutants may be the main drives for antimicrobial resistance (AR) (Warnes, Highmore, and Keevil Citation2012; Antonova and Hammer Citation2011; Hastings, Rosenberg, and Slack Citation2004), however, other environmental stressors may contribute to the development of resistance by a mechanism that is not well studied. The high amount of antibiotics released into the wastewater may promote the selection of antimicrobial-resistant bacteria which find their way into natural environments. Numerous studies focus on emerging microbial pathogens with multidrug resistance in municipal wastewater treatment plants (Li et al. Citation2016; Barancheshme and Munir Citation2018). From the bioaerosol and liquid samples collected at the plants, qPCR assays identified 44 antibiotic-resistant genes (ARGs) that confer resistance to a wide range of antibiotics (Gaviria-Figueroa et al. Citation2019). A comparison of the ARG profiles across samples showed that the downwind bioaerosol profile was 68% similar to the profile found in liquid sludge samples. Animal operations are another source of ARGs (Zhou et al. Citation2016), that were detected both within and 150 m downwind from swine production facilities (Gibbs et al. Citation2004; Gibbs et al. Citation2006), posing a threat to workers and residents nearby. Arfken, Song, and Sung (Citation2015) found that fecal spray fields could be a source of antibiotic-resistant bacteria into the environment.

The widely applied Kirby Bauer disk diffusion method (Bauer et al. Citation1966) uses Mueller-Hinton agar (Oxoid, Hampshire, England) to analyze antibiotic resistance. Article discs impregnated with different antimicrobials are placed on inoculated plates and incubated overnight at 37 °C to assess the sensitive, intermediate and resistant cells based on the clearance of growth of the bacterial lawn around the discs (Bacto-Dickinson, England). Resistance data are interpreted according to National Committee for Clinical Laboratory Standards (NCCLS). For quantitative data, ETEST strips, which are calibrated with drug gradients, can be used to determine the Minimum Inhibitory Concentration (MIC) of the antibiotics (BioMerieux, Durham, NC, USA).

For the quantification of the genomic copy number (GCN) RT-PCR can be used. Stress from aerosol formation could lead to genome rearrangements or favor mutations that enhance antibiotic resistance. Furthermore, the selection of ARGs conferring resistance to β-lactams and aminoglycosides are often reported to be present on mobile genetic elements that are able to move within or between DNA strands, which could promote the acquisition and spread of resistance genes (Devarajan et al. Citation2015; Partridge et al. Citation2018). To identify molecular signatures for elevated antibiotic resistance, target genes that regulate multi-drug resistance functions can be analyzed from the collected aerosols (Griffith et al. Citation2018). Specific target genes may include marR and rfaC, which acquire two of the most prevalent alleles that affect the regulation of several genes involved in antibiotic resistance (Weiss et al. Citation2016). Alternate target genes include katE, sodC, and mdtM, which also have mutant alleles associated with enhanced resistance (Weiss et al. Citation2016). For each aerosol sample, DNA extraction can be performed with three replicate samples using standard methods (Zhou, Yang, and Jong Citation1990). Following isolation and purification, DNA from the replicates and time points for each aerosol sample can be prepared as barcoded libraries and sequenced on an Illumina HiSeq 2000. The sequence outputs can be analyzed using QIIME software to determine the allele distributions among the bacteria before and after aerosolization (Paulson et al. Citation2013). As an alternative approach, apparent resistant bacteria can be isolated from the Kirby-Bauer test plates and the presence of resistance alleles can be determined using PCR amplification of the target genes followed by standard Sanger sequencing.

6. Limitations, problems and perspectives

6.1. Bioaerosols

Most bioaerosol samples are a mixture of living, damaged, dead cells and inerts. Depending on environmental factors and collector devices, the ratio of living to the dead can be highly variable (). In non-viable bioaerosols with damaged cell membrane, the DNA is leaking out, unfolding and becoming hydrated and oxidized. However, in viable bioaerosols, the DNA often remains inside the cell, in a compact, folded form. The bioaerosols can also be non-viable, albeit with intact cell membranes; and viable, when the organism is metabolically active, but not culturable. As viability enhances a wide range of analysis, the airborne microorganisms should remain viable and culturable, maintaining intact cell wall and DNA integrity during collection. Preserving viability during collection may improve sensitivity of detection and identification assays and provide more information about agents of interest or concern. The increasingly wide-spread use of DNA-based molecular techniques requires the maintenance of intact cell wall to protect the integrity of the nucleic acids in the collected bioaerosols.

6.2. Bioaerosol detection

There is an increasing demand in the food processing industry for portable collector units with analytical capabilities to enumerate and identify the bioaerosol particles on-site, in real-time. Currently, the analysis of the samples requires a dedicated laboratory with trained personnel. Another challenge is the in-line, on-site collection and detection of nanosize bioaerosols. Viral epidemics including the recent Ebola infections necessitate the efficient sampling of large air volumes for the timely detection of a potential outbreak. However, a collector system coupled with a detector via microfluidics for the analysis of bioaerosols, even less for nanosize particles in room-size spaces is not readily available.

Acknowledgments

The authors acknowledge organizers of the special issue “Bioaerosol Research: Methods, Challenges, and Perspectives,” including Shanna Ratnesar-Shumate and Alex Huffman, based on requests from the Bioaerosol Working Group and after discussion during the Bioaerosol Standardization Workshop at the International Aerosol Conference in St Louis, Missouri in September 2018.

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

References

  • Acuña-Villaorduña, C., L. G. Schmidt-Castellani, P. Marques-Rodrigues, L. F. White, D. J. Hadad, M. Gaeddert, J. J. Ellner, K. P. Fennelly, M. Palaci, R. Dietze, et al. 2018. Cough-aerosol cultures of Mycobacterium tuberculosis in the prediction of outcomes after exposure. A household contact study in Brazil. PLoS One 13 (10):e0206384. doi:10.1371/journal.pone.0206384.
  • Adhikari, A., J. Jung, T. Reponen, J. S. Lewis, E. C. DeGrasse, L. F. Grimsley, G. L. Chew, and S. A. Grinshpun. 2009. Aerosolization of fungi, 1 → 3-[beta]-d glucan, and endotoxin from flood-affected materials collected in New Orleans homes. Environ. Res. 109 (3):215–24. doi:10.1016/j.envres.2008.12.010.
  • Adhikari, A., S. Kurella, P. Banerjee, and A. Mitra. 2017. Aerosolized bacteria and microbial activity in dental clinics during cleaning procedures. J. Aerosol Sci. 114:209–18. doi:10.1016/j.jaerosci.2017.09.019.
  • Agranovski, I. E., V. Agranovski, T. Reponen, K. Willeke, and S. A. Grinshpun. 2002. Development and evaluation of a new personal sampler for culturable airborne microorganisms. Atmos. Environ. 36 (5):889–98. doi:10.1016/S1352-2310(01)00488-5.
  • Aketagawa, J., S. Tanaka, H. Tamura, Y. Shibata, and H. Sait. 1993. Activation of limulus coagulation factor G by several 1→3-β-D-glucans: Comparison of the potency of glucans with identical degree of polymerization but different conformations. J. Biochem. 113 (6):683–6. doi:10.1093/oxfordjournals.jbchem.a124103.
  • Allen, R. J., and B. Waclaw. 2019. Bacterial growth: A statistical physicist’s guide. Reports on progress in physics. Phys. Soc. Great Br. 82 (1):016601. doi:10.1088/1361-6633/aae54.
  • Alvarez, A. J., M. P. Buttner, and L. D. Stetzenbach. 1995. PCR for bioaerosol monitoring: Sensitivity and environmental interference. Appl. Environ. Microbiol. 61 (10):3639–44. doi:10.1128/AEM.61.10.3639-3644.1995.
  • Amir, A., D. McDonald, J. A. Navas-Molina, E. Kopylova, J. T. Morton, Z. Z. Xu, E. P. Kightley, L. R. Thompson, E. R. Hyde, A. Gonzalez, et al. 2017. Deblur rapidly resolves single-nucleotide community sequence patterns. mSystems 2 (2):e00191–16. doi:10.1128/mSystems.00191-16.
  • An, H. R., G. Mainelis, and L. White. 2006. Development and calibration of real-time PCR for quantification of airborne microorganisms in air samples. Atmos. Environ. 40 (40):7924–39. doi:10.1016/j.atmosenv.2006.07.020.
  • Andersen, A. A. 1958. New sampler for the collection, sizing and enumeration of viable airborne particles. J. Bacteriol. 76:357–75.
  • Anderson, B. D., J. A. Lednicky, M. Torremorell, and G. C. Gray. 2017. The use of bioaerosol sampling for airborne virus surveillance in swine production facilities: A mini review. Front. Vet. Sci. 4:121. doi:10.3389/fvets.2017.00121.
  • Antonova, E. S., and B. K. Hammer. 2011. Quorum-sensing autoinducer molecules produced by members of a multispecies biofilm promote horizontal gene transfer to Vibriocholerae. FEMS Microbiol. Lett. 322 (1):68–76. doi:10.1111/j.1574-6968.2011.02328.x.
