Skip to main content

REVIEW article

Front. Public Health, 23 November 2023
Sec. Environmental health and Exposome
This article is part of the Research Topic Indoor Environment and Respiratory Diseases View all 3 articles

A comprehensive review of microbial contamination in the indoor environment: sources, sampling, health risks, and mitigation strategies

Hitikk Chawla&#x;Hitikk Chawla1Purnima Anand&#x;Purnima Anand2Kritika GargKritika Garg3Neeru BhagatNeeru Bhagat2Shivani G. VarmaniShivani G. Varmani4Tanu BansalTanu Bansal5Andrew J. McBain&#x;Andrew J. McBain6Ruchi Gulati Marwah
&#x;Ruchi Gulati Marwah2*
  • 1Institute for Cell Biology and Neuroscience, Goethe University Frankfurt, Frankfurt, Germany
  • 2Department of Microbiology, Bhaskaracharya College of Applied Sciences, University of Delhi, New Delhi, India
  • 3Department of Biosciences and Bioengineering, Indian Institute of Technology Roorkee, Roorkee, India
  • 4Department of Biomedical Science, Bhaskaracharya College of Applied Sciences, University of Delhi, New Delhi, India
  • 5Department of Biochemistry, All India Institute of Medical Sciences, New Delhi, India
  • 6School of Health Sciences, Faculty of Biology, Medicine and Health, The University of Manchester, Manchester, United Kingdom

The quality of the indoor environment significantly impacts human health and productivity, especially given the amount of time individuals spend indoors globally. While chemical pollutants have been a focus of indoor air quality research, microbial contaminants also have a significant bearing on indoor air quality. This review provides a comprehensive overview of microbial contamination in built environments, covering sources, sampling strategies, and analysis methods. Microbial contamination has various origins, including human occupants, pets, and the outdoor environment. Sampling strategies for indoor microbial contamination include air, surface, and dust sampling, and various analysis methods are used to assess microbial diversity and complexity in indoor environments. The review also discusses the health risks associated with microbial contaminants, including bacteria, fungi, and viruses, and their products in indoor air, highlighting the need for evidence-based studies that can relate to specific health conditions. The importance of indoor air quality is emphasized from the perspective of the COVID-19 pandemic. A section of the review highlights the knowledge gap related to microbiological burden in indoor environments in developing countries, using India as a representative example. Finally, potential mitigation strategies to improve microbiological indoor air quality are briefly reviewed.

1 Introduction

Around 90% of our time is spent indoors due to lifestyle changes and work habits; thus, indoor air quality is closely related to our health and comfort (1, 2). Many reports confirm that the concentration of pollutants can be 2 to 5 times higher inside than outside (3). In addition to the chemical pollutants, biological contaminants, such as bacteria, viruses, fungi, insects, mites, pollen, and pet dander, are also present in the indoor environment (4, 5). Microbial contaminants in the indoor environment can exist on surfaces, as suspended cells, or as bioaerosols (6). They show variations in numbers and types depending on the kind of indoor environment, sources of contamination, and other environmental factors. These chemical and biological pollutants may adversely impact indoor air quality leading to Indoor Air Pollution (IAP).

IAP is considered one of the five top risks influencing public health and presents a much higher risk than outdoor pollution. An estimated 3.2 million people die every year due to IAP (7, 8). Poor indoor air can have both immediate and long-term health effects, which are commonly referred to as Sick building syndrome (SBS), Building-related illnesses (BRIs), and Multiple chemical sensitivities (1, 9, 10). SBS refers to a collection of symptoms reported by the occupants or workers of a given building and is generally not attributed to a particular cause (10). These may include immediate or short-term effects such as irritation of the eyes, nose, and throat, headaches, dizziness, and fatigue (11, 12). This may also affect the nervous and cardiovascular systems and cause reduced fertility and congenital disabilities in the long term (8). BRI’s on the other hand refers to medical conditions such as hypersensitivities, asthma, and respiratory infections such as pneumonia, linked to a specific cause (13). Generally, exposure to microbes or their components is associated with three main groups of illness: toxicity, infections, and allergic reactions, including respiratory infections and other related diseases (1, 14, 15). Neonates, young children, older adults, and especially people suffering from co-morbid conditions are highly vulnerable to IAP (2, 16).

Particularly in developing nations, including India, the situation is more critical because of overpopulation and other socio-economic factors (17, 18). Tackling indoor air pollution is challenging and requires interdisciplinary efforts (19). There remains a significant knowledge gap and a lack of standards and guidelines, especially for assessing the microbiological quality of indoor air (20). Since the COVID-19 pandemic, IAP has been brought to the forefront because, globally, people are working remotely and spending more time indoors (21).

This review provides a comprehensive overview of microbial contaminants in diverse indoor environments, their sources, sampling and assessment strategies, the factors affecting their prevalence, and the associated health risks (Figure 1). The significance of IAP in the context of the COVID-19 pandemic is also highlighted. The severity of the problem is highlighted for developing countries with a particular reference to the scenario in India. The last section discusses various strategies to mitigate this challenge.

FIGURE 1
www.frontiersin.org

Figure 1. Various aspects of microbiological contamination of indoor air.

2 Methodology to conduct literature search

A thorough search of academic databases such as Google scholar and PubMed was conducted. We searched for specific terms such as ‘Microbial Indoor air quality,’ ‘Indoor air and Bioaerosols,’ ‘Bacteria in indoor air, ‘Fungi and molds in Indoor air,’ ‘Microbes in Indoor air and health’. The titles and abstracts of the articles were screened and assessed for the inclusion in the review. Articles were cross checked, duplicated and irrelevant articles were excluded. Further, relevant papers from the reference list of the included papers were also considered to ensure maximum coverage of the literature. A list of 430 papers were extensively discussed among the authors to resolve any discrepancies and reach a consensus. Finally, 314 references were included in this review, out of which 76% are from 2010 onwards.

3 Sources of microbial contaminants in indoor air

Outdoor pollutants influence the number and kind of microorganisms that enter the indoor environment displaying a source-sink relationship (22, 23). Apart from these, various inherent sources in the built environment contribute to the abundant microbial numbers and variety. Abiotic factors like moisture, relative humidity, and temperature can influence indoor microbiology (24, 25).

3.1 Humans

Human occupancy and their activities affect the microbial numbers and diversity in the indoor environment (25, 26). In the indoor air, respirable particulate matter and bacterial DNA increased with human occupancy (27). It has been estimated that humans shed approximately a billion skin cells daily, likely influencing microbial concentrations in the indoor air (28). Abundant bacteria and viruses from human oral and respiratory fluids can become aerosolized via talking, breathing, coughing, sneezing, etc. and reach varying distances depending on the droplet size (29). The frequency of cooking, vacuum cleaning, showering, and other human activities also affect microbial numbers and diversity (30). Patients, doctors, visitors, and hospital staff and their activities contribute to the pathogen load in the hospital environment (5, 31). Fungal concentrations have been reported to be elevated due to increased human movement from outside to inside, specifically linked to increased dust levels (32).

3.2 Pets

Pets also influence the indoor air microbiome. Animal skin, saliva, hair, fecal matter, and fleas are expected to contribute to microbial diversity in indoor air (28, 33, 34). Bacterial diversity and community richness have been reported to increase in households owning dogs and cats (35, 36). In a study involving 70 pet dogs, about 44 fungal isolates were obtained from the hair and skin of these animals (37).

3.3 Dust

House dust consists of hair, cotton fibers, bacteria, molds, and other particulate matter (3840). Dust-borne microbes can become resuspended into indoor air, increasing the risk of inhalation (41). Dust is reported to be dominated by skin-associated Gram-positive bacteria (35, 42, 43). Leppänen et al. (44) quantitatively assessed 259 house dust samples from both rural and urban homes in Finland. This study indicated that the fungal composition and seasonal variations correlated well between indoor air, settled, and reservoir dust samples. The latter showed reproducibility in repeated sampling over time. In another study by Wu et al. (45), bacterial composition and diversity in indoor dust samples were found to be affected by abiotic factors in the indoor environment. Marked differences in bacteria isolated from dust samples collected from university dormitory rooms and printing shops were reported. Human activities or strong air currents can also suspend respiratory viruses deposited in dust or indoor surfaces (46).

3.4 Building characteristics

Building design and maintenance, moisture build-up, and inadequate ventilation are common triggers for microbial build-up in the indoor environment (47). Plumbing systems can impact indoor air quality by adding bioaerosols to the built environment. Toilet flushing can generate large numbers of aerosolized bacteria from human faces, especially in case of poor ventilation, and if pathogenic, they carry the risk of transmission to healthy occupants (48, 49). In a toilet-seeding experiment by Barker and Jones (50), colony-forming units of Serratia marcescens rose sharply from 0 to 1,370 CFU/m3 after the first flush. Plumbing faults or leaks, water-damped carpets, ceilings, walls, and cramped building design can lead to increased dampness in the indoor environment, further increasing mold growth, odor, and microbial counts (47, 51). Additionally, the events of flooding can substantially increase the dampness inside houses which may favor the growth and dispersal of mold and bacteria. This was observed in the aftermath of flooded houses due to hurricane Katrina in New Orleans, Louisiana, United States (52). Construction and other materials used in buildings significantly contribute as potential sources as they can degrade and convert into organic compounds that support the growth of microorganisms (5). Both natural and mechanical ventilation practices influence the microbial ecology of indoor air (26, 53). Heating, ventilation, and air-conditioning systems (HVAC), primarily installed for air exchange, cleaning, and thermal comfort, can favor microorganisms’ growth if not properly maintained and cleaned (54). These microbes can remain viable on the internal filters of air-conditioning units for a long time. They can re-enter the indoor environment due to filter clogging, inefficient operation, improper maintenance, and malfunction (55). Further, leakage or condensation can wet the filters, favoring mold and bacterial proliferation (56).

3.5 Indoor plants

Microbial diversity and numbers are significantly contributed by indoor plants in the built environment (57, 58). Indoor plants can increase humidity levels and favor the excessive growth of molds and other airborne microorganisms (59). Due to agitation from the watering of plants or through the generation of strong air currents from fans, levels of airborne fungi were found to increase (60).

4 Sampling strategies

There is a lack of international consensus and standard operating procedures for collecting samples from indoor environments for both qualitative and quantitative estimation of microbial contaminants. International standards, including ISO 16000 series such as 16,000-17, 16,000-18, 16,000-19, 16,000-34, offer guidelines concerning sampling techniques, kinds of samplers, and sample analysis (47, 61, 62). Indoor air and dust samples are mainly used for microbial assessment of the indoor environment. Dust samples could be settled dust (on various surfaces like floors and tables) or reservoir dust (in mattresses, carpets, and bedding) (44). Dust collection is a quick but indirect method for sampling bioaerosols in the indoor environment. Various kinds of passive and active sampling techniques are widely and routinely used for the assessment of microbial contamination of indoor air.

4.1 Passive sampling

Passive sampling using ‘agar settle plates’ is one of the most widely practiced procedures to collect settled dust samples under the force of gravity. The air sample is collected according to the 1/1/1 scheme, plates are incubated, and results are expressed as CFU/m3 using the equation described by Omeliansky (63, 64). It is a simple, inexpensive, and unobtrusive sampling method (65). It gives comparable results, requires no special powered instruments or personnel, and is not influenced by engineering factors. Further, it provides a valid risk assessment if passive sampling is performed in an operation theater or near a surgical site (63, 66).

A simple and cost-effective dustfall collector developed by Wurtz et al. (67) is useful for prolonged airborne dust collection to measure significant concentrations of culturable fungi. Passive samplers based on electrostatic attractions, such as electrostatic dust clothes and the recently invented Rutgers electrostatic sampler, are also used (68, 69). Also, a vacuum cleaner can collect reservoir dust like those from mattresses/carpets or settled dust samples above floor surfaces (44, 70). Swabs are used to collect surface dust from table tops, computers, doors, walls, and cupboards. However, it allows only a qualitative measure of microorganisms’ airborne concentrations, and the sample’s age is unknown (47).

4.2 Active sampling

Active sampling collects airborne microorganisms present in inhalable dust in the indoor environment. In this, an air sampler physically draws a pre-set volume of air with the help of a pump, through or over a particle collection device into a liquid or solid culture medium or a nitrocellulose membrane (71). Various types of active air samplers like Anderson, Active Casella slit, Surface air system, and Coriolis cyclone samplers are available commercially based on sampling techniques such as filtration, impaction, impingement, and cyclone, each with its own sets of advantages and limitations (4, 61, 65). Active samplers require trained operators and a power supply, which might constrain their use in remote areas. Frequent replacement of collection media can help avoid reducing sample viability and overloading (65). Massoudinejad et al. (72) stated that active sampling was more sensitive and precise than passive sampling. This sampling technique is mainly applicable when the concentration of microorganisms is not very high, such as in an operating theater of a hospital (73).

These two sampling procedures have shown a good correlation in some studies (68, 74, 75) while others reported no correlation (70, 73, 76, 77). These variations in results could occur due to the type of sample and sampler used, the volume of air sampled, and the place or time of sampling (61, 78). Frankel et al. (79) pointed towards an urgent need for standardized sampling methodologies as they observed significant differences in the levels of culturable fungi, bacteria, endotoxin, and total inflammatory potential, in various air and dust samples collected from homes.

5 Sample analysis

Various culture-based and culture-independent methods employed to enumerate and identify microorganisms in the indoor environment are represented in Figure 2.

FIGURE 2
www.frontiersin.org

Figure 2. Techniques for the microbiological analysis of the indoor environment.

5.1 Culture-based methods

5.1.1 Culturing and microscopy

Microbial culturing techniques are simple, traditional, low-cost, and well-developed, primarily involving different types of microbiological media. Bacteria are routinely grown on general-purpose media such as nutrient agar, tryptic soy agar, blood agar, and casein soy peptone agar (80, 81). In addition to this, various selective media such as endo agar, eosin methylene blue agar, and mannitol salt agar are also employed (82). This method provides qualitative and quantitative measures of culturable bacteria and population diversity. Several broad-spectrum complex media, such as potato dextrose agar, malt extract agar, rose bengal agar, and dichloran glycerol 18 agar are used for fungal isolation and quantification (80, 81). Compared to bacteria, enumerating fungi is more complex and thus limits the comparison of data from different studies (7).

After processing the sample and using appropriate staining techniques, identification and enumeration of bacterial or fungal contaminants can be carried out by bright field microscopy (6). Alternatively, airborne microbial particles can be directly sampled onto glass slides, semisolid media, or membrane filters and microscopically examined (4). For enumeration and identification of fungi and their spores, hemocytometer chambers and adhesive tapes appressed onto indoor surfaces are also combined with microscopy (83). Both culturing and microscopic techniques are however time-consuming and tedious. Culturing methods show intrinsic constraints in obtaining actual microbial numbers and diversity. Direct microscopic counts are generally exceedingly higher than viable colony counts, commonly phrased as ‘the great plate count anomaly’ (84).

5.1.2 Matrix-assisted laser desorption/ionization time-of-flight mass spectrometry

MALDI-TOF MS is a highly reliable and rapid spectrometric technique compared to traditional microbiological and molecular methods and is thus useful for high-throughput microbial identification (85). In this, the mass-to-charge (m/z) ratio of ribosomal protein/peptide (analyte of interest) of the given microorganism is measured by determining the time required for it to travel the length of the flight tube following ionization with a laser beam (86). A unique mass spectrum called peptide mass fingerprint is obtained for the organism quickly. Identification is made using an algorithm-based approach from databases containing the MS reference spectra of peptides and proteins extracted from known microorganisms (87, 88). This technique is gaining popularity for identifying bacteria and fungi from indoor dust and air samples (8991). The microbial culture is transferred directly onto the sample target from the culture medium. MALDI-TOF MS is simple, cost-effective, and does not require highly skilled personnel (92).

5.1.3 Laser-induced fluorescence and laser-induced breakdown spectroscopy

LIF spectroscopy is a relatively novel technique for non-contact real-time online detection of bioaerosols. In this technique, the given bacterial or fungal culture is excited by laser light, showing the largest cross section with the selected wavelength (4, 93). Following de-excitation, the light of a longer wavelength is emitted, which is recorded by a photomultiplier tube resulting in the generation of a spectral fingerprint for each microbial species. Bioaerosols can be differentiated from other airborne particles based on their distinct fluorescence spectra and assessed quantitatively based on fluorescence intensity (94). It efficiently detects microbial contamination of surfaces in industrial and hospital cleanroom facilities, the food-processing industry, operating theaters, and as a biosafety measure in microbiological or medical laboratories or cases of suspicion of a deliberate or accidental release of biological warfare agents (95, 96). However, not all excited species may fluoresce, causing improper measurements (4). This technique has also been recently investigated for the first time for detecting Picorna viruses, thus showing the potential for rapid virological analysis with a substantial cost reduction (97, 98).

LIBS is yet another rapid, flexible, and real-time monitoring technique based on the unique atomic composition of plasma produced when a high-power pulsed laser is focused on a minimal area of the sample surface (99). The sample gets ablated, generating a plasma plume with temperatures over 100,000 K, breaking the sample into excited ionic and atomic species. The plasma then expands and cools within a short time, emitting radiation. Thus, the characteristic emission lines give information about the sample (4). Saari et al. (94, 100) demonstrated the potential of LIBS and LIF techniques as promising tools for the online detection and differentiation of bioaerosol types.