  • Arfken, A. M., B. Song, and J.-S. Sung. 2015. Comparison of airborne bacterial communities from a hog farm and spray field. J. Microbiol. Biotechnol. 25 (5):709–17. doi:10.4014/jmb.1408.08005.
  • Atlas, R. M. 2005. Preparation of media. In Handbook of media for environmental microbiology, 4–5. 2nd ed. Boca Raton, FL: CRC Press.
  • Barancheshme, F., and M. Munir. 2018. Strategies to combat antibiotic resistance in the wastewater treatment plants. Front. Microbiol. 8:2603. doi:10.3389/fmicb.2017.02603.
  • Barghouthi, S. A. 2011. A universal method for the identification of bacteria based on general PCR primers. Indian J. Microbiol. 51 (4):430–44. doi:10.1007/s12088-011-0122-5.
  • Bauer, A. W., W. M. M. Kirby, J. C. Sherris, and M. Turck. 1966. Antibiotic susceptibility testing by standard single disc method. Am. J. Clin. Pathol. 45 (4_ts):493–6. doi:10.1093/ajcp/45.4_ts.493.
  • Bayraktar, I., D. Doganay, S. Coskun, C. Kaynak, G. Akca, and H. E. Unalan. 2019. 3D printed antibacterial silver nanowire/polylactide nanocomposites. Compos. Part B: Eng. 172:671–8. doi:10.1016/j.compositesb.2019.05.059.
  • Beck, J. M., V. B. Young, and G. B. Huffnagle. 2012. The microbiome of the lung. Transl. Res. 160 (4):258–66. doi:10.1016/j.trsl.2012.02.005.
  • Blair, J. M., M. A. Webber, A. J. Baylay, D. O. Ogbolu, and L. J. Piddock. 2015. Molecular mechanisms of antibiotic resistance. Nat. Rev. Microbiol. 13 (1):42–51. doi:10.1038/nrmicro3380.
  • Boissy, R. J., D. J. Romberger, W. A. Roughead, L. Weissenburger-Moser, J. A. Poole, and T. D. LeVan. 2014. Shotgun pyrosequencing metagenomic analyses of dusts from swine confinement and grain facilities. PLoS One 9 (4):e95578. doi:10.1371/journal.pone.0095578.
  • Bond, D. R., and J. B. Russel. 2000. Protonmotive force regulates the membrane conductance of Streptococcus bovis in a non-ohmic fashion. Microbiology 146 (3):687–94. doi:10.1099/00221287-146-3-687.
  • Bonk, F., D. Popp, H. Harms, and F. Centler. 2018. PCR-based quantification of taxa-specific abundances in microbial communities: Quantifying and avoiding common pitfalls. J. Microbiol. Methods 153:139–47. doi:10.1016/j.mimet.2018.09.015.
  • Bold, H. C., C. J. Alexopoulos, and T. Delevoryas. 1980. Morphology of plants and fungi. 4th ed. New York: Harper & Row.
  • Boratyn, G., C. Camacho, P. Cooper, G. Coulouris, A. Fong, N. Ma, T. Madden, W. Matten, S. McGinnis, Y. Merezhuk, et al. 2013. BLAST: A more efficient report with usability improvements. Nucleic Acids Res. 41 (W1):W29–W33. doi:10.1093/nar/gkt282.
  • Brodie, E. L., T. Z. DeSantis, D. C. Joyner, S. M. Baek, J. T. Larsen, G. L. Andersen, T. C. Hazen, P. M. Richardson, D. J. Herman, T. K. Tokunaga, et al. 2006. Application of a high-density oligonucleotide microarray approach to study bacterial population dynamics during uranium reduction and reoxidation. Appl. Environ. Microbiol. 72 (9):6288–98. doi:10.1128/AEM.00246-06.
  • Bustin, S. A., V. Benes, J. A. Garson, J. Hellemans, J. Hugget, M. Kubista, R. Mueller, T. Nolan, M. W. Pfaffl, G. L. Shipley, et al.. 2009. The MIE guidelines: Minimum information for publication of quantitative real-time PCR experiments. Clin. Chem. 55 (4):611–22. doi:10.1373/clinchem.2008.112797.
  • Byeon, J. H., C. W. Park, K. Y. Yoon, J. H. Park, and J. Hwang. 2008. Size distributions of total airborne particles and bioaerosols in a municipal composting facility. Bioresour. Technol. 99 (11):5150–4. doi:10.1016/j.biortech.2007.09.014.
  • Callahan, B., P. McMurdie, and S. Holmes. 2017. Exact sequence variants should replace operational taxonomic units in marker-gene data analysis. ISME J. 11 (12):2639–43. doi:10.1038/ismej.2017.119.
  • Callahan, B., P. McMurdie, M. Rosen, A. Han, A. Johnson, and S. Holmes. 2016. DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Methods 13 (7):581–3. doi:10.1038/nmeth.3869.
  • Carvalho, E., C. Sindt, A. Verdier, C. Galan, L. O’Donoghue, S. Parks, and M. Thibaudon. 2008. Performance of the Coriolis air sampler, a high-volume aerosol-collection system for quantification of airborne spores and pollen grains. Aerobiologia 24 (4):191–201. doi:10.1007/s10453-008-9098-y.
  • Chang, C.-W., Y. H. Hwang, S. A. Grinshpun, J. M. Macher, and K. Willeke. 1994. Evaluation of counting error due to colony masking in bioaerosol sampling. Appl. Environ. Microbiol. 60 (10):3732–8. doi:10.1128/AEM.60.10.3732-3738.1994.
  • Chang, C.-W., and L.-J. Wang. 2014. Methods for quantifying Staphylococcus aureus in indoor air. Indoor Air 25 (1):59–67. doi:10.1111/ina.12124.
  • Chang, C.-W., and L.-J. Wang. 2015. Impact of culture media and sampling methods on Staphylococcus aureus aerosols. Indoor Air 25 (5):488–98. doi:10.1111/ina.12162.
  • Chang, R. Y., J. Wong, A. Mathai, S. Morales, E. Kutter, W. Britton, J. Li, and H.-K. Chan. 2017. Production of highly stable spray dried phage formulations for treatment of Pseudomonas aeruginosa lung infection. Eur. J. Pharm. Biopharm. 121:1–13. doi:10.1016/j.ejpb.2017.09.002.
  • Chen, B. T., G. A. Feather, A. Maynard, and C. Y. Rao. 2004. Development of a personal sampler for collecting fungal spores. Aerosol Sci. Technol. 38 (9):926–37. doi:10.1080/027868290511218.
  • Chen, P. S., F. T. Tsai, C. K. Lin, C. Y. Yang, C. C. Chan, C. Y. Young, and C. H. Lee. 2010. Ambient influenza and avian influenza virus during dust storm days and background days. Environ. Health Perspect. 118 (9):1211–6. doi:10.1289/ehp.0901782.
  • Chen, Q., and L. M. Hildemann. 2009. The effects of human activities on exposure to particulate matter and bioaerosols in residential homes. Environ. Sci. Technol. 43 (13):4641–6. doi:10.1021/es802296j.
  • Cheng, S., Y. Chen, J. A. Monforte, R. Higuchi, and B. Van Houten. 1995. Template integrity is essential for PCR amplification of 20- to 30-kb sequences from genomic DNA. PCR Methods Appl. 4 (5):294–8. doi:10.1101/gr.4.5.294.
  • Chou, W. P., P. H. Chen, M. Miao, Jr., L. S. Kuo, S. H. Yeh, and P. J. Chen. 2011. Rapid DNA amplification in a capillary tube by natural convection with a single isothermal heater. BioTechniques 50 (1):52–7. doi:10.2144/000113589.
  • Chuang, C., S. Yang, M. Chang, H. Huang, C. Luo, P. Hung, and W. Fang. 2013. Inactivation efficiency to Bacillus subtilis and Escherichia coli bacterial aerosols of spraying neutral electrolyzed water. J. Air Waste Manage. Assoc. 63 (12):1447–56. doi:10.1080/10962247.2013.827604.
  • Chun, D. T., K. Bartlett, T. Gordon, R. R. Jacobs, B. Larsson, L. Larsson, D. M. Lewis, J. Liesivuori, O. Michel, D. K. Milton, et al.. 2006. History and results of the two inter‐laboratory round robin endotoxin assay studies on cotton dust. Am. J. Ind. Med. 49 (4):301–6. doi:10.1002/ajim.20266.
  • Coico, R. 2005. Gram staining. Curr. Protoc. Immunol. A.3C.1–A.3C.2. doi:10.1002/9780471729259.mca03cs00.
  • Couch, R. B., T. R. Cate, R. G. Douglas, P. J. Gerone, and V. Knight. 1966. Effect of route of inoculation on experimental respiratory viral disease in volunteers and evidence for airborne transmission. Bacteriol. Rev. 30 (3):517–29. doi:10.1128/MMBR.30.3.517-529.1966.
  • Cox, C. S., and C. M. Wathes. 1995. Bioaerosols handbook. Boca Raton: CRC Press.
  • Cullinan, P., J. M. Harris, A. N. Taylor, A. M. Hole, M. Jones, F. Barnes, and G. Jolliffe. 2000. An outbreak of asthma in a modern detergent factory. Lancet 356 (9245):1899–900. doi:10.1016/S0140-6736(00)03264-5.
  • Cyprowski, M., A. Stobnicka-Kupiec, A. Ławniczek-Wałczyk, A. Bakal-Kijek, M. Gołofit-Szymczak, and R. L. Górny. 2018. Anaerobic bacteria in wastewater treatment plant. Int. Arch. Occup. Environ. Health 91 (5):571–9. doi:10.1007/s00420-018-1307-6.
  • Damit, B. 2017. Droplet-based microfluidics detector for bioaerosol detection. Aerosol Sci. Technol. 51 (4):488–500. doi:10.1080/02786826.2016.1275515.
  • Dannemiller, K. C., N. Lang-Yona, N. Yamamoto, Y. Rudich, and J. Peccia. 2014. Combining real-time pcr and next-generation DNA sequencing to provide quantitative comparisons of fungal aerosol populations. Atmos. Environ. 84:113–21. doi:10.1016/j.atmosenv.2013.11.036.
  • Dawson, A. M., J. Bettgenhaeuser, M. Gardiner, P. Green, I. Hernández-Pinzón, A. Hubbard, and M. J. Moscou. 2015. The development of quick, robust, quantitative phenotypic assays for describing the host-nonhost landscape to stripe rust. Front. Plant Sci. 6:876. doi:10.3389/fpls.2015.00876.
  • Deagle, B. E., J. P. Eveson, and S. N. Jarman. 2006. Quantification of damage in DNA recovered from highly degraded samples-a case study on DNA in faeces. Front. Zool. 3:11. doi:10.1186/1742-9994-3-11.
  • Dedesko, S., and J. A. Siegel. 2015. Moisture parameters and fungal communities associated with gypsum drywall in buildings. Microbiome 3 (1):71. doi:10.1186/s40168-015-0137-y.
  • Devarajan, N., A. Laffite, N. D. Graham, M. Meijer, K. Prabakar, J. I. Mubedi, V. Elongo, P. T. Mpiana, B. W. Ibelings, W. Wildi, et al.. 2015. Accumulation of clinically relevant antibiotic-resistance genes, bacterial load, and metals in freshwater lake sediments in Central Europe. Environ. Sci. Technol. 49 (11):6528–37. doi:10.1021/acs.est.5b01031.
  • Dommergue, A., P. Amato, R. Tignat-Perrier, O. Magand, A. Thollot, M. Joly, L. Bouvier, K. Sellegri, T. Vogel, J. E. Sonke, et al.. 2019. Methods to investigate the global atmospheric microbiome. Front. Microbiol. 10:1–12. doi:10.3389/fmicb.2019.00243.
  • Dong, S., and M. Yao. 2010. Exposure assessment in Beijing, China: Biological agents, ultrafine particles, and lead. Environ. Monit. Assess. 170 (1-4):331–43. doi:10.1007/s10661-009-1236-7.
  • Douwes, J., G. Doekes, R. Montijn, D. Heederik, and B. Brunekreef. 1996. Measurement of β1→3-glucans in the occupational and home environment with an inhibition enzyme immunoassay. Appl. Environ. Microbiol. 62 (9):3176–82. doi:10.1128/AEM.62.9.3176-3182.1996.
  • Douwes, J., P. Thorne, N. Pearce, and D. Heederik. 2003. Bioaerosol health effects and exposure assessment: Progress and prospects. Bull. WHO 47:187–200. doi:10.1093/annhyg/meg032.
  • Duan, D. 2006. From the smallest virus to the biggest gene: Marching towards gene therapy for duchenne muscular dystrophy. Discov Med. 6 (33):103–8.
  • Dubovi, E. J., and T. G. Akers. 1970. Airborne stability of tailless bacterial viruses S13 and MS2. Appl. Microbiol. 19 (4):624–8. doi:10.1128/AEM.19.4.624-628.1970.
  • Duquenne, P. 2018. On the identification of culturable microorganisms for the assessment of biodiversity in bioaerosols. Ann. Work Expos. Health 622:139–46. doi:10.1093/annweh/wxx096.
  • Duquenne, P., G. Marchand, and C. Duchaine. 2013. Measurement of endotoxins in bioaerosols at workplace: A critical review of literature and a standardization issue. Ann. Occup. Hyg. 57 (2):137–72. doi:10.1093/annhyg/mes051.
  • Edgar, R. C. 2018. Updating the 97% identity threshold for 16S ribosomal RNA OTUs. Bioinformatics 34 (14):2371–5. doi:10.1093/bioinformatics/bty113.
  • Eduard, W., D. Heederik, C. Duchaine, and B. J. Green. 2012. Bioaerosol exposure assessment in the workplace: The past, present and recent advances. J. Environ. Monit. 14 (2):334–9. doi:10.1039/c2em10717a.
  • Estrada-Perez, C. E., K. A. Kinney, J. P. Maestre, Y. A. Hassan, and M. D. King. 2018. Droplet distribution and airborne bacteria in an experimental shower unit. Water Res. 130:47–57. doi:10.1016/j.watres.2017.11.039.
  • Ferguson, R. M., S. Garcia-Alcega, F. Coulon, A. J. Dumbrell, C. Whitby, and I. Colbeck. 2019. Bioaerosol biomonitoring: Sampling optimization for molecular microbial ecology. Mol. Ecol. Resour. 19 (3):672–90. doi:10.1111/1755-0998.13002.
  • Forootan, A., R. Sjöback, J. Björkman, B. Sjögreen, L. Linz, and M. Kubista. 2017. Methods to Determine limit of detection and limit of quantification in quantitative real-time PCR qPCR. Biomol. Detect. Quantif. 12:1–6. doi:10.1016/j.bdq.2017.04.001.
  • Frankel, M., G. Bekö, M. Timm, S. Gustavsen, E. W. Hansen, and A. M. Madsen. 2012. Seasonal variations of indoor microbial exposures and their relation to temperature, relative humidity, and air exchange rate. Appl. Environ. Microbiol. 78 (23):8289–97. doi:10.1128/AEM.02069-12.
  • Garre, A., J. A. Egea, A. Esnoz, A. Palop, and P. S. Fernandez. 2019. Tail or artefact? Illustration of the impact that uncertainty of the serial dilution and cell enumeration methods has on microbial inactivation. Food Res. Int. 119:76–83. doi:10.1016/j.foodres.2019.01.059.
  • Garvey, J. S., N. E. Cremer, and D. H. Sussdorf. 1977. Methods in immunology. 3rd ed., vol. 87, 301–12, 431–4. Reading, MA: W.A. Benjamin, Inc.
  • Gaviria-Figueroa, A., E. C. Preisner, S. Hoque, C. E. Feigley, and R. S. Norman. 2019. Emission and dispersal of antibiotic resistance genes through bioaerosols generated during the treatment of municipal sewage. Sci. Total Environ. 686:402–12. doi:10.1016/j.scitotenv.2019.05.454.
  • Georgakopoulos, D. G., V. Després, J. Fröhlich-Nowoisky, R. Psenner, P. A. Ariya, M. Pósfai, H. E. Ahern, B. F. Moffett, and T. C. J. Hill. 2009. Microbiology and atmospheric processes: Biological, physical and chemical characterization of aerosol particles. Biogeosciences 6 (4):721–37. doi:10.5194/bg-6-721-2009.
  • Ghosh, B., H. Lal, and A. Srivastava. 2015. Review of bioaerosols in indoor environment with special reference to sampling, analysis and control mechanisms. Environ. Int. 85:254–72. doi:10.1016/j.envint.2015.09.018.
  • Gibbs, S. G., C. F. Green, P. M. Tarwater, L. C. Mota, K. D. Mena, and P. V. Scarpino. 2006. Isolation of antibiotic-resistant bacteria from the air plume downwind of a swine confined or concentrated animal feeding operation. Environ. Health Perspect. 114 (7):1032–7. doi:10.1289/ehp.8910.
  • Gibbs, S. G., C. F. Green, P. M. Tarwater, and P. V. Scarpino. 2004. Airborne antibiotic resistant and nonresistant bacteria and fungi recovered from two swine herd confined animal feeding operations. J. Occup. Environ. Hyg. 1 (11):699–706. doi:10.1080/15459620490515824.
  • Glassman, S. I., and J. Martiny. 2018. Broadscale ecological patterns are robust to use of exact sequence variants versus operational taxonomic units. mSphere 3 (4):e00148–18. doi:10.1128/mSphere.00148-18.
  • Gorny, A. M., F. S. Hay, X. Wang, and S. J. Pethybridge. 2018. Isolation of nematode DNA from 100 g of soil using Fe3O4 super paramagnetic nanoparticles. Nematology 20 (3):271–83. doi:10.1163/15685411-00003140.
  • Górny, R. L., and A. Ławniczek-Wałczyk. 2012. Effect of two aerosolization methods on the release of fungal propagules from a contaminated agar surface. Ann. Agric. Environ. Med. 19:279–84.
  • Green, B. J., L. L. Millecchia, F. M. Blachere, E. R. Tovey, D. H. Beezhold, and D. Schmechel. 2006. Dual fluorescent halogen immunoassay for bioaerosols using confocal microscopy. Anal. Biochem. 354 (1):151–3. doi:10.1016/j.ab.2006.03.035.