5.2 Culture-independent methods

5.2.1 Epifluorescence microscopy

Samples of airborne microbes collected in liquid buffer solutions or on membrane filters from the indoor environment are stained with fluorescent dyes and counted using epifluorescence microscopy (81, 101). DAPI and Acridine orange are popularly used for bioaerosol monitoring (4). It is a high throughput automated technique that allows the counting and differentiation of viable cells of culturable and non-culturable bacteria and fungi from non-viable cells or particles. Compared to conventional culturing, this technique gives much higher estimates of airborne microbial numbers for bacteria and fungi (102, 103), which was confirmed by Chi and Li (104) in a study conducted on bioaerosol concentrations and viability in swine buildings. However, the technique has the limited ability to differentiate between bacteria and fungi in case of size overlap and gives false positive results if the dye binds to the organic material (4, 102).

5.2.2 Flow cytometry

Flow cytometry is a rapid real-time technique for enumerating, detecting, and sorting microscopic particles suspended in a fluid stream with a high sample throughput (105, 106). It has been widely used for microbial analysis of environmental samples (107, 108). Different cell types cannot be distinguished only based on light scattering characteristics for bioaerosols, so using fluorescent dyes are recommended with this technique (106). After collecting indoor air samples in a liquid, flow cytometry with fluorescence is applied to quantify cells (4). The sample is hydrodynamically focused to create a single cell stream that passes in front of a laser beam, and the fluorescence emitted by each particle is then measured with photon detectors (105). The combination of FCM with fluorescence has been used to assess bacterial counts in hospital wards, reporting significantly higher counts than culture-based methods (109). This method can also detect and quantify airborne fungi (83). The significant drawbacks of FCM are high cost, the need for highly skilled labor, low-temperature requirements, and false positive results due to abiotic particles (4).

5.2.3 Single cell-based techniques: Raman microspectroscopy and nanoscale secondary ion mass spectrometry

Normal or spontaneous Raman spectroscopy and its variations have been used for the spectroscopic analysis of biochemical components of microorganisms. It is an effective, non-invasive, and label-free technique to extract the chemical fingerprint of an individual microbial cell (110). However, the spectra are obscured in standard Raman spectroscopy due to intrinsic fluorescence. Also, a burning effect is caused due to high-energy UV photons in UV resonance Raman enhancement spectroscopy (111). Surface-enhanced Raman spectroscopy (SERS) increases the Raman signal by suppressing the fluorescence and is effective under infrared and visible excitation and has thus emerged as an upcoming rapid, cultivation-independent technique for selective and real-time detection of single microbial aerosols (111, 112). The method involves adsorption or proximity of the analyte to a roughened gold or silver surface or nanoparticles. Following excitation by the incident light, the total enhancement of the SERS process is due to chemical and electromagnetic enhancement effects that generate a sizeable spectroscopic signal for the analyte (113). The major limitation is the reproducibility of the spectral data, as the laser has to strike at a point where the substrate and the analyte are present at an appropriate geometry. This limitation is overcome by combining Raman spectroscopy with other techniques. One study involved the combination of powerful magnification provided by scanning electron microscope in conjunction with the Raman interface to successfully target SERS active regions of the sample matrix to produce reproducible spectra (113).

Similarly in another study, bioaerosols were impacted and transferred to colloidal silver nanoparticles to obtain efficient Raman spectra (111). Schwarzmeier et al. (114) employed a Coriolis μ wet particle sampler with SERS to detect E. coli aerosols, facilitated using an advanced microarray readout system. Distinct Raman spectra of spores from several fungi associated with a damp indoor environment were obtained, suggesting the need to develop a library of Raman spectra for maximal differentiation of fungal spores at the species level (115).

NanoSIMS uses a sufficiently small beam size of approximately 50 nm to analyze single microbial cells or their parts. It combines features of microbial imaging techniques, stable isotope probing, and molecular biomarkers to study environmental microbial communities and draw comparisons (116, 117). Both Raman microspectroscopy and NanoSIMS, though have the potential to dissect and compare microbial communities in environmental samples but have not been used widely for profiling microbes in the indoor environment.

5.2.4 Molecular methods

Several culture-independent molecular methods have vastly improved the quality of data obtained for indoor air microbiomes in terms of greater microbial abundance and diversity and thus have changed the way we think about built environments (118, 119). The polymerase chain reaction is the most widely used molecular technique for environmental microbial samples. It is frequently used to amplify genes coding for 16S rRNA for prokaryotes and 18S rRNA and internal transcribed spacer region (ITS) for eukaryotes (28, 120). This technique is rapid and particularly useful when concentrations of airborne microorganisms are low. Indoor airborne bacteria and fungi associated have been analyzed by conventional PCR in many studies (22, 53, 121). In recent years, quantitative PCR (qPCR), also called real-time PCR, has gained popularity in providing accurate data on total microbial numbers in environmental samples (122). It has been used to assess bacterial (123) and fungal concentrations and types (70) in indoor air and dust samples. The limitations of PCR include a prior requirement of sequence knowledge to design the primers and non-specific binding of the primer to similar sequences on template DNA (124).

Amplicons obtained from PCR can be further analyzed using various genetic fingerprinting techniques like amplified ribosomal DNA restriction analysis, terminal restriction fragment length polymorphism (T-RFLP), automated ribosomal intergenic spacer analysis, denaturing gradient gel electrophoresis, temperature gradient gel electrophoresis, single-strand conformation polymorphism, and denaturing high-performance liquid chromatography (125). These techniques are widely used for studying microbial communities in environmental samples and have also been applied for indoor air microbial profiling (116). For instance, in a study by Weikl et al. (126) involving T-RFLP, the variation, and diversity of bacterial and fungal microbiomes were studied in dust samples from 286 households. The study indicated that both bacterial and fungal communities followed a temporal and seasonal pattern, with indoor and outdoor determinants affecting the fungal microbiome more strongly.

Our knowledge of indoor air microbiome diversity and community dynamics has been substantially enhanced by metagenomic-based sequencing, having the advantage of rapidly analyzing millions of samples in parallel with high sensitivity (127). The classic first-generation Sanger sequencing is often replaced by advanced sequencing approaches using platforms such as the second-generation Illumina MiSeq/HiSeq and Ion Torrent systems, third generation PacBio SMRT sequencing, and recently commercialized fourth-generation Oxford Nanopore MinION technology (128, 129).

These advanced identification methods are being increasingly applied to analyze microbial diversity in indoor air and dust samples (130, 131). In a study, indoor air samples from elementary schools and day-care centers in Korea showed the presence of a wide variety of taxa in microbial communities previously not identified by culture-based methods (132). Nygaard et al. (128) characterized microbiomes in the building-dust by comparing Nanopore MinION and Illumina MiSeq 16S rRNA gene sequencing. This study did not find significant differences in microbial composition between the two methods; however, a greater taxonomic resolution was achieved with MinION technology. Diverse indoor airborne viral communities associated with disparate hosts were also observed using metagenomic sequencing of dust samples obtained from the HVAC filters by Rosario et al. (133).

However, metagenomic-based sequencing studies of the indoor environment are limited by the availability of reference sequences in genome databases, low biomass of microbial aerosols leading to poor DNA extraction, and missing out the DNA of microbes present in low-abundance. Also, a significant deterrent to adopting these advanced methods for routine microbial surveillance is offset by the high cost of resources and machinery, especially in low- and middle-income countries (134, 135).

6 Microbial diversity of indoor air

The indoor microflora is complex and dynamic, with their numbers and diversity guided by their sources and associated environmental factors. There are no uniform standards for the acceptable levels of microbial contaminants in the indoor environment. A study by the WHO expert group to assess the health risks of biological agents in indoor environments suggested that the total microbial concentration should not exceed 1,000 CFU/m3 (47). However, some recommend 300 CFU/m3 and 750 CFU/m3, the maximum limit for fungi and bacteria (136). Also, the standards followed may vary from one country to another and from one kind of indoor environment to another. Table 1 provides an insight into the vast diversity of microflora in various indoor environments such as homes, offices, schools, universities, libraries, healthcare facilities, shopping centers, public restrooms, and gyms. In comparison to bacteria and fungi, there are limited studies on viruses.

TABLE 1
www.frontiersin.org

Table 1. An overview of the microbial diversity in different indoor environments (2017-2023).

6.1 Bacteria

The most common bacterial genera observed in dust and air samples collected from various types of indoor environments are Acinetobacter, Actinobacteria, Arthrobacter, Alcaligenes, Bacillus, Corynebacterium, Kocuria, Micrococcus, Propionibacterium, Staphylococcus, and Streptococcus species (43, 45, 140, 147). Barberan et al. (35) indicated that there could be differences in bacterial types isolated depending on the gender occupancy of a particular environment. Relatively, a greater abundance of Corynebacterium, Dermabacter, and Roseburia spp. was found in homes dominated by males, whereas those occupied mainly by females had more Lactobacillus spp. Further, seasonal variations are observed in bacterial populations in the indoor environment. In a study by Rintala et al. (157), the relative abundance of alpha- and beta-proteobacteria increased slightly toward summer in house dust samples. The proportion of firmicutes and gamma-proteobacteria was highest in winter and that of actinobacteria, alpha- and beta-proteobacteria in spring or summer, whereas the diversity of Bacteroides peaked in fall. Barberan et al. (35) reported that species of Bacteroides, Porphyromonas, Arthrobacter, Moraxella, Blautia, and Neisseria were associated with dogs, while species of Prevotella, Porphyromonas, Jeotgalicoccus, Sporosarcina, Moraxella, and Bifidobacterium were associated with cats, in household dust samples.

The indoor air of hospitals and healthcare settings can also be a source of harmful bacteria like Mycobacterium tuberculosis, Staphylococcus aureus, Escherichia coli, Legionella pneumophila, and Pseudomonas aeruginosa. Consequently, patients, healthcare workers, visitors, and mainly immunocompromised patients are at a higher risk of acquiring nosocomial infections (138, 144, 158). In particular, Legionella pneumophila, is a highly concerning indoor pathogen, that can easily get aerosolised from toilet or sink faucets, shower heads, bathtubs, plumbing systems, and is abundantly found in various healthcare facilities (159, 160). It has also been found to contaminate the dust in air conditioning and HVAC systems. Both water and air sampling are performed to evaluate the pathogen exposure and to suggest measures for its control. In a study using real time TaqMan PCR, Legionella pneumophila was found in 66.7% of the dust samples collected from the air ducts of central air conditioning system (161). Montagna et al. (138) conducted a multicentre study in Italy, to detect airborne contamination of Legionella from the water sources. They found that the liquid impingement technique (using Coriolis®μ) was effective in the collection of airborne pathogens combined with molecular investigations in comparison to passive and active dust sampling. Transmission of airborne antibiotic-resistant bacteria like methicillin-resistant Staphylococcus aureus, beta-lactam-resistant Acinetobacter sp., vancomycin-resistant streptococci, and their inhalable antibiotic-resistance genes in the hospital environment are also a significant threat to public health (162). In a study assessing bacterial contamination of surfaces, medical equipment, and indoor air of the pediatric ward and neonatal intensive care unit of an Ethiopian hospital, a quarter of the bacteria isolated were found to be multi-drug resistant, and most were associated with nosocomial infections (163).

6.2 Fungi

Exposure to fungi, their spores, hyphal fragments, and compounds like mycotoxins and beta-glucans in the indoor environment, is considered a major health risk mainly associated with hypersensitivity and allergic risks to the occupants (164). Outdoor air fungi mostly dominate the patterning of indoor air and can easily get suspended in the dust or settle on surfaces of any built environment (22, 32). The indoor temperature range of 10-35°C is highly conducive to the growth of fungi. Fungi can obtain nutrients from plant or animal matter in house dust, building, construction, and painting materials, cooking oil deposits, clothes, books, and paper (47). Dampness is a significant contributor to fungal growth and their spores, fragments, and allergens can easily get aerosolized in the indoor environment (47, 165). Both natural and mechanical ventilations are also considered as likely sources of fungal contamination. High humidity due to condensation of water and settled dust inside air-conditioning ducts, filters, and collecting trays are considered potential sites for high fungal contamination (165, 166). The prevalence of fungi in the indoor air of healthcare facilities poses a high risk of invasive fungal infections. This has been extensively studied and reviewed by Belizario et al. (167). The fungal genera predominating various indoor environments are Aspergillus, Penicillium, Fusarium, Alternaria, Cladosporium, Stachybotrys, Trichoderma, and yeasts like Candida spp. (137, 154, 168). They have frequently been related to hospital-acquired infections and exacerbation of respiratory symptoms (20, 167). The microbial content of rug dust and vacuum cleaner bag, dust samples determined by qPCR showed Penicillium, Aspergillus, and Paecilomyces to be present in the highest concentrations (169).

6.3 Viruses

Though bacteria and fungi are monitored frequently in various indoor environments, not many efforts have been made to investigate the viral diversity of the indoor air (101, 133). Respiratory infections, like the common cold, bronchiolitis, and influenza, can transmit through aerosolized droplets (170, 171). Respiratory syncytial, influenza, parainfluenza, corona, and adeno viruses are implicated in these infections (172). Low relative humidity and cold temperatures are believed to increase the transmission of respiratory viruses (173, 174). Human activities like breathing, coughing, sneezing, talking, and laughing release viral aerosols abundantly in the indoor air (175). Moon et al. (176) showed that the indoor air of residential apartments in South Korea mainly contained adenoviruses and influenza A virus, with concentrations of adenoviruses higher in winter than in spring. In another study by Yang et al. (177), one-third of the samples from the health center and two-thirds from the day-care center were confirmed to contain aerosolized influenza A virus.

The diversity of viruses in the indoor environment could be beyond human viruses. Rosario et al. (133) used a metagenomic approach to investigate the diversity of DNA and RNA viruses in the dust samples accumulated in HVAC filters of university dormitory rooms. They detected a broad diversity of viruses, including human papillomaviruses (HPVs) and polyomaviruses (HPyVs), which were found to be associated with a range of hosts. Rare zoonotic diseases caused by deadly viruses like Marburg, Ebola, Hanta, and Lassa can pose a potential nosocomial threat (178). During the SARS outbreak of 2003, infectious viral droplets entered buildings via sewage and drainage systems due to inadequacies in plumbing (179). The profuse transmission of SARS-CoV-2 and cluster cases were primarily due to the confined, crowded, and poorly ventilated indoor environment (180).

7 Microbe-associated health risks in the indoor environment

Exposure to microbes in the indoor environment can lead to adverse health effects like respiratory illnesses, including allergies, microbial infections, acute toxic effects, and cancer, as described below. The major routes of human exposure to airborne microorganisms and their products are mainly inhalation followed by ingestion and dermal contact (15). Various health effects following microbial exposure depend on the dose, route, and timing of exposure and the person’s genetic disposition (181). Molds and bacteria can also produce components like endotoxin, beta-1-3-glucan, muramic acid, ergosterol, allergenic proteins, volatile organic compounds of microbial origin (MVOCs), and mycotoxins (32, 182, 183) which can also pose additional health hazards.

7.1 Respiratory illnesses, allergies, and hypersensitive reactions

Exposure of the respiratory tract to microbes or their products can lead to mucous membrane irritation, allergic rhinitis, asthma, bronchitis, organic dust toxic syndrome (ODTS), chronic obstructive pulmonary disease (COPD), or allergic alveolitis (184, 185). Moldy and dusty environments are highly uncomfortable to susceptible individuals, with fungi such as Aspergillus, Penicillium, Cladosporium, Acremonium, Paecilomyces, and Mucor found to be commonly associated with respiratory infections and allergies (168). Asthma is common in children, with allergens and molds influencing the development of the disease (137). Fungal DNA has also been considered a risk factor for childhood asthma at home (181). However, according to hygiene hypothesis, exposure to microbes in early childhood could protect against allergies and asthma by stimulating the immune system, and that excessive cleanliness in the human environment might not necessarily be beneficial (186). Among fungi, Aspergillus species are the most commonly found indoor fungi. Aspergillus fumigatus and Aspergillus flavus are frequently encountered and responsible for sinusitis, allergic broncho-pulmonary aspergillosis (ABPA), and hospital-acquired infections (187, 188). High fungal or dust exposure has been linked to hypersensitive pneumonitis (189). Fungal metabolism produces volatile compounds such as 3-methyl furan that can irritate eyes, nose, and airways leading to headache, nausea, dizziness, and fatigue (40, 190). Inhalation of high doses of endotoxins, glucans, fungal spores, and mycotoxins can contribute to airway irritation and inflammation, decrease lung function, exacerbate asthma, and chronic conditions like ODTS and COPD (164, 191).

7.2 Infections

Airborne transmission is the most likely route for many microbial infections in the indoor environment. Legionella bacteria causing Pontiac fever and highly fatal pneumonia can rapidly spread in indoor environments through aerosolized contaminated water and inhalation of infected mist or vapor (138). Mycobacterium tuberculosis can be transmitted by inhaling the infectious droplet nuclei expectorated from sputum-positive patients in the indoor air of hospitals or other environments (192). Staphylococcus aureus, Pseudomonas aeruginosa, Enterococcus faecalis, Enterococcus faecium, Acinetobacter baumanii, and Escherichia coli are some of the most frequently encountered bacteria in the indoor air of hospitals, causing various nosocomial infections (193).

Immunocompromised patients are at potential risk for more severe opportunistic and systemic fungal infections caused by Blastomyces, Coccidioides, Cryptococcus, Histoplasma, Alternaria, Heliminthosporium, Cladosporium, Fusarium, Aspergillus, Phoma, and Penicillium (194, 195). Similarly, the presence of these fungi in the indoor air of critical hospital areas, especially neonatal and pediatric ICU, has been linked to the increased incidence of mucocutaneous colonization and a high risk of invasive fungal infections in neonates (167).

Human corona, respiratory syncytial, rhino, adeno, measles, mumps, rubella, and enteric viruses are readily transmitted through air droplets in the indoor environment causing various infections (196198). The COVID-19 pandemic has recently shown the extreme contagiousness of the Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) virus in the indoor environment, which has been responsible for high morbidity and mortality (199). A special section is dedicated here to the potential implications of poor indoor air quality in transmitting SARS-CoV2.

7.3 Cancer

Several studies have tried to ascertain the link between cancer and bioaerosol exposure (200202). Mycotoxins have been established as non-viral biological carcinogens (203, 204). Among them, aflatoxin from Aspergillus flavus is the best-known carcinogenic mycotoxin, particularly linked to liver cancer (205, 206). Ochratoxin A is also a possible human carcinogen (183, 207). The lung cancer risk of workers in the meat/poultry industry was reviewed with evidence pointing to oncogenic viruses associated with animals and heavy exposure to airborne material, including micro-organisms, fecal material, dander, and feathers (201).

8 Indoor air pollution from the COVID-19 perspective

The global pandemic of COVID-19 caused by the SARS-CoV-2 has left an unprecedented impact on various facets of human life, including health, lifestyle, economy, and the environment (208, 209). Terms like disinfection, masks, social distancing, and quarantine have become a part of our daily life. On the one hand, a few studies indicated improvement in ambient air quality in many cities during the lockdown, attributed to reduced industrial activities, vehicular movements, and various human activities (210213). On the other hand, the reduced outdoor pollution did not contribute to any health gains, as prolonged exposure to polluted indoor air negated this effect (214, 215). Poor indoor air quality increased COVID-19 infection and viral transmission in enclosed spaces (216). It took several months before the WHO acknowledged the airborne transmission of SARS CoV2, especially when people spent prolonged periods in indoor, crowded, inadequately ventilated spaces (217). The aerosolized virus is stable for 3 h and can travel long distances in closed and open environments (214). These aerosols need to be controlled to reduce the risk of new infections. Also, in a study conducted in Bergamo, Northern Italy, viral RNA was isolated from the particulate matter, possibly explaining a higher COVID-19 burden in areas with high air pollution (218).

People were compelled to stay inside owing to home quarantine and lockdowns. This may have impacted indoor air quality due to insufficient ventilation, overcrowding, recirculation of polluted air, and increased household activities and office work (214, 215). Further, aerosolized virus released from infected people quarantined at homes, or hospitals could have increased the risk and deteriorated the indoor air (219). A study showed that the cough from an individual with a high SARS-CoV-2 load could contain 7.44 million viral copies/m3 (220). If the quality of the air is poor to begin with, people may be inadvertently exposed to various SBS-related symptoms (221) putting people with pre-existing medical conditions at a higher risk. Due to the high transmission of the SARS-CoV-2 virus in indoor spaces, IAP has thus become a table talk worldwide (180). The COVID-19 pandemic has thus strongly taught us to focus equally on sustainable and safe indoor spaces, as poor indoor air, can likely intensify the impact of air-borne viral outbreaks and other microbial infections (222, 223).

9 Indoor air pollution in India

In developing countries, including India, people are exposed to the highest air pollution in the indoor environment (224, 225). There is a general lack of awareness among people in India as air pollution is only thought to be associated with the outdoor environment, and the inhabitants are considered safe indoors (18). Here, rural and urban poor populations still use simple solid fuels that release abundant amounts of suspended particulate matter, harmful chemicals and gasses and have been associated with severe respiratory and other health risks (226). Overcrowded and confined indoor spaces, improper ventilation, inadequacies in building design, and various socio-economic factors also aggravate poor indoor air quality (213, 227). Microbial contaminants also significantly contribute to IAP; however, in India, this aspect has not been dealt with extensively (228) and can pose various health hazards to the occupants of the indoor environment, as discussed earlier.

Further, there are no standard guidelines to monitor and assess various parameters related to IAP, including microbiological assessment (18). The Central Pollution Control Board (229) under the Ministry of Environment and Forests (MoEF), Government of India, does not include IAP in its agenda and mainly focuses on outdoor air pollution. Thus, this knowledge and skill gap is highly concerning (18). Damp, overcrowded houses, and poorly ventilated buildings are a hot spot for microbial contaminants and a source of respiratory illnesses like tuberculosis, contributing significantly to the national disease burden (230). India hosts diverse climatic conditions, and thus the indoor microflora diversity and concentration show a lot of seasonal variations (5, 231, 232). Some representative studies on the isolation and assessment of microflora in various indoor environments conducted in India’s urban and rural settings are highlighted in Table 2.

TABLE 2
www.frontiersin.org

Table 2. Studies on microbiological assessment of the indoor environment in India (2012-2022).

However, it is clear from these studies that primarily passive sampling procedures combined with conventional culturing methods have been used. Thus, a large proportion of non-culturable microbes still remain unassessed. There is a need to introduce advanced molecular and spectrometric techniques to routinely monitor microbiological indoor air quality and bridge this knowledge gap. Further, mainly bacterial and fungal contaminants have been studied, with only a few studies assessing viruses and microbial products such as endotoxins, mycotoxins, and microbial allergens in the indoor environment (252, 254).

10 Strategies for microbial control in the indoor environment

Various mitigation strategies aim to control microorganisms and improve indoor air quality by optimizing ventilation systems, controlling emission sources, and developing air purification technologies, as discussed here (255).

10.1 Ventilation

Ventilation can help remove or dilute pollutants and control humidity or dampness in the built environment, thereby maintaining the health and comfort of the occupants (256, 257). International agencies such as ASHRAE (American Society of Heating, Refrigerating, and Air-Conditioning Engineers), WHO, REHVA (Representatives of European Heating and Ventilation Associations), and CDC (Center for Disease Control and Prevention) have also recognized that the transmission of airborne diseases can be effectively managed by ventilation (257259). For instance, CDC (258) has provided detailed guidelines including cost considerations for improving ventilation in buildings. Ventilation can be provided by natural or mechanical means (260, 261). Natural or passive ventilation is achieved through airflow provided by doors, windows, louvers, and vents (7, 259). Various studies have reported that natural ventilation is effective in controlling the transmission of the infectious airborne microbes (260, 262). In low resource healthcare settings, Escombe et al. (263) demonstrated that it could be used as a low-cost and energy efficient measure to reduce the risk of transmission of nosocomial tuberculosis. Further, the rampant spread of corona virus brought natural ventilation strategy to the forefront of the discussion as a crucial factor that can help in lowering the viral build-up in the indoor spaces. In this context, Vignolo et al. (264) qualitatively assessed airborne viral transmission and natural ventilation in school classroom in Uruguay and reported that periodic ventilation can act as a useful strategy to reduce transmission of airborne microbes. However, in another study, no impact was reported (265). Despite the low cost, energy efficiency, low maintenance, and higher airflow changes, natural ventilation is difficult to maintain in a tight building without cross ventilation. Further, it is not a feasible strategy for healthcare facilities and countries with colder climates as natural ventilation can increase the thermal discomfort for the occupants (266). Thus, a cross-sectional study involving all the factors can potentially give an insight for optimal use of natural ventilation.

Mechanical ventilation uses an air handling system like HVAC that circulates fresh and recycled air via ducts (261). Healthcare facilities with efficiently maintained and operated mechanical air conditioning, and ventilation systems were found to be less contaminated than buildings with naturally ventilated systems (267). In a study by Gołofit-Szymczak and Górny (268), different ventilation systems were assessed for their impact on microbiological indoor air quality of 15 office buildings. Higher concentrations of both bacteria and fungi were observed in naturally ventilated office in comparisons with offices having air-conditioning and mechanical ventilation systems. Similarly, HVAC system was found to effectively remove fungal spores (269). However, it is essential to note that mechanical ventilation systems may be a source of biological contamination, as discussed earlier. Further, the high transmission of SARS-CoV-2 in indoor spaces has made us all think critically about indoor air quality, especially in hospitals, and the approaches to controlling nosocomial infections (270). Fonseca et al. (271) suggested that the mechanical ventilation systems in healthcare facilities could be complemented with natural ventilation to adequately control relative humidity and CO2 concentrations to minimize the risk of airborne infections. Both mechanical and natural ventilation systems combined in a hybrid technology can also provide sustainability in terms of energy efficiency (272), and the shortcomings of natural ventilation can be overcome by the mechanical components (273). The frequency of regular monitoring, cleaning, and maintenance procedures can be adjusted based on operating hours and human occupancy to achieve good air quality (20).

Studies show that HVAC filtration can effectively control pathogens penetrating the building envelope and their transmission in the indoor environment (274, 275). Thus, to improve the performance of HVAC systems, panel filters are installed in their ducts (274, 276, 277). These filters are installed as per the standard guidelines and used widely (278, 279). HEPA or High-Efficiency Particulate Air filters must be cleaned and replaced periodically to maintain proper function (280). By following the air route and sources of contamination through sequential sampling, Cabo Verde et al. (20) assessed the efficacy of HEPA filters installed in the ventilation systems of operation theaters and found an effective reduction in bacterial concentration from outdoors to indoors. These filters have the limited ability to inactivate the collected biological agents, which may regrow with adequate humidity levels (281). Nanofiber filters can overcome this problem and are very effective in indoor environments where stringent air purification is needed, such as hospitals, food, and pharmaceutical industries (282).

The inactivation of microbial contaminants in indoor air could also be brought about by utilizing other methods integrated with the HVAC system (275, 280, 283). These include ultraviolet germicidal irradiation (UVGI) (284, 285); chlorine dioxide gas (286); antimicrobial compounds (287); electrostatic technology (288); photocatalytic oxidation (289); plasma cluster ion technology (290); microwave heating (291); ozone sterilization (292, 293); and dispersion of atomized nanoparticles in the air (294). UVGI system uses UVC rays (200–280 nm) to form photo dimers in nucleic acids to bring about disinfection and has been widely recognized and recommended for installation, especially in healthcare facilities (295, 296). However, it is pertinent to mention that ozone-based technologies can themselves pose an additional health risk by releasing ozone as a harmful by-product (297). Recently, silver nanoparticles and multiwall carbon nanotubes coated with hybrid polypropylene nano-filter were tested in air-conditioning and found to be very effective in killing Pseudomonas aeruginosa, Salmonella enterica, and Staphylococcus aureus (298). Thus, two or more such technologies can be synergized to improve the efficient functioning of the HVAC systems to remove microbial contaminants (283, 299).

10.2 Air cleaners and purifiers

Air Purifiers/Cleaners (AP) have become essential for reducing IAP. There has been a considerable increase in sales of air purifiers from 0.8 million units in 2015 to almost 2 million units in 2018, with a similar projection for the future (300, 301). Portable air purifiers utilizing HEPA filtration are commonly used for homes. Air purifiers’ advantages are simple installation, portability, and the lack of harmful by-products (302). Their configuration generally employs a multilayer filter system consisting of a prefilter, a carbon filter, an antibacterial filter, and a HEPA filter (255, 303). Also, APs with disinfection capability are more effective than those with only HEPA filtration (216). In these, UVGI lights are installed within the body of the air cleaners. Combined HEPA and UVGI, air cleaning technologies could effectively remove airborne pathogens (304, 305). Air purifiers or cleaners based on ionization, electrostatic precipitation, cold plasma generation, and photocatalytic oxidation are emerging technologies (216, 255, 306309). However, currently, no single technology can sufficiently achieve “cleaner” indoor air. Current research is focused on combined and innovative alternatives, such as plasma-catalytic hybrid systems, hybrid ozonation systems, and biofilter-adsorption systems, that can work synergistically to achieve optimum indoor air quality (310312).

10.3 Housekeeping measures

The frequency of housekeeping can also help in lowering the microbial content of the indoor air. This includes basic measures like dusting with a damp cloth or using electrostatic cloth, vacuuming, carpet and rug cleaning, and removal of shoes while entering the house. Persistent dampness and moldy growth on indoor surfaces and in building structures can be minimized. For instance, USEPA has provided guidelines for the control of moisture and mold in homes (313). Franke et al. (314) carried long-term monitoring of a building for biological, chemical, and particulate matter, following a standardized cleaning program. This improved house-keeping initiative was found to be effective in the reduction of culturable bacteria and fungi, airborne dust mass and total volatile organic compounds. However, barring a few reports, there are no direct studies that have ascertained the link between house-keeping and indoor microbial concentrations. This is definitely a thrust area that will also aim to increase awareness about the impact of poor indoor air quality on an individual’s health especially children and older people and those with underlying health conditions.

11 Conclusion

Rapid industrialization and urbanization have led people to spend most of their time indoors in tight-built environments. Thus, indoor air quality plays a significant role in their general state of health and comfort. They are exposed to a variety of pollutants, including microorganisms. A vast ecological niche has been created for these minuscule organisms leading to their presence in various microhabitats of the indoor environment, thus presenting us with a complex ecosystem that requires a more profound understanding. Compared to bacteria and fungi, there are limited studies on the viral diversity of the indoor environment. Protocols for sampling and assessing microorganisms need to be standardized and validated as the threshold limits can vary from one indoor environment to another, region to region, and country to country. Culture-independent methods can complement culture-based methods in comprehensively monitoring microbial composition and community structure. Advancements in metagenomic next-generation sequencing approaches have paved the way for better analysis of viral diversity which has been relatively unexplored in comparison to bacteria and fungi. To minimize the knowledge gap, a repository of microbial sequences obtained from different kinds of built environments must be maintained as a referential database source.

Developed countries mostly follow specific guidelines and standards for building design and management of indoor air quality. However, the situation is grave in developing countries with a high disease burden aggravated by various socio-economic factors. A concerted effort is needed to raise awareness about the issue, as simple source control measures can make a massive difference in the quality of the built environment. For example, basic housekeeping measures to prevent dust build-up or humidity control can curtail microbial growth, for which people need to be educated through awareness programs and surveys. Cross-sectional and evidence-based studies will provide a realistic analysis of the microbial impact and help design feasible mitigation strategies. We need to focus on hybrid air cleaning technologies and sustainable building architecture that combines energy saving with appropriate ventilation to minimize the buildup of indoor air pollutants. A multifactorial interdisciplinary approach will involve a synergy between microbiologists, public health experts, engineers, architects, environmental agencies, and the government to devise sustainable solutions to this serious public health issue.

Author contributions

HC: Writing – original draft, Data curation, Investigation. PA: Data curation, Writing – original draft, Writing – review & editing. KG: Investigation, Writing – original draft. NB: Data curation, Investigation, Writing – review & editing. SV: Investigation, Writing – review & editing. TB: Writing – original draft. AM: Writing – review & editing. RM: Conceptualization, Supervision, Writing – original draft.

Funding

The author(s) declare that no financial support was received for the research, authorship, and/or publication of this article.

Acknowledgments

The authors acknowledge the use of Biorender.com for creating the figures used in the manuscript.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1. Van, TV, Park, D, and Lee, YC. Indoor air pollution, related human diseases, and recent trends in the control and improvement of indoor air quality. Int J Environ Res Public Health. (2020) 17:2927. doi: 10.3390/ijerph17082927

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Sundell, J. On the history of indoor air quality and health. Indoor Air. (2004) 14:51–8. doi: 10.1111/j.1600-0668.2004.00273.x

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Smith, KR. National burden of disease in India from indoor air pollution. Proc Natl Acad Sci. (2000) 97:13286–93. doi: 10.1073/pnas.97.24.13286

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Ghosh, B, Lal, H, and Srivastava, A. Review of bioaerosols in indoor environment with special reference to sampling, analysis and control mechanisms. Environ Int. (2015) 85:254–72. doi: 10.1016/j.envint.2015.09.018

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Kumar, P, MohdA, K, Singh, AB, and Singh, R. Biological contaminants in the indoor air environment and their impacts on human health. Air Qual Atmos. Health. (2021) 14:1723–36. doi: 10.1007/s11869-021-00978-z

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Stetzenbach, LD, Buttner, MP, and Cruz, P. Detection and enumeration of airborne biocontaminants. Curr Opin Biotechnol. (2004) 15:170–4. doi: 10.1016/j.copbio.2004.04.009

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Flannigan, B, Samson, RA, and David, MJ. Microorganisms in home and indoor work environments. Boca Raton, FL: CRC Press (2002).

Google Scholar

8. WHO. Household air pollution and health. Geneva: World Health Organization (2022).

Google Scholar

9. Spachos, P, and Plataniotis, K. Environmental monitoring for smart buildings In: P Spachos, editor. Transportation and power grid in smart cities. Chichester, UK: John Wiley and Sons, Ltd. (2018). 327–54.

Google Scholar

10. Nag, PK. Sick building syndrome and other building-related illnesses. Berlin: Springer, pp. 53–103. (2019).

Google Scholar

11. USEPA. Introduction to indoor air quality. United States Environmental Protection Agency, United States Government (2021) Available at: https://www.epa.gov/indoor-air-quality-iaq/introduction-indoor-air-quality.

Google Scholar

12. Berglund, B, Brunekreef, B, Knoppe, H, Lindvall, T, Maroni, M, Molhave, L, et al. Effects of indoor air pollution on human health. Indoor Air. (1992) 2:2–25. doi: 10.1111/j.1600-0668.1992.02-21.x

CrossRef Full Text | Google Scholar

13. Horvath, EP. Building-related illness and sick building syndrome: from the specific to the vague. Cleve Clin J Med. (1997) 64:303–9. doi: 10.3949/ccjm.64.6.303

CrossRef Full Text | Google Scholar

14. Bernstein, JA, Alexis, N, Bacchus, H, Bernstein, IL, Fritz, P, Horner, E, et al. The health effects of nonindustrial indoor air pollution. J Allergy Clin Immunol. (2008) 121:585–91. doi: 10.1016/j.jaci.2007.10.045

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Srikanth, P, Sudharsanam, S, and Steinberg, R. Bio-aerosols in indoor environment: composition, health effects and analysis. Indian J Med Microbiol. (2008) 26:302–12. doi: 10.1016/S0255-0857(21)01805-3

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Holgate, S, Grigg, J, Arshad, H, Carslaw, N, Cullinan, P, Dimitroulopoulou, S, et al. Health effects of indoor air quality on children and young people In: S Holgate, editor. Issues in environmental science and technology. London, UK: Royal Society of Chemistry (2020). 151–88.