  • Green, B. J., E. R. Tovey, J. K. Sercombe, F. M. Blachere, D. H. Beezhold, and D. Schmechel. 2006. Airborne fungal fragments and allergenicity. Med. Mycol. 44 (s1):245–S255. doi:10.1080/13693780600776308.
  • Griffith, E. C., M. J. Wallace, Y. Wu, G. Kumar, S. Gajewski, P. Jackson, G. A. Phelps, Z. Zheng, C. O. Rock, R. E. Lee, et al.. 2018. The structural and functional basis for recurring sulfa drug resistance mutations in Staphylococcus aureus dihydropteroate synthase. Front. Microbiol. 9:1369. doi:10.3389/fmicb.2018.01369.
  • Grinshpun, S. A., M. P. Buttner, and K. Willeke. 2007. Sampling for airborne microorganisms. In Manual of environmental microbiology, ed. C. J. Hurst, R. L. Crawford, J. L. Garland, D. A. Lipson, A. L. Mills, and L. D. Stetzenbach, 939–51. 3rd ed. Washington, DC: ASM Press.
  • Gutarowska, B., J. Skóra, Ł. Stępień, B. Szponar, A. Otlewska, and K. Pielech-Przybylska. 2015. Assessment of microbial contamination within working environments of different types of composting plants. J. Air Waste Manage. Assoc. 65 (4):466–78. doi:10.1080/10962247.2014.960954.
  • Hall, M., and R. Beiko. 2018. 16S rRNA gene analysis with QIIME2. In Microbiome analysis. Methods in molecular biology, ed. R. Beiko, W. Hsiao, and J. Parkinson, vol. 1849, 113–29. New York, NY: Humana Press.
  • Han, T., N. Thomas, and G. Mainelis. 2018. Performance of personal electrostatic bioaerosol sampler PEBS when collecting airborne microorganisms. J. Aerosol Sci. 124:54–67. doi:10.1016/j.jaerosci.2018.07.004.
  • Hastings, P. J., S. M. Rosenberg, and A. Slack. 2004. Antibiotic-induced lateral transfer of antibiotic resistance. Trends Microbiol. 12 (9):401–4. doi:10.1016/j.tim.2004.07.003.
  • He, Q., and M. Yao. 2011. Integration of high volume portable aerosol-to-hydrosol sampling and qPCR in monitoring bioaerosols. J. Environ. Monit. 13 (3):706–12. doi:10.1039/c0em00559b.
  • Heida, H., F. Bartman, and S. C. van der Zee. 1995. Occupational exposure and indoor air quality monitoring in composting facility. Am. Ind. Hyg. Assoc. J. 56 (1):39–43. doi:10.1080/15428119591017295.
  • Hemmillä, I. 1985. Fluoroimmunoassays and immunofluorometric assays. Clin. Chem. 31 (3):359–70.
  • Hinds, W. C. 1982. Aerosol Technology, vol. 78, 104–26, 165. New York, NY: John Wiley and Sons, Ltd.
  • Hoisington, A., J. P. Maestre, M. D. King, J. A. Siegel, and K. A. Kinney. 2014. Characterizing the indoor microbiome with pyrosequencing: Impact of sampler selection. Build. Environ. 80:274–9. doi:10.1016/j.buildenv.2014.04.021.
  • Holland, P. M., R. D. Abramson, R. Watson, and D. H. Gelfand. 1991. Detection of specific polymerase chain reaction product by utilizing the 5’-3’ exonuclease activity of Thermus aquaticus DNA polymerase. Proc. Natl. Acad. Sci. USA 88 (16):7276–80. doi:10.1073/pnas.88.16.7276.
  • Hospodsky, D., N. Yamamoto, and J. Peccia. 2010. Accuracy, precision, and method detection limits of quantitative PCR for airborne bacteria and fungi. Appl. Environ. Microbiol. 76 (21):7004–12. doi:10.1128/AEM.01240-10.
  • Hu, S., and A. R. McFarland. 2007. Numerical performance simulation of a wetted wall bioaerosol sampling cyclone. Aerosol Sci. Technol. 41 (2):160–8. doi:10.1080/02786820601124065.
  • Hua, Y. T., and R. C. Tong. 1992. Use of the polymerase chain reaction for specific detection of type A, D and E enterotoxigenic Staphylococcus aureus in foods. Appl. Microbiol. Biotechnol. 37:685–90.
  • Huang, K., Y. Tian, D. Salvi, M. Karwe, and N. Nitin. 2018. Influence of exposure time, shear stress, and surfactants on detachment of Escherichia coli O157:H7 from fresh lettuce leaf surfaces during washing process. Food Bioprocess Technol. 11 (3):621–33. doi:10.1007/s11947-017-2038-5.
  • Hubad, B., and A. Lapanje. 2013. The efficient method for simultaneous monitoring of the culturable as well as nonculturable airborne microorganisms. PLoS One 8 (12):e82186. doi:10.1371/journal.pone.0082186.
  • Hubbard, J. A., Haglund, J. S. Ezekoye, O. A. A. R. and McFarland. A. R. 2011. Liquid consumption of wetted wall bioaerosol sampling cyclones: Characterization and control. Aerosol Sci. Technol. 45 (2):172–82. doi:10.1080/02786826.2010.528806.
  • Huffman, J. A., A. E. Perring, N. J. Savage, B. Clot, B. Crouzy, F. Tummon, O. Shoshanim, B. Damit, J. Schneider, V. Sivaprakasam, et al. 2019. Real-time sensing of bioaerosols: Review and current perspectives. Aerosol Sci. Technol. 1–31. doi:10.1080/02786826.2019.1664724.
  • Huggett, J. F., T. Novak, J. A. Garson, C. Green, S. D. Morris-Jones, R. F. Miller, and A. Zumla. 2008. Differential susceptibility of PCR reactions to inhibitors: An important and unrecognized phenomenon. BMC Res. Notes 1 (1):70. doi:10.1186/1756-0500-1-70.
  • Illumina. 2016. An introduction to next-generation sequencing technology. https://www.illumina.com/science/technology/next-generation-sequencing.html
  • Janda, J., and S. Abbott. 2007. 16S rRNA gene sequencing for bacterial identification in the diagnostic laboratory: Pluses, perils, and pitfalls. J. Clin. Microbiol. 45 (9):2761–4. doi:10.1128/JCM.01228-07.
  • Jensen, P. A., and M. P. Schafer. 1998. Sampling and characterization of bioaerosols. NIOSH Man. Anal. Methods 2:82–112. doi:10.26616/nioshpub82100.
  • Jung, J. H., and J. E. Lee. 2016. Real-time bacterial microcolony counting using on-chip microscopy. Sci Rep. 6:21473. doi:10.1038/srep21473.
  • Jürgensen, C. W., and A. M. Madsen. 2016. Influence of everyday activities and presence of people in common indoor environments on exposure to airborne fungi. AIMS Environ. Sci 3 (1):77–95.
  • Kakde, U. 2012. Fungal bioaerosols: Global diversity, distribution and its impact on human beings and agricultural crops. Bionano Front. 5:323–9.
  • Kapoor, G., S. Saigal, and A. Elongavan. 2017. Action and resistance mechanisms of antibiotics: A guide for clinicians. J. Anaesthesiol. Clin. Pharmacol. 33 (3):300–30. doi:10.4103/joacp.JOACP_349_15.
  • Karakonstantis, S., and D. Kalemaki. 2019. Antimicrobial overuse and misuse in the community in Greece and link to antimicrobial resistance using methicillin-resistant S. aureus as an example. J. Infect. Public Health 12 (4):460–4. pii: S1876-03411930126-1. doi:10.1016/j.jiph.2019.03.017.
  • Karl, D. M. 1980. Cellular nucleotide measurements and applications in microbial ecology. Microbiol. Mol. Biol. Rev. 44 (4):739–96. doi:10.1128/MMBR.44.4.739-796.1980.
  • Kaur, P., A. Ghosh, R. V. Krishnamurthy, D. G. Bhattacharjee, V. Achar, S. Datta, S. Narayanan, A. Anbarasu, and S. Ramaiah. 2015. A high-throughput cidality screen for Mycobacterium tuberculosis. PLoS One 10 (2):e0117577. doi:10.1371/journal.pone.0117577.
  • Keswani, J., M. L. Kashon, and B. T. Chen. 2005. Evaluation of interference to conventional and real-time PCR for detection and quantification of fungi in dust. J. Environ. Monit. 7 (4):311–8. doi:10.1039/b415498c.
  • Khot, P. D., D. L. Ko, and D. N. Fredricks. 2009. Sequencing and analysis of fungal rRNA operons for development of broad-range fungal PCR assays. Appl. Environ. Microbiol. 75 (6):1559–65. doi:10.1128/AEM.02383-08.
  • Kim, K., E. Kabir, and S. Jahan. 2018. Airborne bioaerosols and their impact on human health. J. Environ. Sci. 67:23–35. 1016. doi:10.1016/j.jes.2017.08.027.
  • Kim, K., Y. Kim, and D. Kim. 2010. Distribution characteristics of airborne bacteria and fungi in the general hospitals of Korea. Ind. Health 48 (2):236–43. doi:10.2486/indhealth.48.236.