Google Scholar

17. Bruce, N, Perez-Padilla, R, and Albalak, R. Indoor air pollution in developing countries: a major environmental and public health challenge. Bull World Health Organ. (2000) 78:1078–92.

PubMed Abstract | Google Scholar

18. Goyal, R, and Khare, M. Indoor air quality: current status, missing links and future road map for India. J Civil Environ Eng. (2012) 2:118. doi: 10.4172/2165-784X.1000118

CrossRef Full Text | Google Scholar

19. Baecher, C, Sohui, F, Ball, L, and Masson, O. Indoor air quality: a multifaceted public health problem. Field Actions Sci Rep. (2020) 2020:6–7.

Google Scholar

20. Cabo Verde, S, Almeida, SM, Matos, J, Guerreiro, D, Meneses, M, Faria, T, et al. Microbiological assessment of indoor air quality at different hospital sites. Res Microbiol. (2015) 166:557–63. doi: 10.1016/j.resmic.2015.03.004

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Mumtaz, R, Zaidi, SMH, Shakir, MZ, Shafi, U, Malik, MM, Haque, A, et al. Internet of things (IoT) based indoor air quality sensing and predictive analytic–a COVID-19 perspective. Electronics (Basel). (2021) 10:184. doi: 10.3390/electronics10020184

CrossRef Full Text | Google Scholar

22. Adams, RI, Miletto, M, Taylor, JW, and Bruns, TD. Dispersal in microbes: fungi in indoor air are dominated by outdoor air and show dispersal limitation at short distances. ISME J. (2013) 7:1262–73. doi: 10.1038/ismej.2013.28

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Miletto, M, and Lindow, SE. Relative and contextual contribution of different sources to the composition and abundance of indoor air bacteria in residences. Microbiome. (2015) 3:61. doi: 10.1186/s40168-015-0128-z

CrossRef Full Text | Google Scholar

24. Prussin, AJ, and Marr, LC. Sources of airborne microorganisms in the built environment. Microbiome. (2015) 3:78. doi: 10.1186/s40168-015-0144-z

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Adams, RI, Bhangar, S, Pasut, W, Arens, EA, Taylor, JW, Lindow, SE, et al. Chamber bioaerosol study: outdoor air and human occupants as sources of indoor airborne microbes. PLoS One. (2015) 10:e0128022. doi: 10.1371/journal.pone.0128022

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Meadow, JF, Altrichter, AE, Kembel, SW, Kline, J, Mhuireach, G, Moriyama, M, et al. Indoor airborne bacterial communities are influenced by ventilation, occupancy, and outdoor air source. Indoor Air. (2014) 24:41–8. doi: 10.1111/ina.12047

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Hospodsky, D, Qian, J, Nazaroff, WW, Yamamoto, N, Bibby, K, Rismani-Yazdi, H, et al. Human occupancy as a source of indoor airborne Bacteria. PLoS One. (2012) 7:e34867. doi: 10.1371/journal.pone.0034867

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Kelley, ST, and Gilbert, JA. Studying the microbiology of the indoor environment. Genome Biol. (2013) 14:202. doi: 10.1186/gb-2013-14-2-202

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Xie, X, Li, Y, Sun, H, and Liu, L. Exhaled droplets due to talking and coughing. J R Soc Interface. (2009) 6:S703–14. doi: 10.1098/rsif.2009.0388.focus

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Chen, Q, and Hildemann, LM. The effects of human activities on exposure to particulate matter and bioaerosols in residential homes. Environ Sci Technol. (2009) 43:4641–6. doi: 10.1021/es802296j

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Nevalainen, A, and Seuri, M. Of microbes and men. Indoor Air. (2005) 15:58–64. doi: 10.1111/j.1600-0668.2005.00344.x

CrossRef Full Text | Google Scholar

32. Nevalainen, A, Täubel, M, and Hyvärinen, A. Indoor fungi: companions and contaminants. Indoor Air. (2015) 25:125–56. doi: 10.1111/ina.12182

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Fujimura, KE, Johnson, CC, Ownby, DR, Cox, MJ, Brodie, EL, Havstad, SL, et al. Man’s best friend? The effect of pet ownership on house dust microbial communities. J Allergy Clin Immunol. (2010) 126:410–412.e3. doi: 10.1016/j.jaci.2010.05.042

CrossRef Full Text | Google Scholar

34. Kettleson, EM, Adhikari, A, Vesper, S, Coombs, K, Indugula, R, and Reponen, T. Key determinants of the fungal and bacterial microbiomes in homes. Environ Res. (2015) 138:130–5. doi: 10.1016/j.envres.2015.02.003

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Barberán, A, Dunn, RR, Reich, BJ, Pacifici, K, Laber, EB, Menninger, HL, et al. The ecology of microscopic life in household dust. Proc R Soc B Biol Sci. (2015) 282:20151139. doi: 10.1098/rspb.2015.1139

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Mäki, JM, Kirjavainen, PV, Täubel, M, Piippo-Savolainen, E, Backman, K, Hyvärinen, A, et al. Associations between dog keeping and indoor dust microbiota. Sci Rep. (2021) 11:5341. doi: 10.1038/s41598-021-84790-w

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Jang, K-S, Yun, Y-H, Yoo, H-D, and Kim, SH. Molds isolated from pet dogs. Mycobiology. (2007) 35:100. doi: 10.4489/MYCO.2007.35.2.100

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Harrison, RM, Jones, AM, Biggins, PDE, Pomeroy, N, Cox, CS, Kidd, SP, et al. Climate factors influencing bacterial count in background air samples. Int J Biometeorol. (2005) 49:167–78. doi: 10.1007/s00484-004-0225-3

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Maron, P-A, Lejon, DPH, Carvalho, E, Bizet, K, Lemanceau, P, Ranjard, L, et al. Assessing genetic structure and diversity of airborne bacterial communities by DNA fingerprinting and 16S rDNA clone library. Atmos Environ. (2005) 39:3687–95. doi: 10.1016/j.atmosenv.2005.03.002

CrossRef Full Text | Google Scholar

40. Korpi, A, Pasanen, A-L, Pasanen, P, and Kalliokoski, P. Microbial growth and metabolism in house dust. Int Biodeterior Biodegradation. (1997) 40:19–27. doi: 10.1016/S0964-8305(97)00032-2

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Ferro, AR, Kopperud, RJ, and Hildemann, LM. Source strengths for indoor human activities that Resuspend particulate matter. Environ Sci Technol. (2004) 38:1759–64. doi: 10.1021/es0263893

CrossRef Full Text | Google Scholar

42. Pakarinen, J, Hyvärinen, A, Salkinoja-Salonen, M, Laitinen, S, Nevalainen, A, Mäkelä, MJ, et al. Predominance of gram-positive bacteria in house dust in the low-allergy risk Russian Karelia. Environ Microbiol. (2008) 10:3317–25. doi: 10.1111/j.1462-2920.2008.01723.x

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Rintala, H, Pitkäranta, M, and Täubel, M. Microbial communities associated with house dust. Adv Appl Microbiol. (2012) 78:75–120. doi: 10.1016/B978-0-12-394805-2.00004-X

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Leppänen, HK, Täubel, M, Jayaprakash, B, Vepsäläinen, A, Pasanen, P, and Hyvärinen, A. Quantitative assessment of microbes from samples of indoor air and dust. J Expo Sci Environ Epidemiol. (2018) 28:231–41. doi: 10.1038/jes.2017.24

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Wu, Z, Lyu, H, Ma, X, Ren, G, Song, J, Jing, X, et al. Comparative effects of environmental factors on bacterial communities in two types of indoor dust: potential risks to university students. Environ Res. (2022) 203:111869. doi: 10.1016/j.envres.2021.111869

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Rawat, MS, and Ferro, AR. Respiratory virus deposition and resuspension from indoor surfaces In: D Barry and H Kanematsu, editors. Studies to combat COVID-19 using science and engineering. Singapore: Springer Nature Singapore (2022). 107–18.

Google Scholar

47. WHO. WHO guidelines for indoor air quality: Dampness and mould. Geneva: WHO, pp. 1–248. (2009).

Google Scholar

48. Lai, ACK, Tan, TF, Li, WS, and Ip, DKM. Emission strength of airborne pathogens during toilet flushing. Indoor Air. (2018) 28:73–9. doi: 10.1111/ina.12406

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Lou, M, Liu, S, Gu, C, Hu, H, Tang, Z, Zhang, Y, et al. The bioaerosols emitted from toilet and wastewater treatment plant: a literature review. Environ Sci Pollut Res. (2021) 28:2509–21. doi: 10.1007/s11356-020-11297-8

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Barker, J, and Jones, MV. The potential spread of infection caused by aerosol contamination of surfaces after flushing a domestic toilet. J Appl Microbiol. (2005) 99:339–47. doi: 10.1111/j.1365-2672.2005.02610.x

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Cho, SJ, Cox-Ganser, JM, and Park, J-H. Observational scores of dampness and mold associated with measurements of microbial agents and moisture in three public schools. Indoor Air. (2016) 26:168–78. doi: 10.1111/ina.12191

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Solomon, GM, Hjelmroos-Koski, M, Rotkin-Ellman, M, and Hammond, SK. Airborne Mold and endotoxin concentrations in New Orleans, Louisiana, after flooding, October through November 2005. Environ Health Perspect. (2006) 114:1381–6. doi: 10.1289/ehp.9198

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Kwan, SE, Shaughnessy, R, Haverinen-Shaughnessy, U, Kwan, TA, and Peccia, J. The impact of ventilation rate on the fungal and bacterial ecology of home indoor air. Build Environ. (2020) 177:106800. doi: 10.1016/j.buildenv.2020.106800

CrossRef Full Text | Google Scholar

54. Stanley, HO, Onwuna, B, and Ugboma, CJ. Microbial assessment of indoor air quality of ventilation systems. Asian J Adv Res Rep. (2019) 2019:1–7. doi: 10.9734/ajarr/2019/v3i430095

CrossRef Full Text | Google Scholar

55. Lee, BU, Yun, SH, Ji, J-H, and Bae, G-N. Inactivation of S. epidermidis, B. subtilis, and E. coli bacteria bioaerosols deposited on a filter utilizing airborne silver nanoparticles. J Microbiol Biotechnol. (2008) 18:176–82.

PubMed Abstract | Google Scholar

56. Hassan Al-abdalall, A, Abdullah Al-dakheel, S, and Abdulhadi, A-AH. Impact of air-conditioning filters on microbial growth and indoor air pollution In: T Morosuk and M Sultan, editors. Low-temperature technologies. London: IntechOpen (2020)

Google Scholar

57. Mahnert, A, Moissl-Eichinger, C, and Berg, G. Microbiome interplay: plants alter microbial abundance and diversity within the built environment. Front Microbiol. (2015) 6:887. doi: 10.3389/fmicb.2015.00887

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Berg, G, Mahnert, A, and Moissl-Eichinger, C. Beneficial effects of plant-associated microbes on indoor microbiomes and human health? Front Microbiol. (2014) 5:15. doi: 10.3389/fmicb.2014.00015

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Wolverton, BC, and Wolverton, JD. Interior plants: their influence on airborne microbes inside energy-efficient buildings. J Miss Acad Sci. (1996) 41:100–5.

Google Scholar

60. Burge, HA, Solomon, WR, and Muilenberg, ML. Evaluation of indoor plantings as allergen exposure sources. J Allergy Clin Immunol. (1982) 70:101–8. doi: 10.1016/0091-6749(82)90236-6

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Napoli, C, Marcotrigiano, V, and Montagna, MT. Air sampling procedures to evaluate microbial contamination: a comparison between active and passive methods in operating theatres. BMC Public Health. (2012) 12:594. doi: 10.1186/1471-2458-12-594

CrossRef Full Text | Google Scholar

62. Saffell, J, and Nehr, S. Improving indoor air quality through standardization. Standards. (2023) 3:240–67. doi: 10.3390/standards3030019

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Viani, I, Colucci, ME, Pergreffi, M, Rossi, D, Veronesi, L, Bizzarro, A, et al. Passive air sampling: the use of the index of microbial air contamination. Acta Biomed. (2020) 91:92–105. doi: 10.23750/abm.v91i3-S.9434

CrossRef Full Text | Google Scholar

64. Kotgire, S, Akhtar, R, Damle, A, Siddiqui, S, and Padekar, H. Bioaerosol assessment of indoor air in hospital wards from a tertiary care hospital. Indian J Microbiol Res. (2020) 7:28–34. doi: 10.18231/j.ijmr.2020.007

CrossRef Full Text | Google Scholar

65. Haig, CW, Mackay, WG, Walker, JT, and Williams, C. Bioaerosol sampling: sampling mechanisms, bioefficiency and field studies. J Hosp Infect. (2016) 93:242–55. doi: 10.1016/j.jhin.2016.03.017

CrossRef Full Text | Google Scholar

66. French, MLV, Eitzen, HE, Ritter, MA, and Leland, DS. Wound healing and wound infection In: TK Hunt, editor. Environmental control of microbial contamination in the operating room. New York: Appleton-Century Crois (1980). 254–61.

Google Scholar

67. Wurtz, H, Sigsgaard, T, Valbjorn, O, Doekes, G, and Meyer, HW. The dustfall collector - a simple passive tool for long-term collection of airborne dust: a project under the Danish Mould in buildings program (DAMIB). Indoor Air. (2005) 15:33–40. doi: 10.1111/j.1600-0668.2005.00342.x

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Manibusan, S, and Mainelis, G. Passive bioaerosol samplers: a complementary tool for bioaerosol research. A review. J Aerosol Sci. (2022) 163:105992. doi: 10.1016/j.jaerosci.2022.105992

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Therkorn, J, Thomas, N, Scheinbeim, J, and Mainelis, G. Field performance of a novel passive bioaerosol sampler using polarized ferroelectric polymer films. Aerosol Sci Technol. (2017) 51:787–800. doi: 10.1080/02786826.2017.1316830

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Cox, J, Indugula, R, Vesper, S, Zhu, Z, Jandarov, R, and Reponen, T. Comparison of indoor air sampling and dust collection methods for fungal exposure assessment using quantitative PCR. Environ Sci Process Impacts. (2017) 19:1312–9. doi: 10.1039/C7EM00257B

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Saha, R, Agarawal, S, and Khan, AM. Air sampling procedures to evaluate microbial contamination: a comparison between active and passive methods at high-risk areas in a tertiary care hospital of Delhi. J Patient Saf Infect Control. (2017) 5:18–23. doi: 10.4103/jpsic.jpsic_12_17

CrossRef Full Text | Google Scholar

72. Massoudinejad, MR, Ghajari, A, Hezarkhani, N, and Aliyari, A. Survey of fungi bioaersols in ICU ward of Taleghani hospital in Tehran by petri-dish trapping technique and bioaerosol sampler in 2013. J Saf Promot Injury Prevent. (2015) 3:147–54.

Google Scholar

73. Petti, S, Iannazzo, S, and Tarsitani, G. Comparison between different methods to monitor the microbial level of indoor air contamination in the dental office. Ann Ig. (2003) 15:725–33.

PubMed Abstract | Google Scholar

74. Balyan, P, Ghosh, C, Das, S, and Banerjee, BD. Comparison of efficiency of active and passive methods of bioaerosols’ estimation In: P Balyan, editor. Lecture notes in civil engineering. Berlin: Springer (2020). 85–94. doi: 10.1007/978-981-15-1334-3_10

CrossRef Full Text | Google Scholar

75. Canha, N, Almeida, SM, Freitas Do, CM, and Wolterbeek, HT. Assessment of bioaerosols in urban and rural primary schools using passive and active sampling methodologies. Archives of. Environ Prot. (2015) 41:11–22. doi: 10.1515/aep-2015-0034

CrossRef Full Text | Google Scholar

76. Sudharsanam, S, Swaminathan, S, Ramalingam, A, Thangavel, G, Annamalai, R, Steinberg, R, et al. Characterization of indoor bioaerosols from a hospital ward in a tropical setting. Afr Health Sci. (2012) 12:217–25. doi: 10.4314/ahs.v12i2.22

CrossRef Full Text | Google Scholar

77. Dallongeville, A, le Cann, P, Zmirou-Navier, D, Chevrier, C, Costet, N, Annesi-Maesano, I, et al. Concentration and determinants of molds and allergens in indoor air and house dust of French dwellings. Sci Total Environ. (2015) 536:964–72. doi: 10.1016/j.scitotenv.2015.06.039

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Pasquarella, C, Albertini, R, Dall’aglio, P, Saccani, E, Sansebastiano, GE, and Signorelli, C. Air microbial sampling: the state of the art. Ig Sanita Pubbl. (2008) 64:79–120.

PubMed Abstract | Google Scholar

79. Frankel, M, Timm, M, Hansen, EW, and Madsen, AM. Comparison of sampling methods for the assessment of indoor microbial exposure. Indoor Air. (2012) 22:405–14. doi: 10.1111/j.1600-0668.2012.00770.x

CrossRef Full Text | Google Scholar

80. Jensen, PA, and Schafer, MP. Sampling and characterization of bioaerosols. In: PA Jensen. NIOSH Manual of Analytical Methods. Atlanta, Georgia: CDC pp. 82–112. (1998).

Google Scholar

81. Mandal, J, and Brandl, H. Bioaerosols in indoor environment–a review with special reference to residential and occupational locations. Open Environ Biol Monit J. (2011) 4:83–96. doi: 10.2174/1875040001104010083

CrossRef Full Text | Google Scholar

82. Kalogerakis, N, Paschali, D, Lekaditis, V, Pantidou, A, Eleftheriadis, K, and Lazaridis, M. Indoor air quality—bioaerosol measurements in domestic and office premises. J Aerosol Sci. (2005) 36:751–61. doi: 10.1016/j.jaerosci.2005.02.004

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Prigione, V, Lingua, G, and Marchisio, VF. Development and use of flow cytometry for detection of airborne Fungi. Appl Environ Microbiol. (2004) 70:1360–5. doi: 10.1128/AEM.70.3.1360-1365.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Hugenholtz, P. Exploring prokaryotic diversity in the genomic era. Genome Biol. (2002) 3:REVIEWS0003. doi: 10.1186/gb-2002-3-2-reviews0003

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Rahi, P, Prakash, O, and Shouche, YS. Matrix-assisted laser desorption/ionization time-of-flight mass-spectrometry (MALDI-TOF MS) based microbial identifications: challenges and scopes for microbial ecologists. Front Microbiol. (2016) 7:1359. doi: 10.3389/fmicb.2016.01359

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Singhal, N, Kumar, M, Kanaujia, PK, and Virdi, JS. MALDI-TOF mass spectrometry: an emerging technology for microbial identification and diagnosis. Front Microbiol. (2015) 6:791. doi: 10.3389/fmicb.2015.00791

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Ashfaq, MY, Da’na, DA, and Al-Ghouti, MA. Application of MALDI-TOF MS for identification of environmental bacteria: a review. J Environ Manag. (2022) 305:114359. doi: 10.1016/j.jenvman.2021.114359

PubMed Abstract | CrossRef Full Text | Google Scholar

88. Jang, K-S, and Kim, YH. Rapid and robust MALDI-TOF MS techniques for microbial identification: a brief overview of their diverse applications. J Microbiol. (2018) 56:209–16. doi: 10.1007/s12275-018-7457-0

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Fykse, EM, Tjärnhage, T, Humppi, T, Eggen, VS, Ingebretsen, A, Skogan, G, et al. Identification of airborne bacteria by 16S rDNA sequencing, MALDI-TOF MS and the MIDI microbial identification system. Aerobiologia (Bologna). (2015) 31:271–81. doi: 10.1007/s10453-015-9363-9

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Fox, K, Fox, A, Elßner, T, Feigley, C, and Salzberg, D. MALDI-TOF mass spectrometry speciation of staphylococci and their discrimination from micrococci isolated from indoor air of schoolrooms. J Environ Monit. (2010) 12:917–23. doi: 10.1039/b925250a

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Motteu, N, Goemaere, B, Bladt, S, and Packeu, A. Implementation of MALDI-TOF mass spectrometry to identify Fungi from the indoor environment as an added value to the classical morphology-based identification tool. Front Allergy. (2022) 3:826148. doi: 10.3389/falgy.2022.826148

PubMed Abstract | CrossRef Full Text | Google Scholar

92. Fenselau, CC. Rapid characterization of microorganisms by mass spectrometry–what can be learned and how? J Am Soc Mass Spectrom. (2013) 24:1161–6. doi: 10.1007/s13361-013-0660-7

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Geng, Z, Yang, W, Wei, S, Wang, L, and Pan, J. Verification of real-time microbial monitoring technology based on the principle of laser-induced fluorescence under USP <1223> standard and prospective application in our hospital PIVAS clean room. Int J Pept Res Ther. (2022) 28:4. doi: 10.1007/s10989-021-10315-8

CrossRef Full Text | Google Scholar

94. Saari, S, Järvinen, S, Reponen, T, Mensah-Attipoe, J, Pasanen, P, Toivonen, J, et al. Identification of single microbial particles using electro-dynamic balance assisted laser-induced breakdown and fluorescence spectroscopy. Aerosol Sci Technol. (2016) 50:126–32. doi: 10.1080/02786826.2015.1134764

CrossRef Full Text | Google Scholar

95. Babichenko, S, Gala, J, Bentahir, M, Piette, A, Poryvkina, L, Rebane, O, et al. Real-time laser-induced fluorescence detection and monitoring of microbial contaminants on solid surfaces before, during and after decontamination. J Biosens Bioelectron. (2018) 9:255. doi: 10.4172/2155-6210.1000255

CrossRef Full Text | Google Scholar

96. Leone, N, D’Arthur, G, Adam, P, and Amouroux, J. Detection of bacterial deposits and bio aerosols by time-resolved laser-induced breakdown spectroscopy (trelibs). High Temp Mater Proc. (2004) 8:1–22. doi: 10.1615/HighTempMatProc.v8.i1.10

CrossRef Full Text | Google Scholar

97. Gabbarini, V, Rossi, R, Ciparisse, J-F, Malizia, A, Divizia, A, de Filippis, P, et al. Laser-induced fluorescence (LIF) as a smart method for fast environmental virological analyses: validation on picornaviruses. Sci Rep. (2019) 9:12598. doi: 10.1038/s41598-019-49005-3

CrossRef Full Text | Google Scholar

98. Owoicho, O, Olwal, CO, and Quaye, O. Potential of laser-induced fluorescence-light detection and ranging for future stand-off virus surveillance. Microb Biotechnol. (2021) 14:126–35. doi: 10.1111/1751-7915.13698

PubMed Abstract | CrossRef Full Text | Google Scholar

99. Singh, VK, Sharma, J, Pathak, AK, Ghany, CT, and Gondal, MA. Laser-induced breakdown spectroscopy (LIBS): a novel technology for identifying microbes causing infectious diseases. Biophys Rev. (2018) 10:1221–39. doi: 10.1007/s12551-018-0465-9

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Saari, S, Putkiranta, MJ, and Keskinen, J. Fluorescence spectroscopy of atmospherically relevant bacterial and fungal spores and potential interferences. Atmos Environ. (2013) 71:202–9. doi: 10.1016/j.atmosenv.2013.02.023

CrossRef Full Text | Google Scholar

101. Prussin, AJ, Garcia, EB, and Marr, LC. Total concentrations of virus and Bacteria in indoor and outdoor air. Environ Sci Technol Lett. (2015) 2:84–8. doi: 10.1021/acs.estlett.5b00050

CrossRef Full Text | Google Scholar

102. Li, C-S, and Huang, T-Y. Fluorochrome in monitoring indoor bioaerosols. Aerosol Sci Technol. (2006) 40:237–41. doi: 10.1080/02786820500543308

PubMed Abstract | CrossRef Full Text | Google Scholar

103. Fabian, MP, Miller, SL, Reponen, T, and Hernandez, MT. Ambient bioaerosol indices for indoor air quality assessments of flood reclamation. J Aerosol Sci. (2005) 36:763–83. doi: 10.1016/j.jaerosci.2004.11.018

CrossRef Full Text | Google Scholar

104. Chi, M-C, and Li, C-S. Fluorochrome and fluorescent in situ hybridization to monitor bioaerosols in swine buildings. Aerosol Sci Technol. (2005) 39:1101–10. doi: 10.1080/02786820500421539

CrossRef Full Text | Google Scholar

105. Bergquist, PL, Hardiman, EM, Ferrari, BC, and Winsley, T. Applications of flow cytometry in environmental microbiology and biotechnology. Extremophiles. (2009) 13:389–401. doi: 10.1007/s00792-009-0236-4

PubMed Abstract | CrossRef Full Text | Google Scholar

106. Lange, JL, Thorne, PS, and Lynch, N. Application of flow cytometry and fluorescent in situ hybridization for assessment of exposures to airborne bacteria. Appl Environ Microbiol. (1997) 63:1557–63. doi: 10.1128/aem.63.4.1557-1563.1997

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Wang, Y, Hammes, F, de Roy, K, Verstraete, W, and Boon, N. Past, present and future applications of flow cytometry in aquatic microbiology. Trends Biotechnol. (2010) 28:416–24. doi: 10.1016/j.tibtech.2010.04.006

PubMed Abstract | CrossRef Full Text | Google Scholar

108. Chen, P-S, and Li, C-S. Bioaerosol characterization by flow cytometry with fluorochrome. J Environ Monit. (2005) 7:950–9. doi: 10.1039/b505224f

PubMed Abstract | CrossRef Full Text | Google Scholar

109. Nimra, A, Ali, Z, Sultan, S, Nasir, ZA, Sidra, S, and Hussain, A. Molecular sequencing and phylogenetic analysis of bioaerosols in hospital wards with different ventilation conditions. J Infect Dev Countr. (2022) 16:157–65. doi: 10.3855/jidc.13682

CrossRef Full Text | Google Scholar

110. Jayan, H, Pu, H, and Sun, D-W. Recent developments in Raman spectral analysis of microbial single cells: techniques and applications. Crit Rev Food Sci Nutr. (2022) 62:4294–308. doi: 10.1080/10408398.2021.1945534

PubMed Abstract | CrossRef Full Text | Google Scholar

111. Sengupta, A, Laucks, ML, Dildine, N, Drapala, E, and Davis, EJ. Bioaerosol characterization by surface-enhanced Raman spectroscopy (SERS). J Aerosol Sci. (2005) 36:651–64. doi: 10.1016/j.jaerosci.2004.11.001

CrossRef Full Text | Google Scholar

112. Witkowska, E, Jagielski, T, Kamińska, A, Kowalska, A, Hryncewicz-Gwóźdź, A, and Waluk, J. Detection and identification of human fungal pathogens using surface-enhanced Raman spectroscopy and principal component analysis. Anal Methods. (2016) 8:8427–34. doi: 10.1039/C6AY02957D

PubMed Abstract | CrossRef Full Text | Google Scholar

113. Jarvis, RM, Brooker, A, and Goodacre, R. Surface-enhanced Raman spectroscopy for bacterial discrimination utilizing a scanning Electron microscope with a Raman spectroscopy Interface. Anal Chem. (2004) 76:5198–202. doi: 10.1021/ac049663f

CrossRef Full Text | Google Scholar

114. Schwarzmeier, K, Knauer, M, Ivleva, NP, Niessner, R, and Haisch, C. Bioaerosol analysis based on a label-free microarray readout method using surface-enhanced Raman scattering. Anal Bioanal Chem. (2013) 405:5387–92. doi: 10.1007/s00216-013-6984-0

PubMed Abstract | CrossRef Full Text | Google Scholar

115. Ghosal, S, Macher, JM, and Ahmed, K. Raman microspectroscopy-based identification of individual fungal spores as potential indicators of indoor contamination and moisture-related building damage. Environ Sci Technol. (2012) 46:6088–95. doi: 10.1021/es203782j

PubMed Abstract | CrossRef Full Text | Google Scholar

116. Su, C, Lei, L, Duan, Y, Zhang, K-Q, and Yang, J. Culture-independent methods for studying environmental microorganisms: methods, application, and perspective. Appl Microbiol Biotechnol. (2012) 93:993–1003. doi: 10.1007/s00253-011-3800-7

PubMed Abstract | CrossRef Full Text | Google Scholar

117. Kuypers, MMM, and Jørgensen, BB. The future of single-cell environmental microbiology. Environ Microbiol. (2007) 9:6–7. doi: 10.1111/j.1462-2920.2006.01222_5.x

PubMed Abstract | CrossRef Full Text | Google Scholar

118. Adams, RI, Bateman, AC, Bik, HM, and Meadow, JF. Microbiota of the indoor environment: a meta-analysis. Microbiome. (2015) 3:49. doi: 10.1186/s40168-015-0108-3

CrossRef Full Text | Google Scholar

119. Stephens, B. What have we learned about the microbiomes of indoor environments? mSystems. (2016) 1:16. doi: 10.1128/mSystems.00083-16

PubMed Abstract | CrossRef Full Text | Google Scholar

120. Georgakopoulos, DG, Després, V, Fröhlich-Nowoisky, J, Psenner, R, Ariya, PA, Pósfai, M, et al. Microbiology and atmospheric processes: biological, physical and chemical characterization of aerosol particles. Biogeosciences. (2009) 6:721–37. doi: 10.5194/bg-6-721-2009

PubMed Abstract | CrossRef Full Text | Google Scholar

121. Leung, MHY, Wilkins, D, Li, EKT, Kong, FKF, and Lee, PKH. Indoor-air microbiome in an urban Subway network: diversity and dynamics. Appl Environ Microbiol. (2014) 80:6760–70. doi: 10.1128/AEM.02244-14

CrossRef Full Text | Google Scholar

122. Smith, CJ, and Osborn, AM. Advantages and limitations of quantitative PCR (Q-PCR)-based approaches in microbial ecology. FEMS Microbiol Ecol. (2009) 67:6–20. doi: 10.1111/j.1574-6941.2008.00629.x

PubMed Abstract | CrossRef Full Text | Google Scholar

123. Kärkkäinen, PM, Valkonen, M, Hyvärinen, A, Nevalainen, A, and Rintala, H. Determination of bacterial load in house dust using qPCR, chemical markers and culture. J Environ Monit. (2010) 12:759. doi: 10.1039/b917937b

PubMed Abstract | CrossRef Full Text | Google Scholar

124. Garibyan, L, and Avashia, N. Polymerase chain reaction. J Investig Dermatol. (2013) 133:1–4. doi: 10.1038/jid.2013.1

CrossRef Full Text | Google Scholar

125. Nocker, A, Burr, M, and Camper, AK. Genotypic microbial community profiling: a critical technical review. Microb Ecol. (2007) 54:276–89. doi: 10.1007/s00248-006-9199-5

CrossRef Full Text | Google Scholar

126. Weikl, F, Tischer, C, Probst, AJ, Heinrich, J, Markevych, I, Jochner, S, et al. Fungal and bacterial communities in indoor dust follow different environmental determinants. PLoS One. (2016) 11:4131. doi: 10.1371/journal.pone.0154131

PubMed Abstract | CrossRef Full Text | Google Scholar

127. Thomas, T, Gilbert, J, and Meyer, F. Metagenomics–a guide from sampling to data analysis. Microb Inform Exp. (2012) 2:3. doi: 10.1186/2042-5783-2-3

CrossRef Full Text | Google Scholar

128. Nygaard, AB, Tunsjø, HS, Meisal, R, and Charnock, C. A preliminary study on the potential of Nanopore MinION and Illumina MiSeq 16S rRNA gene sequencing to characterize building-dust microbiomes. Sci Rep. (2020) 10:3209. doi: 10.1038/s41598-020-59771-0

CrossRef Full Text | Google Scholar

129. Madsen, AM, Zervas, A, Tendal, K, and Nielsen, JL. Microbial diversity in bioaerosol samples causing ODTS compared to reference bioaerosol samples as measured using Illumina sequencing and MALDI-TOF. Environ Res. (2015) 140:255–67. doi: 10.1016/j.envres.2015.03.027

PubMed Abstract | CrossRef Full Text | Google Scholar

130. Nazzal, Y, Howari, FM, Yaslam, A, Iqbal, J, Maloukh, L, Ambika, LK, et al. A methodological review of tools that assess dust microbiomes, Metatranscriptomes and the particulate chemistry of indoor dust. Atmosphere (Basel). (2022) 13:1276. doi: 10.3390/atmos13081276

CrossRef Full Text | Google Scholar

131. Gangneux, J-P, Sassi, M, Lemire, P, and Le Cann, P. Metagenomic characterization of indoor dust bacterial and fungal microbiota in homes of asthma and non-asthma patients using next generation sequencing. Front Microbiol. (2020) 11:1671. doi: 10.3389/fmicb.2020.01671

PubMed Abstract | CrossRef Full Text | Google Scholar

132. Shin, S-K, Kim, J, Ha, S, Oh, H-S, Chun, J, Sohn, J, et al. Metagenomic insights into the bioaerosols in the indoor and outdoor environments of childcare facilities. PLoS One. (2015) 10:e0126960. doi: 10.1371/journal.pone.0126960

PubMed Abstract | CrossRef Full Text | Google Scholar

133. Rosario, K, Fierer, N, Miller, S, Luongo, J, and Breitbart, M. Diversity of DNA and RNA viruses in indoor air as assessed via metagenomic sequencing. Environ Sci Technol. (2018) 52:1014–27. doi: 10.1021/acs.est.7b04203

PubMed Abstract | CrossRef Full Text | Google Scholar

134. Ko, KKK, Chng, KR, and Nagarajan, N. Metagenomics-enabled microbial surveillance. Nat Microbiol. (2022) 7:486–96. doi: 10.1038/s41564-022-01089-w

PubMed Abstract | CrossRef Full Text | Google Scholar

135. Behzad, H, Gojobori, T, and Mineta, K. Challenges and opportunities of airborne metagenomics. Genome Biol Evol. (2015) 7:1216–26. doi: 10.1093/gbe/evv064

PubMed Abstract | CrossRef Full Text | Google Scholar

136. Cappitelli, F, Fermo, P, Vecchi, R, Piazzalunga, A, Valli, G, Zanardini, E, et al. Chemical–physical and microbiological measurements for indoor air quality assessment at the ca’ Granda historical archive, Milan (Italy). Water Air Soil Pollut. (2009) 201:109–20. doi: 10.1007/s11270-008-9931-5

CrossRef Full Text | Google Scholar

137. Fsadni, P, Frank, B, Fsadni, C, and Montefort, S. The impact of microbiological pollutants on school indoor air quality. J Geosci Environ Prot. (2017) 5:54–65. doi: 10.4236/gep.2017.55004

PubMed Abstract | CrossRef Full Text | Google Scholar

138. Montagna, MT, De Giglio, O, Cristina, ML, Napoli, C, Pacifico, C, Agodi, A, et al. Evaluation of Legionella air contamination in healthcare facilities by different sampling methods: An Italian multicenter study. Int J Environ Res Public Health. (2017) 14:670. doi: 10.3390/ijerph14070670