  • King, M. D., and A. R. McFarland. 2012a. Use of an Andersen bioaerosol sampler to simultaneously provide culturable particle and culturable organism size distributions. Aerosol Sci. Technol. 46 (8):852–61. doi:10.1080/02786826.2012.669507.
  • King, M. D., and A. R. McFarland. 2012b. Bioaerosol sampling with a wetted wall cyclone: Cell culturability and DNA integrity of Escherichia coli bacteria. Aerosol Sci. Technol. 46 (1):82–93. doi:10.1080/02786826.2011.605400.
  • King, M. D., B. F. Thien, J. S. Tiirikainen, and A. R. McFarland. 2009. Collection characteristics of a batch-type wetted wall bioaerosol sampling cyclone. Aerobiologia 25 (4):239–47. doi:10.1007/s10453-009-9129-3.
  • Knopp, M., and D. I. Andersson. 2018. Predictable phenotypes of antibiotic resistance mutations. Am. Soc. Microbiol. 9 (3):e00770-18. doi:10.1128/mBio.00770-18.
  • Kozajda, A., K. Jeżak, M. Cyprowski, and I. Szadkowska-Stańczyk. 2017. Inhalable dust, endotoxins and 1-3-β-D-glucans as indicators of exposure in waste sorting plant environment. Aerobiologia 33 (4):481–91. doi:10.1007/s10453-017-9484-4.
  • Kralik, P., and M. Ricchi. 2017. A basic guide to real time PCR in microbial diagnostics: Definitions, parameters, and everything. Front. Microbiol. 8:108. doi:10.3389/fmicb.2017.00108.
  • Kuehn, A., and C. Hilger. 2015. Animal allergens: Common protein characteristics featuring their allergenicity. Front. Immunol. 6:40. doi:10.3389/fimmu.2015.00040.
  • Lagier, J.-C., S. Edouard, I. Pagnier, O. Mediannikov, M. Drancourt, and D. Raoult. 2015. Current and past strategies for bacterial culture in clinical microbiology. Clin. Microbiol. Rev. 28 (1):208–36. doi:10.1128/CMR.00110-14.
  • Lauterbach, S., C. Wright, M. Zentkovich, S. Nelson, J. Lorbach, N. Bliss, J. Nolting, R. Pierson, M. King, and A. Bowman. 2018. Detection of influenza A virus from agricultural fair environment: Air and surfaces. Prev. Vet. Med. 153:24–9. doi:10.1016/j.prevetmed.2018.02.019.
  • Lawley, B., and G. Tannock. 2016. Analysis of 16S rRNA gene amplicon sequences using the QIIME software package. In Oral biology. Methods in molecular biology, ed. G. Seymour, M. Cullinan, and N. Heng, vol. 1537: 153–63. New York, NY: Humana Press.
  • Le, J., Q. Dam, T. Tran, A. Nguyen, F. C. Adler-Shohet, S. Kim, K. Schmidt, J. M. Lieberman, and J. S. Bradley. 2017. Epidemiology and hospital readmission associated with complications of Staphylococcus aureus bacteremia in pediatrics over a 25-year period. Epidemiol. Infect. 145 (12):2631–9. doi:10.1017/S0950268817001571.
  • Lee, J. H., G. B. Hwang, J. H. Jung, D. H. Lee, and B. U. Lee. 2010. Generation characteristics of fungal spore and fragment bioaerosols by airflow control over fungal cultures. J. Aerosol Sci. 41 (3):319–25. doi:10.1016/j.jaerosci.2009.11.002.
  • Lee, T., S. A. Grinshpun, D. Martuzevicius, A. Adhikari, C. M. Crawford, J. Luo, and T. Reponen. 2006. Relationship between indoor and outdoor bioaerosols collected with a button inhalable aerosol sampler in urban homes. Indoor Air 16 (1):37–47. doi:10.1111/j.1600-0668.2005.00396.x.
  • Leung, S. S., T. Parumasivam, F. G. Gao, N. B. Carrigy, R. Vehring, W. H. Finlay, S. Morales, W. J. Britton, E. Kutter, and H. K. Chan. 2016. Production of inhalation phage powders using spray freeze drying and spray drying techniques for treatment of respiratory infections. Pharm. Res. 33 (6):1486–96. doi:10.1007/s11095-016-1892-6.
  • Li, B., G. Sun, Z. Cheng, C. Mei, X. Liao, J. Li, and Y. Yuan. 2018. Analysis of nosocomial infections in post–cardiac surgery extracorporeal membrane oxygenation support therapy. HeartSurgForum 21 (5):387–91. doi:10.1532/hsf.1789.
  • Li, C. S., and Y. C. Lin. 1999. Sampling performance of impactors for bacterial bioaerosols. Aerosol Sci. Technol. 30 (3):280–7. doi:10.1080/027868299304633.
  • Li, J., L. Zhou, X. Zhang, C. Xu, L. Dong, and M. Yao. 2016. Bioaerosol emissions and detection of airborne antibiotic resistance genes from a wastewater treatment plant. Atmos. Environ. 124:404–12. doi:10.1016/j.atmosenv.2015.06.030.
  • Li, L., M. Gao, and J. Liu. 2011. Distribution characterization of microbial aerosols emitted from a wastewater treatment plant using the orbal oxidation ditch process. Proc. Biochem. 46 (4):910–5. doi:10.1016/j.procbio.2010.12.016.
  • Li, R., U. Goswami, M. King, J. Chen, T. C. Cesario, and P. M. Rentzepis. 2018. In situ detection of live-to-dead bacteria ratio after inactivation by means of synchronous fluorescence and PCA. Proc. Natl. Acad. Sci. USA. 115 (4):668–73. doi:10.1073/pnas.1716514115.
  • Liang, R., P. Xiao, R. She, S. Han, L. Chang, and L. Zheng. 2013. Culturable airborne bacteria in outdoor poultry-slaughtering facility. Microbes Environ. 28 (2):251–6. doi:10.1264/jsme2.ME12178.
  • Liang, W., and J. P. Johnson. 1988. Rapid plasmid insert amplification with polymerase chain reaction. Nucleic Acids Res. 16 (8):3579. doi:10.1093/nar/16.8.3579.
  • Liebers, V., H. Stubel, M. Düser, T. Brüning, and M. Raulf-Heimsoth. 2009. Standardization of whole blood assay for determination of pyrogenic activity in organic dust samples. Int. J. Hyg. Environ. Health. 212 (5):547–56. doi:10.1016/j.ijheh.2009.03.003.
  • Libert, X., C. Chasseur, S. Bladt, A. Packeu, F. Bureau, N. H. Roosens, and S. C. De Keersmaecker. 2015. Development and performance assessment of a qualitative SYBR® green real-time PCR assay for the detection of Aspergillus versicolor in indoor air. Appl. Microbiol. Biotechnol. 99 (17):7267–82. doi:10.1007/s00253-015-6785-9.
  • Lin, W.-H. 1999. Evaluation of impingement and filtration methods for yeast bioaerosol sampling. Aerosol Sci. Technol. 30 (2):119–26. doi:10.1080/027868299304723.
  • Lindsley, W. G., F. M. Blachere, R. E. Thewlis, A. Vishnu, K. A. Davis, G. Cao, J. E. Palmer, K. E. Clark, M. A. Fisher, R. Khakoo, et al. 2010. Measurements of airborne influenza virus in aerosol particles from human coughs. PLoS One 5 (11):e15100. doi:10.1371/journal.pone.0015100.
  • Lindsley, W. G., D. Schmechel, and B. T. Chen. 2006. A two-stage cyclone using microcentrifuge tubes for personal bioaerosol sampling. J. Environ. Monit. 8 (11):1136–42. doi:10.1039/b609083d.
  • Mainelis, G., K. Willeke, P. Baron, T. Reponen, S. A. Grinshpun, R. L. Gorny, and S. Trakumas. 2001. Electrical charges on airborne microorganisms. J. Aerosol Sci. 32 (9):1087–110. doi:10.1016/S0021-8502(01)00039-8.
  • Mandal, J., and H. Brandl. 2011. Bioaerosols in indoor environment-a review with special reference to residential and occupational locations. Open Environ. Biol. Monitor. J. 4:83–96. doi:10.2174/1875040001104010083.
  • McFarland, A. R., J. S. Haglund, M. D. King, S. Hu, M. S. Phull, B. W. Moncla, and Y. Seo. 2010. Wetted wall cyclones for bioaerosol sampling. Aerosol Sci. Technol. 44 (4):241–52. doi:10.1080/02786820903555552.
  • McTaggart, L. R., J. K. Copeland, A. Surendra, P. W. Wang, S. Husain, B. Coburn, D. S. Guttman, and J. V. Kus. 2019. Mycobiome sequencing and analysis applied to fungal community profiling of the lower respiratory tract during fungal pathogenesis. Front. Microbiol. 10:512. doi:10.3389/fmicb.2019.00512.
  • Michaud, J. M., L. R. Thompson, D. Kaul, J. L. Espinoza, R. A. Richter, Z. Z. Xu, C. Lee, K. M. Pham, C. M. Beall, F. Malfatti, et al. 2018. Taxon-specific aerosolization of bacteria and viruses in an experimental ocean-atmosphere mesocosm. Nat. Commun. 9 (1):2017. doi:10.1038/s41467-018-04409-z.
  • Milton, D. K., R. J. Gere, H. A. Feldman, and I. A. Greaves. 1990. Endotoxin measurement: Aerosol sampling and application of a new Limulus method. Am. Ind. Hyg. Assoc. J. 51 (6):331–7. doi:10.1080/15298669091369754.
  • Mirhoseini, S. H., M. Nikaeen, K. Satoh, and K. Makimura. 2016. Assessment of airborne particles in indoor environments: Applicability of particle counting for prediction of bioaerosol. Aerosol Air Qual. Res. 16 (8):1903–10. doi:10.4209/aaqr.2015.08.0528.
  • Missiakas, D. M., and O. Schneewind. 2013. Growth and laboratory maintenance of Staphylococcus aureus. Curr. Protoc. Microbiol. Chapter 9:Unit 9C.1. doi:10.1002/9780471729259.mc09c01s28.
  • Mitsuboshi, M., Y. Kioka, K. Noguchi, and S. Asakawa. 2016. An evaluation method for the suppression of pathogenic Fusarium oxysporum by soil microorganisms using the dilution plate technique. Microbes Environ. 31 (3):307–13. doi:10.1264/jsme2.ME16052.
  • Morris, K. J. 1995. Modern microscopic methods of bioaerosol analysis. In Bioaerosol handbook, ed. C. S. Cox and C. M. Wathes, 285–316. Boca Raton, LA: CRC Lewis Publishers.
  • Mousavi, B., M. T. Hedayati, N. Hedayati, M. Ilkit, and S. Syedmousavi. 2016. Aspergillus species in indoor environments and their possible occupational and public health hazards. Curr. Med. Mycol. 2 (1):36–42. doi:10.18869/acadpub.cmm.2.1.36.
  • Muddu, R., Y. A. Hassan, and V. M. Ugaz. 2011. Rapid PCR thermocycling using microscale thermal convection. J. Vis. Exp. 49:2366. doi:10.3791/2366.
  • Munita, J. M., and C. A. Arias. 2016. Mechanisms of antibiotic resistance. Microbiol. Spectr. 4 (2):1–29. doi:10.1128/microbiolspec.VMBF-0016-2015.
  • Ng, T. W., W. L. Chan, and K. M. Lai. 2017. Importance of stress-response genes to the survival of airborne Escherichia coli under different levels of relative humidity. AMB Expr. 7 (1):1–7. doi:10.1186/s13568-017-0376-3.
  • Nguyen, H. A., H. Fujii, H. T. Vu, C. M. Parry, A. D. Dang, K. Ariyoshi, and L. Yoshida. 2019. An alarmingly high nasal carriage rate of Streptococcus pneumoniae serotype 19F non-susceptible to multiple beta-lactam antimicrobials among Vietnamese children. BMC Infect. Dis. 19 (1):241. 201919:241 doi:10.1186/s12879-019-3861-2.
  • Nocker, A., and A. K. Camper. 2009. Novel approaches toward preferential detection of viable cells using nucleic acid amplifcation techniques. FEMS Microbiol. Lett. 291 (2):137–42. doi:10.1111/j.1574-6968.2008.01429.x.
  • Nocker, A., C.-Y. Cheung, and A. K. Camper. 2006. Comparison of propidium monoazide with ethidium monoazide for differentiation of live vs. dead bacteria by selective removal of DNA from dead cells. J. Microbiol. Methods 67 (2):310–20. doi:10.1016/j.mimet.2006.04.015.
  • Novoa Rama, E., M. Bailey, D. R. Jones, R. K. Gast, K. Anderson, J. Brar, R. Taylor, H. F. Oliver, and M. Singh. 2018. Prevalence, persistence, and antimicrobial resistance of Campylobacter spp. From eggs and laying hens housed in five commercial housing systems. Foodborne Pathog. Dis. 15 (8):506–16. doi:10.1089/fpd.2017.2404.
  • Pabary, R., C. Singh, S. Morales, A. Bush, K. Alshafi, D. Bilton, E. W. Alton, A. Smithyman, and J. C. Davies. 2016. Antipseudomonal bacteriophage reduces infective burden and inflammatory response in murine lung. Antimicrob. Agents Chemother. 60 (2):744–51. doi:10.1128/AAC.01426-15.
  • Park, C. W., J. W. Park, S. H. Lee, and J. Hwang. 2014. Real-time monitoring of bioaerosols via cell-lysis by air ion and ATP bioluminescence detection. Biosens. Bioelectron. 52:379–83. doi:10.1016/j.bios.2013.09.015.
  • Park, J.-W., C. W. Park, S. H. Lee, and J. Hwang. 2015. Fast monitoring of indoor bioaerosol concentrations with ATP bioluminescence assay using an electrostatic rod-type sampler. PLoS One 10 (5):e0125251. 10.1371. doi:10.1371/journal.pone.0125251.
  • Partridge, S. R., S. M. Kwong, N. Firth, and S. O. Jensen. 2018. Mobile genetic elements associated with antimicrobial resistance. Clin. Microbiol. Rev. 31 (4):e00088–17. doi:10.1128/CMR.00088-17.
  • Pasquarella, C., O. Pitzurra, and A. Savino. 2000. The index of microbial air contamination. J. Hosp. Infect. 46 (4):241–56. doi:10.1053/jhin.2000.0820.
  • Patterson, J. P., D. B. Collins, J. M. Michaud, J. L. Axson, C. M. Sultana, T. Moser, A. C. Dommer, J. Conner, V. H. Grassian, M. D. Stokes, et al. 2016. Sea spray aerosol structure and composition using cryogenic transmission electron microscopy. ACS Cent. Sci. 2 (1):40–7. doi:10.1021/acscentsci.5b00344.
  • Paulson, J. N., O. C. Stine, H. C. Bravo, and M. Pop. 2013. Differential abundance analysis for microbial marker-gene surveys. Nat. Methods 10 (12):1200–2. PMID: 24076764 doi:10.1038/nmeth.2658.
  • Pavšic, J., J. Žel, and M. Milavec. 2016. Assessment of the real-time PCR and different digital PCR platforms for DNA quantification. Anal. Bioanal. Chem. 408 (1):107–21. doi:10.1007/s00216-015-9107-2.
  • Pearson, C., E. Littlewood, P. Douglas, S. Robertson, T. W. Gant, and A. L. Hansell. 2015. Exposures and health outcomes in relation to bioaerosol emissions from composting facilities: A systematic review of occupational and community studies. J. Toxicol. Environ. Health B Crit. Rev. 18 (1):43–69. doi:10.1080/10937404.2015.1009961.
  • Peccia, J., and M. Hernandez. 2006. Incorporating polymerase chain reaction-based identification, population characterization, and quantification of microorganisms into aerosol science: A review. Atmos. Environ. 40 (21):3941–61. doi:10.1016/j.atmosenv.2006.02.029.
  • Peters, M., P. Fritz, and A. Bufe. 2012. A bioassay for determination of lipopolysaccharide in environmental samples. Innate Immun. 18 (5):694–9. doi:10.1177/1753425912436590.
  • Pirtle, E. C., and G. W. Beran. 1991. Virus survival in the environment. Rev. Sci. Tech. 10 (3):733–48. doi:10.20506/rst.10.3.570.
  • Platts-Mills, T. A., and J. A. Woodfolk. 2011. Allergens and their role in the allergic immune response. Immunol. Rev. 242 (1):51–68. doi:10.1111/j.1600-065X.2011.01021.x.
  • Qian, J., D. Hospodsky, N. Yamamoto, W. W. Nazaroff, and J. Peccia. 2012. Size-resolved emission rates of airborne bacteria and fungi in an occupied classroom. Indoor Air 22 (4):339–51. doi:10.1111/j.1600-0668.2012.00769.x.
  • Qiu, X., J. I. Shu, O. Baysal, J. Wu, S. Qian, S. Ge, K. Li, X. Ye, N. Xia, and D. Yu. 2019. Real-time capillary convective PCR based on horizontal thermal convection. Microfluid. Nanofluid 23 (3):39. doi:10.1007/s10404-019-2207-0.
  • Quince, C., A. Walker, J. Simpson, N. Loman, and N. Segata. 2017. Shotgun metagenomics, from sampling to analysis. Nat. Biotechnol. 35 (9):833–44. doi:10.1038/nbt.3935.
  • Reponen, T., J. Lockey, D. I. Bernstein, S. Vesper, L. Levin, G. K. Khurana Hershey, S. Zheng, P. Ryan, S. A. Grinshpun, M. Villareal, et al. 2012. Infant origins of childhood asthma associated with specific molds. J. Allergy Clin. Immunol. 130 (3):639–44. doi:10.1016/j.jaci.2012.05.030.
  • Rittenour, W. R., J. Park, J. M. Cox-Ganser, D. H. Beezhold, and B. J. Green. 2012. Comparison of DNA extraction methodologies used for assessing fungal diversity via ITS sequencing. J. Environ. Monit. 14 (3):766. doi:10.1039/c2em10779a.