PubMed Abstract | CrossRef Full Text | Google Scholar

139. Solomon, FB, Wadilo, F, Tufa, EG, and Mitiku, M. Extended spectrum and metalo beta-lactamase producing airborne Pseudomonas aeruginosa and Acinetobacter baumanii in restricted settings of a referral hospital: a neglected condition. Antimicrob Resist Infect Control. (2017) 6:106. doi: 10.1186/s13756-017-0266-0

CrossRef Full Text | Google Scholar

140. Asif, A, Zeeshan, M, Hashmi, I, Zahid, U, and Bhatti, MF. Microbial quality assessment of indoor air in a large hospital building during winter and spring seasons. Build Environ. (2018) 135:68–73. doi: 10.1016/j.buildenv.2018.03.010

CrossRef Full Text | Google Scholar

141. Brągoszewska, E, and Biedroń, I. Indoor air quality and potential health risk impacts of exposure to antibiotic resistant Bacteria in an office rooms in southern Poland. Int J Environ Res Public Health. (2018) 15:2604. doi: 10.3390/ijerph15112604

CrossRef Full Text | Google Scholar

142. Choi, JY, Zemke, J, Philo, SE, Bailey, ES, Yondon, M, and Gray, GC. Aerosol sampling in a hospital emergency room setting: a complementary surveillance method for the detection of respiratory viruses. Front Public Health. (2018) 6:174. doi: 10.3389/fpubh.2018.00174

PubMed Abstract | CrossRef Full Text | Google Scholar

143. Kozdrój, J, Frączek, K, and Ropek, D. Assessment of bioaerosols in indoor air of glasshouses located in a botanical garden. Build Environ. (2019) 166:106436. doi: 10.1016/j.buildenv.2019.106436

CrossRef Full Text | Google Scholar

144. Sornboot, J, Aekplakorn, W, Ramasoota, P, Bualert, S, Tumwasorn, S, and Jiamjarasrangsi, W. Detection of airborne Mycobacterium tuberculosis complex in high-risk areas of health care facilities in Thailand. Int J Tuberc Lung Dis. (2019) 23:465–73. doi: 10.5588/ijtld.18.0218

CrossRef Full Text | Google Scholar

145. Awad, AHA, Saeed, Y, Shakour, AA, Abdellatif, NM, Ibrahim, YH, Elghanam, M, et al. Indoor air fungal pollution of a historical museum, Egypt: a case study. Aerobiologia. (2020) 36:197–209. doi: 10.1007/s10453-019-09623-w

CrossRef Full Text | Google Scholar

146. Vornanen-Winqvist, C, Järvi, K, Andersson, MA, Duchaine, C, Létourneau, V, Kedves, O, et al. Exposure to indoor air contaminants in school buildings with and without reported indoor air quality problems. Environ Int. (2020) 141:105781. doi: 10.1016/j.envint.2020.105781

CrossRef Full Text | Google Scholar

147. Baudet, A, Baurès, E, Guegan, H, Blanchard, O, Guillaso, M, Le Cann, P, et al. Indoor air quality in healthcare and care facilities: chemical pollutants and microbiological contaminants. Atmosphere (Basel). (2021) 12:1337. doi: 10.3390/atmos12101337

CrossRef Full Text | Google Scholar

148. Cox, J, Stone, T, Ryan, P, Burkle, J, Jandarov, R, Mendell, MJ, et al. Associations of observed home dampness and mold with the fungal and bacterial dust microbiomes. Environ Sci Process Impacts. (2021) 23:491–500. doi: 10.1039/D0EM00505C

CrossRef Full Text | Google Scholar

149. Habibi, N, Uddin, S, Al-Salameen, F, Al-Amad, S, Kumar, V, Al-Otaibi, M, et al. SARS-CoV-2, other respiratory viruses and bacteria in aerosols: Report from Kuwait’s hospitals. Indoor Air. (2021) 31:1815–25. doi: 10.1111/ina.12871

PubMed Abstract | CrossRef Full Text | Google Scholar

150. Lu, R, Pørneki, AD, Lindgreen, JN, Li, Y, and Madsen, AM. Species of Fungi and pollen in the PM1 and the inhalable fraction of indoor air in homes. Atmosphere (Basel). (2021) 12:404. doi: 10.3390/atmos12030404

CrossRef Full Text | Google Scholar

151. Sibanda, T, Selvarajan, R, Ogola, HJ, Obieze, CC, and Tekere, M. Distribution and comparison of bacterial communities in HVAC systems of two university buildings: implications for indoor air quality and public health. Environ Monit Assess. (2021) 193:47. doi: 10.1007/s10661-020-08823-z

CrossRef Full Text | Google Scholar

152. Vosoughi, M, Karami, C, Dargahi, A, Jeddi, F, Jalali, KM, Hadisi, A, et al. Investigation of SARS-CoV-2 in hospital indoor air of COVID-19 patients’ ward with impinger method. Environ Sci Pollut Res. (2021) 28:50480–8. doi: 10.1007/s11356-021-14260-3

PubMed Abstract | CrossRef Full Text | Google Scholar

153. Zavieh, FS, Mohammadi, MJ, Vosoughi, M, Abazari, M, Raesee, E, Fazlzadeh, M, et al. Assessment of types of bacterial bio-aerosols and concentrations in the indoor air of gyms. Environ Geochem Health. (2021) 43:2165–73. doi: 10.1007/s10653-020-00774-1

CrossRef Full Text | Google Scholar

154. Maphossa, V, Langa, JC, Simbine, S, Maússe, FE, Kenga, D, Relvas, V, et al. Environmental bacterial and fungal contamination in high touch surfaces and indoor air of a paediatric intensive care unit in Maputo central hospital, Mozambique in 2018. Infect Prevent Pract. (2022) 4:100250. doi: 10.1016/j.infpip.2022.100250

PubMed Abstract | CrossRef Full Text | Google Scholar

155. An, X-L, Xu, J-X, Xu, M-R, Zhao, C-X, Li, H, Zhu, Y-G, et al. Dynamics of microbial community and potential microbial pollutants in shopping malls. mSystems. (2023) 8:e0057622. doi: 10.1128/msystems.00576-22

PubMed Abstract | CrossRef Full Text | Google Scholar

156. Kim, J, Bae, S, Park, S, Shukla, SK, and Yoo, K. Investigation of surface bacterial diversities and compositions in the global Subway facilities. Atmosphere (Basel). (2023) 14:140. doi: 10.3390/atmos14010140

CrossRef Full Text | Google Scholar

157. Rintala, H, Pitkaranta, M, Toivola, M, Paulin, L, and Nevalainen, A. Diversity and seasonal dynamics of bacterial community in indoor environment. BMC Microbiol. (2008) 8:56. doi: 10.1186/1471-2180-8-56

CrossRef Full Text | Google Scholar

158. Bonadonna, L, Briancesco, R, Coccia, AM, Meloni, P, La Rosa, G, and Moscato, U. Microbial air quality in healthcare facilities. Int J Environ Res Public Health. (2021) 18:6226. doi: 10.3390/ijerph18126226

CrossRef Full Text | Google Scholar

159. Chang, CW, Chou, FC, and Hung, PY. Evaluation of bioaerosol sampling techniques for Legionella pneumophila coupled with culture assay and quantitative PCR. J Aerosol Sci. (2010) 41:1055–65. doi: 10.1016/j.jaerosci.2010.09.002

CrossRef Full Text | Google Scholar

160. Spagnolo, AM, Cristina, ML, Casini, B, and Perdelli, F. Legionella pneumophila in healthcare facilities. Rev Med Microbiol. (2013) 24:70–80. doi: 10.1097/MRM.0b013e328362fe66

PubMed Abstract | CrossRef Full Text | Google Scholar

161. Zhang, BY, Liu, F, and Chen, XD. The effect of Legionella Pneumophila contamination in the surface dust of the air ducts of central air conditioning Systemson indoor air quality. Int J Vent. (2015) 14:231–40. doi: 10.1080/14733315.2015.11684083

CrossRef Full Text | Google Scholar

162. Lee, G, and Yoo, K. A review of the emergence of antibiotic resistance in bioaerosols and its monitoring methods. Rev Environ Sci Biotechnol. (2022) 21:799–827. doi: 10.1007/s11157-022-09622-3

CrossRef Full Text | Google Scholar

163. Bitew, K, Gidebo, DD, and Ali, MM. Bacterial contamination rates and drug susceptibility patterns of bacteria recovered from medical equipment, inanimate surfaces, and indoor air of a neonatal intensive care unit and pediatric ward at Hawassa university comprehensive specialized hospital, Ethiopia. IJID Reg. (2021) 1:27–33. doi: 10.1016/j.ijregi.2021.09.005

CrossRef Full Text | Google Scholar

164. Douwes, J, Thorne, P, Pearce, N, and Heederik, D. Bioaerosol health effects and exposure assessment: Progress and prospects. Ann Occup Hyg. (2003) 47:187–200. doi: 10.1093/annhyg/meg032

PubMed Abstract | CrossRef Full Text | Google Scholar

165. Li, X, Liu, D, and Yao, J. Aerosolization of fungal spores in indoor environments. Sci Total Environ. (2022) 820:153003. doi: 10.1016/j.scitotenv.2022.153003

CrossRef Full Text | Google Scholar

166. Liu, Z, Yin, H, Ma, S, Wei, B, Jensen, B, and Cao, G. Effect of environmental parameters on culturability and viability of dust accumulated fungi in different HVAC segments. Sustain Cities Soc. (2019) 48:101538. doi: 10.1016/j.scs.2019.101538

CrossRef Full Text | Google Scholar

167. Belizario, JA, Lopes, LG, and Pires, RH. Fungi in the indoor air of critical hospital areas: a review. Aerobiologia (Bologna). (2021) 37:379–94. doi: 10.1007/s10453-021-09706-7

PubMed Abstract | CrossRef Full Text | Google Scholar

168. Khalaf, NF, Al-Obaidi, MJ, Mohammed, SW, Al-Malkey, MK, Nayyef, HJ, Al-Hur, FJA, et al. Indoor house dust-borne fungi and risk of allergic respiratory diseases in Baghdad city. Rev Fr Allergol. (2022) 62:401–6. doi: 10.1016/j.reval.2021.05.002

CrossRef Full Text | Google Scholar

169. Kaarakainen, P, Rintala, H, Vepsäläinen, A, Hyvärinen, A, Nevalainen, A, and Meklin, T. Microbial content of house dust samples determined with qPCR. Sci Total Environ. (2009) 407:4673–80. doi: 10.1016/j.scitotenv.2009.04.046

PubMed Abstract | CrossRef Full Text | Google Scholar

170. Barker, J, Stevens, D, and Bloomfield, SF. Spread and prevention of some common viral infections in community facilities and domestic homes. J Appl Microbiol. (2001) 91:7–21. doi: 10.1046/j.1365-2672.2001.01364.x

PubMed Abstract | CrossRef Full Text | Google Scholar

171. López, A, Fuentes, E, Yusà, V, López-Labrador, F, Camaró, M, Peris-Martinez, C, et al. Indoor air quality including respiratory viruses. Toxics. (2021) 9:274. doi: 10.3390/toxics9110274

CrossRef Full Text | Google Scholar

172. Ciencewicki, J, and Jaspers, I. Air pollution and respiratory viral infection. Inhal Toxicol. (2007) 19:1135–46. doi: 10.1080/08958370701665434

CrossRef Full Text | Google Scholar

173. Clements, N, Binnicker, MJ, and Roger, VL. Indoor environment and viral infections. Mayo Clin Proc. (2020) 95:1581–3. doi: 10.1016/j.mayocp.2020.05.028

PubMed Abstract | CrossRef Full Text | Google Scholar

174. Myatt, TA, Kaufman, MH, Allen, JG, MacIntosh, DL, Fabian, MP, and McDevitt, JJ. Modeling the airborne survival of influenza virus in a residential setting: the impacts of home humidification. Environ Health. (2010) 9:55. doi: 10.1186/1476-069X-9-55

CrossRef Full Text | Google Scholar

175. Johnson, GR, and Morawska, L. The mechanism of breath aerosol formation. J Aerosol Med Pulm Drug Deliv. (2009) 22:229–37. doi: 10.1089/jamp.2008.0720

CrossRef Full Text | Google Scholar

176. Moon, KW, Huh, EH, and Jeong, HC. Seasonal evaluation of bioaerosols from indoor air of residential apartments within the metropolitan area in South Korea. Environ Monit Assess. (2014) 186:2111–20. doi: 10.1007/s10661-013-3521-8

PubMed Abstract | CrossRef Full Text | Google Scholar

177. Yang, W, Elankumaran, S, and Marr, LC. Concentrations 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. (2011) 8:1176–84. doi: 10.1098/rsif.2010.0686

PubMed Abstract | CrossRef Full Text | Google Scholar

178. Weinstein, RA, Weber, DJ, and Rutala, WA. Risks and prevention of nosocomial transmission of rare zoonotic diseases. Clin Infect Dis. (2001) 32:446–56. doi: 10.1086/318509

PubMed Abstract | CrossRef Full Text | Google Scholar

179. Yu, ITS, Li, Y, Wong, TW, Tam, W, Chan, AT, Lee, JHW, et al. Evidence of airborne transmission of the severe acute respiratory syndrome virus. N Engl J Med. (2004) 350:1731–9. doi: 10.1056/NEJMoa032867

PubMed Abstract | CrossRef Full Text | Google Scholar

180. Qian, H, Miao, T, Liu, L, Zheng, X, Luo, D, and Li, Y. Indoor transmission of SARS-CoV-2. Indoor Air. (2021) 31:639–45. doi: 10.1111/ina.12766

PubMed Abstract | CrossRef Full Text | Google Scholar

181. Norbäck, D, and Cai, GH. Microbial agents in the indoor environment: associations with health. Environmental Science, pp. 179–198. (2020).

Google Scholar

182. Kim, K-H, Kabir, E, and Jahan, SA. Airborne bioaerosols and their impact on human health. J Environ Sci. (2018) 67:23–35. doi: 10.1016/j.jes.2017.08.027

CrossRef Full Text | Google Scholar

183. Humbal, C, Gautam, S, and Trivedi, U. A review on recent progress in observations, and health effects of bioaerosols. Environ Int. (2018) 118:189–93. doi: 10.1016/j.envint.2018.05.053

PubMed Abstract | CrossRef Full Text | Google Scholar

184. Pomés, A, Chapman, MD, and Wünschmann, S. Indoor allergens and allergic respiratory disease. Curr Allergy Asthma Rep. (2016) 16:43. doi: 10.1007/s11882-016-0622-9

CrossRef Full Text | Google Scholar

185. Sun, Y, Meng, Y, Ou, Z, Li, Y, Zhang, M, Chen, Y, et al. Indoor microbiome, air pollutants and asthma, rhinitis and eczema in preschool children–a repeated cross-sectional study. Environ Int. (2022) 161:107137. doi: 10.1016/j.envint.2022.107137

CrossRef Full Text | Google Scholar

186. Brooks, C, Pearce, N, and Douwes, J. The hygiene hypothesis in allergy and asthma. Curr Opin Allergy Clin Immunol. (2013) 13:70–7. doi: 10.1097/ACI.0b013e32835ad0d2

PubMed Abstract | CrossRef Full Text | Google Scholar

187. Sham, NM, Ahmad, NI, Pahrol, MA, and Leong, Y-H. Fungus and mycotoxins studies in hospital environment: a scoping review. Build Environ. (2021) 193:107626. doi: 10.1016/j.buildenv.2021.107626

PubMed Abstract | CrossRef Full Text | Google Scholar

188. Oshikata, C, Watanabe, M, Saito, A, Ishida, M, Kobayashi, S, Konuma, R, et al. Allergic bronchopulmonary mycosis due to exposure to Eurotium herbariorum after the great East Japan earthquake. Prehosp Disaster Med. (2017) 32:688–90. doi: 10.1017/S1049023X17006914

CrossRef Full Text | Google Scholar

189. Sarkingobir, Y, Zayyanu, MM, and Sarkingobir, S. Sick building syndrome: a review of related literatures. Front Knowl J. (2017) 1:1–11.

Google Scholar

190. Wålinder, R, Ernstgård, L, Johanson, G, Norbäck, D, Venge, P, and Wieslander, G. Acute effects of a fungal volatile compound. Environ Health Perspect. (2005) 113:1775–8. doi: 10.1289/ehp.8193

PubMed Abstract | CrossRef Full Text | Google Scholar

191. Haleem Khan, AA, and Karuppayil, SM. Fungal pollution of indoor environments and its management. Saudi J Biol Sci. (2012) 19:405–26. doi: 10.1016/j.sjbs.2012.06.002

PubMed Abstract | CrossRef Full Text | Google Scholar

192. Nardell, EA. Indoor environmental control of tuberculosis and other airborne infections. Indoor Air. (2016) 26:79–87. doi: 10.1111/ina.12232

PubMed Abstract | CrossRef Full Text | Google Scholar

193. Solomon, FB, Wadilo, FW, Arota, AA, and Abraham, YL. Antibiotic resistant airborne bacteria and their multidrug resistance pattern at university teaching referral Hospital in South Ethiopia. Ann Clin Microbiol Antimicrob. (2017) 16:29. doi: 10.1186/s12941-017-0204-2

CrossRef Full Text | Google Scholar

194. Singh, J. Toxic Moulds and indoor air quality. Indoor Built Environ. (2005) 14:229–34. doi: 10.1177/1420326X05054015

PubMed Abstract | CrossRef Full Text | Google Scholar

195. Holý, O, Matoušková, I, Kubátová, A, Hamal, P, Svobodová, L, Jurásková, E, et al. Monitoring of microscopic filamentous Fungi in indoor air of transplant unit. Cent Eur J Public Health. (2015) 23:331–4. doi: 10.21101/cejph.a4062

CrossRef Full Text | Google Scholar

196. Couch, RB. Viruses and indoor air pollution. Bull N Y Acad Med. (1981) 57:907–21.

PubMed Abstract | Google Scholar

197. Li, Y, Huang, X, Yu, ITS, Wong, TW, and Qian, H. Role of air distribution in SARS transmission during the largest nosocomial outbreak in Hong Kong. Indoor Air. (2005) 15:83–95. doi: 10.1111/j.1600-0668.2004.00317.x

PubMed Abstract | CrossRef Full Text | Google Scholar

198. Ijaz, MK, Zargar, B, Wright, KE, Rubino, JR, and Sattar, SA. Generic aspects of the airborne spread of human pathogens indoors and emerging air decontamination technologies. Am J Infect Control. (2016) 44:S109–20. doi: 10.1016/j.ajic.2016.06.008

PubMed Abstract | CrossRef Full Text | Google Scholar

199. Allen, JG, and Ibrahim, AM. Indoor air changes and potential implications for SARS-CoV-2 transmission. JAMA. (2021) 325:2112. doi: 10.1001/jama.2021.5053

PubMed Abstract | CrossRef Full Text | Google Scholar

200. Hayleeyesus, SF, Ejeso, A, and Derseh, FA. Quantitative assessment of bio-aerosols contamination in indoor air of university dormitory rooms. Int J Health Sci (Qassim). (2015) 9:249–56.