  • Rodrigues, N. M., Bronzato, G. F. Santiago, G. S. L. A. Botelho, B. M. Moreira, I. S. Coelho, M. M. Souza, S. M., and Coelho, S. 2017. The matrix-assisted laser desorption ionization–Time of flight mass spectrometry MALDI-TOF MS identification versus biochemical tests: A study with enterobacteria from a dairy cattle environment. Braz. J. Microbiol. 48 (1):132–8. doi:10.1016/j.bjm.2016.07.025.
  • Rohr, A. C., S. L. Campleman, C. M. Long, M. K. Peterson, S. Weatherstone, W. Quick, and A. Lewis. 2015. Potential occupational exposures and health risks associated with biomass-based power generation. Int. J. Environ. Res. Public Health 12 (7):8542–605. doi:10.3390/ijerph120708542.
  • Rylander, R. 1996. Airway responsiveness and chest symptoms after inhalation of endotoxin or 1→3-β-D-glucan. Indoor Built Environ. 5:106–11. doi:10.1159/000463694.
  • Rylander, R., Y. Peterson, and K. J. Donham. 1990. Questionnaire evaluating organic dust exposure. Am. J. Ind. Med. 17 (1):121–6. doi:10.1002/ajim.4700170142.
  • Saint-Marcoux, D., H. Proust, L. Dolan, and J. A. Langdale. 2015. Identification of reference genes for real-time quantitative PCR experiments in the liverwort Marchia polymorpha. PLoS One 10 (3):e0118678. doi:10.1371/journal.pone.0118678.
  • Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: A laboratory manual. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.
  • Sanchez-Monedero, M. A., and E. I. Stentiford. 2003. Generation and dispersion of airborne microorganisms from composting facilities. Process Saf. Environ. Prot. 81 (B3):166–70. doi:10.1205/095758203765639861.
  • Schloss, P. D., S. L. Westcott, T. Ryabin, J. R. Hall, M. Hartmann, E. B. Hollister, R. R. Lesniewski, B. B. Oakley, D. H. Parks, C. J. Robinson, et al. 2009. Introducing mothur: Open-Source, platform-independent, community-supported software for describing and comparing microbial communities. Appl. Environ. Microbiol. 75 (23):7537–41. doi:10.1128/AEM.01541-09.
  • Schrader, C., A. Schielke, L. Ellerbroek, and R. Johne. 2012. PCR inhibitors - occurrence, properties and removal. J. Appl. Microbiol. 113 (5):1014–26. doi:10.1111/j.1365-2672.2012.05384.x.
  • Schwarz, S., A. Loeffler, and K. Kadlec. 2017. Bacterial resistance to antimicrobial agents and its impact on veterinary and human medicine. Vet. Dermatol. 28 (1):82–e19. doi:10.1111/vde.12362.
  • Selman, M., Y. Lacasse, A. Pardo, and Y. Cormier. 2010. Hypersensitivity pneumonitis caused by fungi. Proc. Am. Thorac. Soc. 7 (3):229–36. doi:10.1513/pats.200906-041AL.
  • Seshadri, S., T. Han, V. Krumins, D. E. Fennell, and G. Mainelis. 2009. Application of ATP bioluminescence method to characterize performance of bioaerosol sampling devices. J. Aerosol Sci. 40 (2):113–21. doi:10.1016/j.jaerosci.2008.10.002.
  • Sharma, G., and R. R. Pandey. 2010. Influence of culture media on growth, colony character and sporulation of fungi isolated from decaying vegetable wastes. J. Yeast Fungal Res. 1 (8):157–64.
  • Sharma, R., Ranjan, R. Kapardar, R. K. A., and Grover, A. 2005. Unculturable bacterial diversity: An untapped resource. Curr. Sci. 89 (1):72–7.
  • Shin, S.-K., J. Kim, S-m Ha, H.-S. Oh, J. Chun, J. Sohn, and H. Yi. 2015. Metagenomic insights into the bioaerosols in the indoor and outdoor environments of childcare facilities. PLoS One 10 (5):e0126960. doi:10.1371/journal.pone.0126960.
  • Siersted, H. C., and S. Gravesen. 1993. Extrinsic allergic alveolitis after exposure to the yeast Rhodotorula rubra. Allergy 48 (4):298–9. doi:10.1111/j.1398-9995.1993.tb00734.x.
  • Sigaev, G. I., A. D. Tolchinsky, V. I. Sigaev, K. G. Soloviev, A. N. Varfolomeev, and B. T. Chen. 2006. Development of a cyclone-based aerosol sampler with recirculating liquid film: Theory and experiment. Aerosol Sci. Technol. 40 (5):293–308. doi:10.1080/02786820600596891.
  • Sil, A., and A. Andrianopoulos. 2015. Thermally dimorphic human fungal pathogens - Polyphyletic pathogens with a convergent pathogenicity trait. Cold Spring Harb. Perspect. Med. 5 (8):a019794. doi:10.1101/cshperspect.a019794.
  • Sinha, M., R. P. Singh, G. S. Kushwaha, N. Iqbal, A. Singh, S. Kaushik, P. Kaur, S. Sharma, and T. P. Singh. 2014. Current overview of allergens of plant pathogenesis related protein families. Sci. World J. 2014:1–19. doi:10.1155/2014/543195.
  • Soppa, J. 2014. Polyploidy in archaea and bacteria: About desiccation resistance, giant cell size, long-term survival, enforcement by a eukaryotic host and additional aspects. J. Mol. Microbiol. Biotechnol. 24 (5-6):409–19. doi:10.1159/000368855.
  • Spaan, S.,D. J. Heederik, P. S. Thorne, and I. M. Wouters. 2007. Optimization of airborne endotoxin exposure assessment: Effects of filter type, transport conditions, extraction solutions, and storage of samples and extracts. Appl. Environ. Microbiol. 73 (19):6134–43. doi:10.1128/AEM.00851-07.
  • Spaan, S., L. A. Smit, W. Eduard, L. Larsson, H. J. Arts, I. M. Wouters, and D. J. Heederik. 2008. Endotoxin exposure in sewage treatment workers: Investigation of exposure variability and comparison of analytical techniques. Ann. Agric. Environ. Med. 15 (2):251–61.
  • Stetzenbach, L. D., M. P. Buttner, and P. Cruz. 2004. Detection and enumeration of airborne biocontaminants. Curr. Opin. Biotechnol. 15 (3):170–4. doi:10.1016/j.copbio.2004.04.009.
  • Stone, B. A., and A. E. Clarke. 1992. Chemistry and biology of 1→3-β-glucans. Bundoora: La Trobe University Press.
  • Stewart, R. J., and T. M. Dowhanick. 1996. Rapid detection of lactic acid bacteria in fermenter samples using a nested polymerase chain reaction. J. Am. Soc. Brew. Chem. 54 (2):78–84. doi:10.1094/ASBCJ-54-0078.
  • Straumfors, A., A. Skogstad, T. Nilsen, O. Syness, I. Skaar, L. Hjeljord, and K. A. Afanou. 2014. Submicronic fungal bioaerosols: High-resolution microscopic characterization and quantification. Appl. Environ. Microbiol. 80 (22):7122–30. doi:10.1128/AEM.01740-14.
  • Suaifan, G. A., S. Alhogail, and M. Zourob. 2017. Paper-based magnetic nanoparticle-peptide probe for rapid and quantitative colorimetric detection of Escherichia coli O157:H7. Biosens. Bioelectron. 92:702–8. doi:10.1016/j.bios.2016.10.023.
  • Svabenska, E. 2012. Systems of detection and identification of biological aerosols. Def. Sci. J. 62 (6):404–11. doi:10.14429/dsj.62.1251.
  • Talley, S. M., P. D. Coley, and T. A. Kursar. 2002. The effects of weather on fungal abundance and richness among 25 communities in the Intermountain West. BMC Ecol. 2 (1):7.
  • Taylor, M. J., R. H. Bentham, and K. E. Ross. 2014. Limitations of using propidium monoazide with qPCR to discriminate between live and dead Legionella in biofilm samples. Microbiol. Insights. 7:15–24. doi:10.4137/MBI.S17723.
  • Tovey, E. R., D. J. M. Taylor, A. H. Graham, T. J. O'Meara, U. Lovborg, A. Jones, and R. Sporik. 2000. New immunodiagnostic system. Aerobiologia 16 (1):113–8. doi:10.1023/A:1007604430700.
  • Troeger, C., B. Blacker, I. A. Khalil, P. C. Rao, J. Cao, S. R. M. Zimsen, S. B. Albertson, A. Deshpande, T. Farag, Z. Abebe, et al. 2018. Estimates of the global, regional, and national morbidity. Lancet 18 (11):1191–210. doi:10.1016/S1473-30991830362-1.
  • Tseng, C. C., and C. S. Li. 2005. Collection efficiencies of aerosol samplers for virus-containing aerosols. J. Aerosol Sci. 36 (5-6):593–607. doi:10.1016/j.jaerosci.2004.12.004.
  • Tseng, C. C., and C. S. Li. 2006. Ozone for inactivation of aerosolized bacteriophages. Aerosol Sci. Technol. 40 (9):683–9. doi:10.1080/02786820600796590.
  • Van Leuken, J. P. G., A. N. Swart, A. H. Havelaar, A. Van Pul, W. Van der Hoek, and D. Heederik. 2016. Atmospheric dispersion modelling of bioaerosols that are pathogenic to humans and livestock—A review to inform risk assessment studies. Microb. Risk Anal. 1:19–39. doi:10.1016/j.mran.2015.07.002.
  • Vaughan, N. P. 1989. The Andersen impactor: Calibration, wall losses and numerical simulation. J. Aerosol Sci. 20 (1):67–90. doi:10.1016/0021-85028990032-3.
  • Verreault, D., S. Moineau, and C. Duchaine. 2008. Methods for sampling of airborne viruses. Microbiol. Mol. Biol. Rev. 72 (3):413–44. doi:10.1128/MMBR.00002-08.
  • Vestlund, A. T., R. Al-Ashaab, S. F. Tyrrel, P. J. Longhurst, S. J. T. Pollard, and G. H. Drew. 2014. Morphological classification of bioaerosols from composting using scanning electron microscopy. Waste Manage. 34:1101–8. doi:10.1016/j.wasman.2014.01.021.
  • Wagner, J., and J. Macher. 2012. Automated spore measurements using microscopy, image analysis, and peak recognition of near-monodisperse aerosols. Aerosol Sci. Technol. 46 (8):862–73. doi:10.1080/02786826.2012.674232.
  • Wang, Z., T. Reponen, S. A. Grinshpun, R. L. Gorny, and K. Willeke. 2001. Effect of sampling time and air humidity on the bioefficiency of filter samplers for bioaerosol collection. J. Aerosol Sci. 32 (5):661–74. doi:10.1016/S0021-8502(00)00108-7.
  • Warnes, S. L., C. J. Highmore, and C. W. Keevil. 2012. Horizontal transfer of antibiotic resistance genes on abiotic touch surfaces: Implications for public health. Mbio 3 (6):e489–12. doi:10.1128/mBio.00489-12.
  • Weiss, S. J., T. J. Mansell, P. Mortazavi, R. Knight, and R. T. Gill. 2016. Parallel mapping of antibiotic resistance alleles in Escherichia coli. PLoS One 11 (1):e0146916–17. E0146916. doi:10.1371/journal.pone.0146916.
  • Wheeler, D., and M. Bhagwat. 2007. Blast quickstart: Example-driven web-based blast tutorial. Methods Mol Biol. 395:149–76. doi:10.1385/1-59745-514-8:149.
  • Wheeler, E. K., C. A. Hara, J. Frank, J. Deotte, S. B. Hall, C. Spadaccini, and N. R. Beer. 2011. Under-three minute PCR: Probing the limits of the fast amplification. Analyst 18:3707–12. doi:10.1039/c1an15365j.
  • Willeke, K., X. Lin, and S. A. Grinshpun. 1998. Improved aerosol collection by combined impaction and centrifugal motion. Aerosol Sci. Technol. 28 (5):439–56. doi:10.1080/02786829808965536.
  • Williams, R. H., E. Ward, and H. A. McCartney. 2001. Methods for integrated air sampling and DNA analysis for detection of airborne fungal spores. Appl. Environ. Microbiol. 67 (6):2453–9. doi:10.1128/AEM.67.6.2453-2459.2001.
  • Wittmaack, K., H. Wehnes, U. Heinzmann, and R. Agerer. 2005. An overview on bioaerosols viewed by scanning electron microscopy. Sci. Total Environ. 346 (1-3):244–55. doi:10.1016/j.scitotenv.2004.11.009.
  • Womack, A. M., P. E. Artaxo, F. Y. Ishida, R. C. Mueller, S. R. Saleska, K. T. Wiedemann, B. J. M. Bohannan, and J. L. Green. 2015. Characterization of active and total fungal communities in the atmosphere over the Amazon rainforest. Biogeosciences 12 (21):6337–49. doi:10.5194/bg-12-6337-2015.
  • Womiloju, T. O., J. D. Miller, P. M. Mayer, and J. R. Brook. 2003. Methods to determine the biological composition of particulate matter collected from outdoor air. Atmos. Environ. 37 (31):4335–44. doi:10.1016/S1352-2310(03)00577-6.
  • World Health Organization (WHO). 1990. Indoor air quality biological contaminants. European series, No. 31. Copenhagen: WHO Regional Publications.
  • World Health Organization (WHO). 2002. Guidelines for concentration and exposure–response measurements of fine and ultra fine particulate matter for use in epidemiological studies. Geneva: World Health Organization.
  • Wu, Y., A. Calis, Y. Luo, C. Chen, M. Lutton, Y. Rivenson, X. Lin, H. C. Koydemir, Y. Zhang, H. Wang, et al. 2018. Label-free bioaerosol sensing using mobile microscopy and deep learning. ACS Photon. 5 (11):4617–27. doi:10.1021/acsphotonics.8b01109.
  • Wu, Y., F. Shen, and M. Yao. 2010. Use of gelatin filter and BioSampler in detecting airborne H5N1 nucleotides, bacteria and allergens. J. Aerosol Sci. 41 (9):869–79. doi:10.1016/j.jaerosci.2010.05.006.
  • Xu, S., and M. Yao. 2013. NanoPCR detection of bacterial aerosols. J. Aerosol Sci. 65:1–9. doi:10.1016/j.jaerosci.2013.06.005.
  • Xu, Z., and M. Yao. 2011. Analysis of culturable bacterial and fungal aerosol diversity obtained using different samplers and culturing methods. Aerosol Sci. Technol. 45 (9):1143–53. doi:10.1080/02786826.2011.582195.
  • Xu, Z., Y. Wu, F. Shen, Q. Chen, M. Tan, and M. Yao. 2011. Bioaerosol science, technology, and engineering: Past, present, and future. Aerosol Sci. Technol. 45 (11):1337–49. doi:10.1080/02786826.2011.593591.
  • Yang, D. C., K. M. Blair, and N. R. Salama. 2016. Staying in shape: The impact of cell shape on bacterial survival in diverse environments. Microbiol. Mol. Biol. Rev. 80 (1):187–203. Doi: doi:10.1128/MMBR.00031-15.
  • Yang, J., and G. Moon. 2018. Isolation of a Lactococcus lactis strain producing anti-staphylococcal bacteriocin. Korean J. Food Sci. Anim. Resour. 36 (6):1315–21. doi:10.5851/kosfa.2018.e67.
  • Yang, P., B. Wang, X. Bian, Y. S. Zhang, Y. Lan, L. Qiao, B. Liu, and W. Zhang. 2016. Microfluidic air sampler for highly efficient bacterial aerosol collection and identification. Anal. Chem. 88:11504–12. doi:10.1021/acs.analchem.6b02708.
  • Yang, W., S. Elankumaran, and L. C. Marr. 2011. Concentration and size distributions of airborne influenza A viruses measured indoors at a health centre, a day-care centre and on aeroplanes. J. R. Soc. Interface 8 (61):1176–84. doi:10.1098/rsif.2010.0686.
  • Yassine, F., N. Bassil, R. Flouty, A. Chokr, A. E. Samrani, G. Boiteux, and M. E. Tahchi. 2016. Culture medium pH influence on Gluconacetobacter physiology: Cellulose production rate and yield enhancement in presence of multiple carbon sources. Carbohydr. Polym. 146:282–91. doi:10.1016/j.carbpol.2016.02.003.
  • Yooseph, S., C. Andrews-Pfannkoch, A. Tenney, J. McQuaid, S. Williamson, M. Thiagarajan, D. Brami, L. Zeigler-Allen, J. Hoffman, J. B. Goll, et al. 2013. A metagenomic framework for the study of airborne microbial communities. PLoS One 12:e81862. doi:10.1371/journal.pone.0081862.
  • Zernike, F. 1942. Phase contrast, a new method for the microscopic observation of transparent objects. Physica 9 (7):686–98. doi:10.1016/S0031-8914(42)80035-X.
  • Zhang, C., and D. Xing. 2009. Parallel DNA amplification by convective polymerase chain reaction with various annealing temperatures on a thermal gradient device. Anal. Biochem. 387 (1):102–12. doi:10.1016/j.ab.2009.01.017.
  • Zhang, G., and S. Ghosh. 2000. Molecular mechanisms of NF-kappaB activation induced by bacterial lipopolysaccharide through Toll-like receptors. J. Endotoxin Res. 6 (6):453–7. doi:10.1179/096805100101532414.
  • Zhou, B., C. Wang, Q. Zhao, Y. Wang, M. Huo, J. Wang, and S. Wang. 2016. Prevalence and dissemination of antibiotic resistance genes and coselection of heavy metals in Chinese dairy farms. J. Hazard. Mater 320:10–7. doi:10.1016/j.jhazmat.2016.08.007.
  • Zhou, C., Y. Yang, and A. Y. Jong. 1990. Mini-prep in ten minutes. Biotechniques 8 (2):172–3.
  • Zhou, S., L. Pape. 2008. and, and D. C. Schwartz. Optical sequencing: Acquisition from mapped single-molecule templates. In Next-generation genome sequencing, ed. M. Janitz, 133–51. Weinheim: Wiley-VCH Verlag GMBH and Co.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.