PubMed Abstract | Google Scholar

201. Johnson, ES, and Choi, K-M. Lung Cancer risk in Workers in the Meat and Poultry Industries - a review. Zoonoses Public Health. (2012) 59:303–13. doi: 10.1111/j.1863-2378.2012.01459.x

PubMed Abstract | CrossRef Full Text | Google Scholar

202. Hoppin, JA, Umbach, DM, Long, S, Rinsky, JL, Henneberger, PK, Salo, PM, et al. Respiratory disease in United States farmers. Occup Environ Med. (2014) 71:484–91. doi: 10.1136/oemed-2013-101983

CrossRef Full Text | Google Scholar

203. Hendry, KM, and Cole, EC. A review of mycotoxins in indoor air. J Toxicol Environ Health. (1993) 38:183–98. doi: 10.1080/15287399309531711

CrossRef Full Text | Google Scholar

204. Claeys, L, Romano, C, De Ruyck, K, Wilson, H, Fervers, B, Korenjak, M, et al. Mycotoxin exposure and human cancer risk: a systematic review of epidemiological studies. Compr Rev Food Sci Food Saf. (2020) 19:1449–64. doi: 10.1111/1541-4337.12567

PubMed Abstract | CrossRef Full Text | Google Scholar

205. Bray, GA, and Ryan, DH, Pennington Biomedical Research Center. Mycotoxins, cancer, and health. Baton Rouge: Louisiana State University Press (1991).

Google Scholar

206. Jarvis, BB, and Miller, JD. Mycotoxins as harmful indoor air contaminants. Appl Microbiol Biotechnol. (2005) 66:367–72. doi: 10.1007/s00253-004-1753-9

PubMed Abstract | CrossRef Full Text | Google Scholar

207. Clark, HA, and Snedeker, SM. Ochratoxin a: its Cancer risk and potential for exposure. J Toxicol Environ Health B. (2006) 9:265–96. doi: 10.1080/15287390500195570

PubMed Abstract | CrossRef Full Text | Google Scholar

208. Verma, AK, and Prakash, S. Impact of COVID-19 on environment and society. J Glob Biosci. (2020) 9:7352–63.

Google Scholar

209. Nazir, R, Ali, J, Rasul, I, Widemann, E, and Shafiq, S. Eco-environmental aspects of COVID-19 pandemic and potential control strategies. Int J Environ Res Public Health. (2021) 18:3488. doi: 10.3390/ijerph18073488

PubMed Abstract | CrossRef Full Text | Google Scholar

210. Mahato, S, Pal, S, and Ghosh, KG. Effect of lockdown amid COVID-19 pandemic on air quality of the megacity Delhi, India. Sci Total Environ. (2020) 730:139086. doi: 10.1016/j.scitotenv.2020.139086

PubMed Abstract | CrossRef Full Text | Google Scholar

211. Shrestha, AM, Shrestha, UB, Sharma, R, Bhattarai, S, Ngoc, H, Tran, T, et al. Lockdown caused by COVID-19 pandemic reduces air pollution in cities worldwide. EarthArXiv. (2020) 2020:4. doi: 10.31223/osf.io/edt4j

CrossRef Full Text | Google Scholar

212. Ching, J, and Kajino, M. Rethinking air quality and climate change after COVID-19. Int J Environ Res Public Health. (2020) 17:5167. doi: 10.3390/ijerph17145167

PubMed Abstract | CrossRef Full Text | Google Scholar

213. Rani, R, Arokiasamy, P, and Sikarwar, A. Household air pollution during COVID-19 pandemic: a concern in India. J Public Aff. (2021) 21:e2729. doi: 10.1002/pa.2729

PubMed Abstract | CrossRef Full Text | Google Scholar

214. Du, W, and Wang, G. Indoor air pollution was nonnegligible during COVID-19 lockdown. Aerosol Air Qual Res. (2020) 20:1851–5. doi: 10.4209/aaqr.2020.06.0281

PubMed Abstract | CrossRef Full Text | Google Scholar

215. Zhang, Y, Zhao, B, Jiang, Y, Xing, J, Sahu, SK, Zheng, H, et al. Non-negligible contributions to human health from increased household air pollution exposure during the COVID-19 lockdown in China. Environ Int. (2022) 158:106918. doi: 10.1016/j.envint.2021.106918

PubMed Abstract | CrossRef Full Text | Google Scholar

216. Agarwal, N, Meena, CS, Raj, BP, Saini, L, Kumar, A, Gopalakrishnan, N, et al. Indoor air quality improvement in COVID-19 pandemic: review. Sustain Cities Soc. (2021) 70:102942. doi: 10.1016/j.scs.2021.102942

PubMed Abstract | CrossRef Full Text | Google Scholar

217. Lewis, D. Why the WHO took two years to say COVID is airborne. Nature. (2022) 604:26–31. doi: 10.1038/d41586-022-00925-7

CrossRef Full Text | Google Scholar

218. Setti, L, Passarini, F, de Gennaro, G, Barbieri, P, Licen, S, Perrone, MG, et al. Potential role of particulate matter in the spreading of COVID-19 in northern Italy: first observational study based on initial epidemic diffusion. BMJ Open. (2020) 10:e039338. doi: 10.1136/bmjopen-2020-039338

PubMed Abstract | CrossRef Full Text | Google Scholar

219. Schijven, J, Vermeulen, LC, Swart, A, Meijer, A, Duizer, E, and de Roda Husman, AM. Quantitative microbial risk assessment for airborne transmission of SARS-CoV-2 via breathing, speaking, singing, coughing, and sneezing. Environ Health Perspect. (2021) 129:047002. doi: 10.1289/EHP7886

PubMed Abstract | CrossRef Full Text | Google Scholar

220. Riediker, M, and Tsai, D-H. Estimation of viral aerosol emissions from simulated individuals with asymptomatic to moderate coronavirus disease 2019. JAMA Netw Open. (2020) 3:e2013807. doi: 10.1001/jamanetworkopen.2020.13807

PubMed Abstract | CrossRef Full Text | Google Scholar

221. Hosseini, MR, Fouladi-Fard, R, and Aali, R. COVID-19 pandemic and sick building syndrome. Indoor Built Environ. (2020) 29:1181–3. doi: 10.1177/1420326X20935644

CrossRef Full Text | Google Scholar

222. D’Alessandro, D, Gola, M, Appolloni, L, Dettori, M, Fara, GM, Rebecchi, A, et al. COVID-19 and living space challenge. Well-being and public health recommendations for a healthy, safe, and sustainable housing. Acta Biomed. (2020) 91:61–75. doi: 10.23750/abm.v91i9-S.10115

PubMed Abstract | CrossRef Full Text | Google Scholar

223. Ciric, L. Microbes in the built environment. Sci Rep. (2022) 12:8732. doi: 10.1038/s41598-022-12254-w

CrossRef Full Text | Google Scholar

224. Smith, KR, Samet, JM, Romieu, I, and Bruce, N. Indoor air pollution in developing countries and acute lower respiratory infections in children. Thorax. (2000) 55:518–32. doi: 10.1136/thorax.55.6.518

PubMed Abstract | CrossRef Full Text | Google Scholar

225. Bruce, N, Perez-Padilla, R, and Albalak, R. The health effects of indoor air pollution exposure in developing countries (no. WHO/SDE/OEH/02.05). (2002). Available at: https://apps.who.int/iris/handle/10665/67496.

Google Scholar

226. Jindal, SK, Aggarwal, AN, and Jindal, A. Household air pollution in India and respiratory diseases. Curr Opin Pulm Med. (2020) 26:128–34. doi: 10.1097/MCP.0000000000000642

PubMed Abstract | CrossRef Full Text | Google Scholar

227. Maharana, S, Paul, B, Garg, S, Dasgupta, A, and Bandyopadhyay, L. Exposure to indoor air pollution and its perceived impact on health of women and their children: a household survey in a slum of Kolkata, India. Indian J Public Health. (2018) 62:182–7. doi: 10.4103/ijph.IJPH_259_18

PubMed Abstract | CrossRef Full Text | Google Scholar

228. Roy, R, Jan, R, Joshi, U, Bhor, R, Pai, K, and Satsangi, PG. Characterization, pro-inflammatory response and cytotoxic profile of bioaerosols from urban and rural residential settings in Pune, India. Environ Pollut. (2020) 264:114698. doi: 10.1016/j.envpol.2020.114698

PubMed Abstract | CrossRef Full Text | Google Scholar

229. Central Pollution Control Board. National Ambient air Quality Status and Trends 2019 (NAAQMS/45/2019-2020). pp. 1–164. (2020).

Google Scholar

230. Manivannan, S, Venugopal, V, Kindo, AJ, and Ravinder, T. Role of residential indoor dampness and mold as determinants of respiratory symptoms among adults. Indian J Public Health Res Dev. (2019) 10:91–7. doi: 10.5958/0976-5506.2019.01245.2

CrossRef Full Text | Google Scholar

231. Dhar, DG, Dhar, P, Das, AK, and Uddin, N. Assessment of airborne Fungi in Childrens hospital located in Kolkata India. Def Life Sci J. (2021) 6:228–34. doi: 10.14429/dlsj.6.16731

CrossRef Full Text | Google Scholar

232. Kumar, P, Singh, AB, and Singh, R. Seasonal variation and size distribution in the airborne indoor microbial concentration of residential houses in Delhi and its impact on health. Aerobiologia (Bologna). (2021) 37:719–32. doi: 10.1007/s10453-021-09718-3

PubMed Abstract | CrossRef Full Text | Google Scholar

233. Bali, R, Sharma, P, Nagrath, S, and Gupta, P. Microbial isolations from maxillofacial operation theatre and its correlation to fumigation in a teaching Hospital in India. J Maxillofac Oral Surg. (2012) 13:128–32. doi: 10.1007/s12663-012-0458-3

PubMed Abstract | CrossRef Full Text | Google Scholar

234. Naruka, K, and Gaur, J. Microbial air contamination in a school. Int J Curr Microbiol Appl Sci. (2013) 2:404–10.

Google Scholar

235. Ghosh, B, Lal, H, Kushwaha, R, Hazarika, N, Srivastava, A, and Jain, VK. Estimation of bioaerosol in indoor environment in the university library of Delhi. Sustain Environ Res. (2013) 23:199–207.

Google Scholar

236. Kumar, B, Prakash Gupta, G, Singh, S, and Kulshrestha, UC. Study of abundance and characterization of Culturable bioaerosol at Delhi, India. Int J Environ Eng Manage. (2013) 4:219–26.

Google Scholar

237. Mohan, KNM, Ramprasad, S, and Maruthi, YA. Microbiological air quality of indoors in primary and secondary schools of Visakhapatnam, India. Int J Curr Microbiol Appl Sci. (2014) 3:880–7.

Google Scholar

238. Paul, D, Biswas, K, Sengupta, C, and Narayan, SS. Studies on environmental monitoring of Aeromicroflora in a Hospital at Kalyani, West Bengal, India. Front Environ Microbiol. (2015) 1:47–50. doi: 10.11648/j.fem.20150103.13

CrossRef Full Text | Google Scholar

239. Banerjee, B, Mukhopadhyay, C, Ke, V, Bupendra, A, and Varma, M. “News on air!” Air surveillance REPORT from intensive care units of a tertiary care hospital. Asian J Pharm Clin Res. (2016) 9:247. doi: 10.22159/ajpcr.2016.v9s3.14867

PubMed Abstract | CrossRef Full Text | Google Scholar

240. Meraj-Ul-Haque, BM, and Nitnawre, S. A study of fungal flora of house dust in Nagpur district, India. Mycopathologia. (2016) 14:43–5.

Google Scholar

241. Anjum, KMR, Sivakumari, K, Ashok, K, and Rajesh, S. A pilot investigation into associations between various indoor airborne fungal particles in Triplicane public toilet of Chennai, Tamil Nadu. Int J Curr Pharm Res. (2017) 9:50–3. doi: 10.22159/ijcpr.2017v9i3.18891

CrossRef Full Text | Google Scholar

242. Gaur, J, and Naruka, K. Assessment of airborne fungal pollution in a hospital room. Int Res J Biol Sci. (2017) 6:1–6.

Google Scholar

243. Quadir, R, Chauhan, PP, and Sofi, R. Airborne microbial population in different health care units of Dehrad. Bio Bulletin. (2017) 3:8–13.

Google Scholar

244. Kumar, H, Kaur, A, Kaur, B, Gupta, RK, Dhanjal, DS, Kaur, A, et al. Assessment of microbial contamination in indoor air of private maternity homes in Moga, Punjab. J Clin Diagn Res. (2018) 12:DC01–5. doi: 10.7860/JCDR/2018/34209.11480

CrossRef Full Text | Google Scholar

245. Pawar, A, Dhapte, V, Kalrao, V, Godha, I, Jacob, J, and Raut, A. Profiling of microbial contamination in internal atmosphere of hospital Ward. Res Pediatr Neonatol. (2018) 2:531. doi: 10.31031/RPN.2018.02.000531

CrossRef Full Text | Google Scholar

246. Priyamvada, H, Priyanka, C, Singh, RK, Akila, M, Ravikrishna, R, and Gunthe, SS. Assessment of PM and bioaerosols at diverse indoor environments in a southern tropical Indian region. Build Environ. (2018) 137:215–25. doi: 10.1016/j.buildenv.2018.04.016

CrossRef Full Text | Google Scholar

247. Sharma, R, Sharma, S, and Rampal, RK. Comparative assessment of indoor air quality within office environments with reference to indoor biopollutants, suspended particulate matter and relative humidity. JK Knowl Initiat. (2018) 2:54–63.

Google Scholar

248. Balyan, P, Ghosh, C, Das, S, and Banerjee, B. Spatio-temporal variations of indoor bioaerosols in different socio-economic zones of an urban Metropolis. Pol J Environ Stud. (2019) 28:4087–97. doi: 10.15244/pjoes/81272

CrossRef Full Text | Google Scholar

249. Raksha, SG. Operation theaters: a source of nosocomial infection. Era’s J Med Res. (2019) 6:09–13. doi: 10.24041/ejmr2019.103

CrossRef Full Text | Google Scholar

250. Valentina, Y, and Umadevi, S. Phenotypic detection and quality assessment of indoor air-borne microorganisms using passive air sampling technique (settle plate) at a tertiary care teaching Hospital in Puducherry. J Pure Appl Microbiol. (2019) 13:241–5. doi: 10.22207/JPAM.13.1.25

CrossRef Full Text | Google Scholar

251. Kumari, H, Chakraborti, T, Singh, M, Chakrawarti, MK, and Mukhopadhyay, K. Prevalence and antibiogram of coagulase negative staphylococci in bioaerosols from different indoors of a university in India. BMC Microbiol. (2020) 20:211. doi: 10.1186/s12866-020-01875-8

CrossRef Full Text | Google Scholar

252. Akila, M, Earappa, R, and Qureshi, A. Ambient concentration of airborne microbes and endotoxins in rural households of southern India. Build Environ. (2020) 179:106970. doi: 10.1016/j.buildenv.2020.106970

CrossRef Full Text | Google Scholar

253. Kumar, P, Singh, AB, and Singh, R. Comprehensive health risk assessment of microbial indoor air quality in microenvironments. PLoS One. (2022) 17:e0264226. doi: 10.1371/journal.pone.0264226

PubMed Abstract | CrossRef Full Text | Google Scholar

254. Balasubramanian, R, Nainar, P, and Rajasekar, A. Airborne bacteria, fungi, and endotoxin levels in residential microenvironments: a case study. Aerobiologia (Bologna). (2012) 28:375–90. doi: 10.1007/s10453-011-9242-y

CrossRef Full Text | Google Scholar

255. Szczotko, M, Orych, I, Mąka, Ł, and Solecka, J. A review of selected types of indoor air purifiers in terms of microbial air contamination reduction. Atmosphere (Basel). (2022) 13:800. doi: 10.3390/atmos13050800

CrossRef Full Text | Google Scholar

256. Qian, H, and Zheng, X. Ventilation control for airborne transmission of human exhaled bio-aerosols in buildings. J Thorac Dis. (2018) 10:S2295–304. doi: 10.21037/jtd.2018.01.24

PubMed Abstract | CrossRef Full Text | Google Scholar

257. Nair, AN, Anand, P, George, A, and Mondal, N. A review of strategies and their effectiveness in reducing indoor airborne transmission and improving indoor air quality. Environ Res. (2022) 213:113579. doi: 10.1016/j.envres.2022.113579

PubMed Abstract | CrossRef Full Text | Google Scholar

258. Centers for Disease Control and Prevention. Ventilation in buildings. (2023). Available at: https://www.cdc.gov/coronavirus/2019-ncov/community/ventilation.html (Accessed October 16, 2023).

Google Scholar

259. Walker, A. Natural Ventilation. National Renewable Energy Laboratory, Whole Building Design Guide. (2016). Available at: https://www.wbdg.org/resources/natural-ventilation (Accessed September 30, 2022).

Google Scholar

260. Qian, H, Li, Y, Seto, WH, Ching, P, Ching, WH, and Sun, HQ. Natural ventilation for reducing airborne infection in hospitals. Build Environ. (2010) 45:559–65. doi: 10.1016/j.buildenv.2009.07.011

CrossRef Full Text | Google Scholar

261. Luongo, JC, Fennelly, KP, Keen, JA, Zhai, ZJ, Jones, BW, and Miller, SL. Role of mechanical ventilation in the airborne transmission of infectious agents in buildings. Indoor Air. (2016) 26:666–78. doi: 10.1111/ina.12267

PubMed Abstract | CrossRef Full Text | Google Scholar

262. Escombe, AR, Oeser, CC, Gilman, RH, Navincopa, M, Ticona, E, Pan, W, et al. Natural ventilation for the prevention of airborne contagion. PLoS Med. (2007) 4:e68. doi: 10.1371/journal.pmed.0040068

PubMed Abstract | CrossRef Full Text | Google Scholar

263. Escombe, AR, Ticona, E, Chávez-Pérez, V, Espinoza, M, and Moore, DAJ. Improving natural ventilation in hospital waiting and consulting rooms to reduce nosocomial tuberculosis transmission risk in a low resource setting. BMC Infect Dis. (2019) 19:88. doi: 10.1186/s12879-019-3717-9

PubMed Abstract | CrossRef Full Text | Google Scholar

264. Vignolo, A, Gómez, AP, Draper, M, and Mendina, M. Quantitative assessment of natural ventilation in an elementary school classroom in the context of COVID-19 and its impact in airborne transmission. Appl Sci. (2022) 12:9261. doi: 10.3390/app12189261

CrossRef Full Text | Google Scholar

265. Núñez, A, and García, AM. Effect of the passive natural ventilation on the bioaerosol in a small room. Build Environ. (2022) 207:108438. doi: 10.1016/j.buildenv.2021.108438

CrossRef Full Text | Google Scholar

266. Atkinson, J, Chartier, Y, Pessoa-Silva, CL, Jensen, P, Li, Y, and Seto, W-H. Natural ventilation for infection control in health-care settings. 2nd ed. Geneva: World Health Organization (2009).

Google Scholar

267. Dai, R, Liu, S, Li, Q, Wu, H, Wu, L, and Ji, C. A systematic review and meta-analysis of indoor bioaerosols in hospitals: the influence of heating, ventilation, and air conditioning. PLoS One. (2021) 16:e0259996. doi: 10.1371/journal.pone.0259996

PubMed Abstract | CrossRef Full Text | Google Scholar

268. Gołofit-Szymczak, M, and Górny, RL. Microbiological air quality in office buildings equipped with dventilation systems. Indoor Air. (2018) 28:792–805. doi: 10.1111/ina.12495

PubMed Abstract | CrossRef Full Text | Google Scholar

269. Lin, W-R, Ho, Y-H, Lee, WK, Cheng, H-M, and Wang, P-H. Spatiotemporal distribution and the passive dispersal of fungal spores through HVAC systems. Aerobiologia (Bologna). (2022) 38:13–21. doi: 10.1007/s10453-021-09730-7

CrossRef Full Text | Google Scholar

270. Mousavi, ES, Kananizadeh, N, Martinello, RA, and Sherman, JD. COVID-19 outbreak and hospital air quality: a systematic review of evidence on air filtration and recirculation. Environ Sci Technol. (2021) 55:4134–47. doi: 10.1021/acs.est.0c03247

CrossRef Full Text | Google Scholar

271. Fonseca, A, Abreu, I, Guerreiro, M, Abreu, C, Silva, R, and Barros, N. Indoor air quality and sustainability management–case study in three Portuguese healthcare units. Sustainability. (2018) 11:101. doi: 10.3390/su11010101

CrossRef Full Text | Google Scholar

272. Chenari, B, Dias Carrilho, J, and Gameiro da Silva, M. Towards sustainable, energy-efficient and healthy ventilation strategies in buildings: a review. Renew Sust Energ Rev. (2016) 59:1426–47. doi: 10.1016/j.rser.2016.01.074

CrossRef Full Text | Google Scholar

273. Vornanen-Winqvist, C, Salonen, H, Järvi, K, Andersson, M, Mikkola, R, Marik, T, et al. Effects of ventilation improvement on measured and perceived indoor air quality in a school building with a hybrid ventilation system. Int J Environ Res Public Health. (2018) 15:1414. doi: 10.3390/ijerph15071414

PubMed Abstract | CrossRef Full Text | Google Scholar

274. Azimi, P, and Stephens, B. HVAC filtration for controlling infectious airborne disease transmission in indoor environments: predicting risk reductions and operational costs. Build Environ. (2013) 70:150–60. doi: 10.1016/j.buildenv.2013.08.025

PubMed Abstract | CrossRef Full Text | Google Scholar

275. Song, L, Zhou, J, Wang, C, Meng, G, Li, Y, Jarin, M, et al. Airborne pathogenic microorganisms and air cleaning technology development: a review. J Hazard Mater. (2022) 424:127429. doi: 10.1016/j.jhazmat.2021.127429

PubMed Abstract | CrossRef Full Text | Google Scholar

276. Polidori, A, Fine, PM, White, V, and Kwon, PS. Pilot study of high-performance air filtration for classroom applications. Indoor Air. (2013) 23:185–95. doi: 10.1111/ina.12013

PubMed Abstract | CrossRef Full Text | Google Scholar

277. Medical Advisory Secretariat. Air cleaning technologies: an evidence-based analysis. Ont Health Technol Assess Ser. (2005) 5:1–52.

PubMed Abstract | Google Scholar

278. ANSI. ANSI/ASHRAE standard 52.2-2017. Method of testing general ventilation air-cleaning devices for removal efficiency by particle size. (2017). Available at: https://www.ashrae.org/technology.

Google Scholar

279. Bastani, A, Lee, C-S, Haghighat, F, Flaherty, C, and Lakdawala, N. Assessing the performance of air cleaning devices–a full-scale test method. Build Environ. (2010) 45:143–9. doi: 10.1016/j.buildenv.2009.05.008

CrossRef Full Text | Google Scholar

280. Navaratnam, S, Nguyen, K, Selvaranjan, K, Zhang, G, Mendis, P, and Aye, L. Designing post COVID-19 buildings: approaches for achieving healthy buildings. Buildings. (2022) 12:74. doi: 10.3390/buildings12010074

CrossRef Full Text | Google Scholar

281. Xu, Z, Wu, Y, Shen, F, Chen, Q, Tan, M, and Yao, M. Bioaerosol science, technology, and engineering: past, present, and future. Aerosol Sci Technol. (2011) 45:1337–49. doi: 10.1080/02786826.2011.593591

CrossRef Full Text | Google Scholar

282. de Almeida, DS, Martins, LD, and Aguiar, ML. Air pollution control for indoor environments using nanofiber filters: a brief review and post-pandemic perspectives. Chem Eng J Adv. (2022) 11:100330. doi: 10.1016/j.ceja.2022.100330

CrossRef Full Text | Google Scholar

283. Shamim, JA, Hsu, W-L, and Daiguji, H. Review of component designs for post-COVID-19 HVAC systems: possibilities and challenges. Heliyon. (2022) 8:e09001. doi: 10.1016/j.heliyon.2022.e09001

CrossRef Full Text | Google Scholar

284. Luo, H, and Zhong, L. Ultraviolet germicidal irradiation (UVGI) for in-duct airborne bioaerosol disinfection: review and analysis of design factors. Build Environ. (2021) 197:107852. doi: 10.1016/j.buildenv.2021.107852

PubMed Abstract | CrossRef Full Text | Google Scholar

285. de Souza, SO, Cardoso, AA Jr, Sarmento, ASC, and d’Errico, F. Effectiveness of a UVC air disinfection system for the HVAC of an ICU. Eur Phys J Plus. (2022) 137:37. doi: 10.1140/epjp/s13360-021-02240-y

CrossRef Full Text | Google Scholar

286. Hsu, C-S, and Huang, D-J. Disinfection efficiency of chlorine dioxide gas in student cafeterias in Taiwan. J Air Waste Manage Assoc. (2013) 63:796–805. doi: 10.1080/10962247.2012.735212

PubMed Abstract | CrossRef Full Text | Google Scholar

287. Joubert, A, Abd Ali, SAZ, Frossard, M, and Andrès, Y. Dust and microbial filtration performance of regular and antimicrobial HVAC filters in realistic conditions. Environ Sci Pollut Res. (2021) 28:39907–19. doi: 10.1007/s11356-021-13330-w

PubMed Abstract | CrossRef Full Text | Google Scholar

288. Elsaid, AM, and Ahmed, MS. Indoor air quality strategies for air-conditioning and ventilation systems with the spread of the global coronavirus (COVID-19) epidemic: improvements and recommendations. Environ Res. (2021) 199:111314. doi: 10.1016/j.envres.2021.111314

PubMed Abstract | CrossRef Full Text | Google Scholar

289. Sánchez, B, Sánchez-Muñoz, M, Muñoz-Vicente, M, Cobas, G, Portela, R, Suárez, S, et al. Photocatalytic elimination of indoor air biological and chemical pollution in realistic conditions. Chemosphere. (2012) 87:625–30. doi: 10.1016/j.chemosphere.2012.01.050

PubMed Abstract | CrossRef Full Text | Google Scholar

290. Adhi, PM, Rahman, H, Nuriskasari, I, and Syamsudin, SA. A simple and low-cost photocatalytic air purification test method. J Phys Conf Ser. (2022) 2243:012075. doi: 10.1088/1742-6596/2243/1/012075

CrossRef Full Text | Google Scholar

291. Kang, Y, and Kato, S. A study on the effectiveness of microwave heating for disinfection of humidifier elements. HVAC&R Res. (2014) 20:113–20. doi: 10.1080/10789669.2013.834780

CrossRef Full Text | Google Scholar

292. Khurana, A. Ozone treatment for prevention of microbial growth in air conditioning systems. [Master’s Thesis]. Florida: University of Florida (2003).

Google Scholar

293. Chan, DWT, Law, KC, Kwan, CHS, and Chiu, WY. Application of an air purification system to control air-borne bacterial contamination in a university clinic. HKIE Trans. (2005) 12:17–21. doi: 10.1080/1023697X.2005.10667988

CrossRef Full Text | Google Scholar

294. Bolashikov, ZD, and Melikov, AK. Methods for air cleaning and protection of building occupants from airborne pathogens. Build Environ. (2009) 44:1378–85. doi: 10.1016/j.buildenv.2008.09.001

PubMed Abstract | CrossRef Full Text | Google Scholar

295. USDHHS, CDC. Guidelines for environmental infection control in health-care facilities recommendations of CDC and the healthcare infection control practices advisory committee (HICPAC), Atlanta. (2019). Available at: https://www.cdc.gov/infectioncontrol/guidelines/environmental/index.htm (Accessed August 25, 2022).

Google Scholar

296. De Robles, D, and Kramer, SW. Improving indoor air quality through the use of ultraviolet Technology in Commercial Buildings. Proc Eng. (2017) 196:888–94. doi: 10.1016/j.proeng.2017.08.021

CrossRef Full Text | Google Scholar

297. Zhong, L, and Haghighat, F. Ozonation air purification technology in HVAC applications. ASHRAE Trans (2014). AVailable at: https://link.gale.com/apps/doc/A371282980/AONE?u=anon~197f2a56&sid=googleScholar&xid=0f07a9a8 (Accessed October 16, 2023).

Google Scholar

298. Haider, AJ, Sukkar, KA, Abdalsalam, AH, Ali, AF, Jaber, SH, and Ridha, TTA. Enhancement of the air quality and heat transfer rate of an air-conditioning system using a hybrid polypropylene nanofilter. Process Saf Environ Prot. (2021) 149:56–66. doi: 10.1016/J.PSEP.2020.10.041

CrossRef Full Text | Google Scholar

299. Al-Rawi, M, Ikutegbe, CA, Auckaili, A, and Farid, MM. Sustainable technologies to improve indoor air quality in a residential house–a case study in Waikato, New Zealand. Energ Buildings. (2021) 250:111283. doi: 10.1016/j.enbuild.2021.111283

CrossRef Full Text | Google Scholar

300. Kim, W, Ko, S, Oh, M, Choi, I, and Shin, J. Is an incentive policy for energy efficient products effective for air purifiers? Case South Korea Energies (Basel). (2019) 12:1664. doi: 10.3390/en12091664

CrossRef Full Text | Google Scholar

301. Brągoszewska, E, and Biedroń, I. Efficiency of air purifiers at removing air pollutants in educational facilities: a preliminary study. Front. Environ Sci. (2021) 9:718. doi: 10.3389/fenvs.2021.709718

CrossRef Full Text | Google Scholar

302. Cooper, E, Wang, Y, Stamp, S, Nijsen, T, de Graaf, P, Hofman, J, et al. Why do people use portable air purifiers? Evidence from occupant surveys and air quality monitoring in homes in three European cities. Build Res Inf. (2022) 50:213–29. doi: 10.1080/09613218.2021.2001303

CrossRef Full Text | Google Scholar

303. Vijayan, V, Paramesh, H, Salvi, S, and Dalal, AK. Enhancing indoor air quality–the air filter advantage. Lung India. (2015) 32:473. doi: 10.4103/0970-2113.164174

PubMed Abstract | CrossRef Full Text | Google Scholar

304. Kujundzic, E, Matalkah, F, Howard, CJ, Hernandez, M, and Miller, SL. UV air cleaners and upper-room air ultraviolet germicidal irradiation for controlling airborne Bacteria and fungal spores. J Occup Environ Hyg. (2006) 3:536–46. doi: 10.1080/15459620600909799

PubMed Abstract | CrossRef Full Text | Google Scholar

305. Nguyen, TT, Johnson, GR, Bell, SC, and Knibbs, LD. A systematic literature review of indoor air disinfection techniques for airborne bacterial respiratory pathogens. Int J Environ Res Public Health. (2022) 19:1197. doi: 10.3390/ijerph19031197

PubMed Abstract | CrossRef Full Text | Google Scholar

306. Grinshpun, SA, Mainelis, G, Trunov, M, Adhikari, A, Reponen, T, and Willeke, K. Evaluation of ionic air purifiers for reducing aerosol exposure in confined indoor spaces. Indoor Air. (2005) 15:235–45. doi: 10.1111/j.1600-0668.2005.00364.x

PubMed Abstract | CrossRef Full Text | Google Scholar

307. Jakober, C, and Phillips, T. Evaluation of ozone emissions from portable indoor air cleaners: Electrostatic precipitators and ionizers. Sacramento, CA. (2008). Available at: http://www.arb.ca.gov/research/indoor/ozone.htm.

Google Scholar

308. Riaz, N, Khan, MS, Bilal, M, Ullah, S, and Al-Sehemi, AG. Photocatalytic inactivation of bioaerosols: a short review on emerging technologies. Curr Anal Chem. (2020) 17:31–7. doi: 10.2174/1573411016999200729115254

CrossRef Full Text | Google Scholar

309. Ruwan Jayakantha, DNP, Bandara, HMN, Gunawardana, NM, Jayantha Rajapakse, RPV, Thilakarathne, DS, Comini, E, et al. Design and construction of a low cost air purifier for killing harmful airborne microorganisms using a combination of a strong multi-directional electric-field and an ultra violet light. HardwareX. (2022) 11:e00279. doi: 10.1016/j.ohx.2022.e00279

CrossRef Full Text | Google Scholar

310. Luengas, A, Barona, A, Hort, C, Gallastegui, G, Platel, V, and Elias, A. A review of indoor air treatment technologies. Rev Environ Sci Biotechnol. (2015) 14:499–522. doi: 10.1007/s11157-015-9363-9

CrossRef Full Text | Google Scholar

311. González-Martín, J, Kraakman, NJR, Pérez, C, Lebrero, R, and Muñoz, R. A state–of–the-art review on indoor air pollution and strategies for indoor air pollution control. Chemosphere. (2021) 262:128376. doi: 10.1016/j.chemosphere.2020.128376

CrossRef Full Text | Google Scholar

312. Kaushik, AK, and Dhau, JS. Photoelectrochemical oxidation assisted air purifiers; perspective as potential tools to control indoor SARS-CoV-2 exposure. Appl Surf Sci Adv. (2022) 9:100236. doi: 10.1016/j.apsadv.2022.100236

CrossRef Full Text | Google Scholar

313. USEPA. Mold cleanup in your home. (2023). Available at: https://www.epa.gov/mold/mold-cleanup-your-home (Accessed October 16, 2023).

Google Scholar

314. Franke, DL, Cole, EC, Leese, KE, Foarde, KK, and Berry, MA. Cleaning for improved indoor air quality: an initial assessment of effectiveness. Indoor Air. (1997) 7:41–54. doi: 10.1111/j.1600-0668.1997.t01-3-00006.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: indoor environments, indoor air quality, public health, bioaerosols, microbial diversity

Citation: Chawla H, Anand P, Garg K, Bhagat N, Varmani SG, Bansal T, McBain AJ and Marwah RG (2023) A comprehensive review of microbial contamination in the indoor environment: sources, sampling, health risks, and mitigation strategies. Front. Public Health. 11:1285393. doi: 10.3389/fpubh.2023.1285393

Received: 29 August 2023; Accepted: 25 October 2023;
Published: 23 November 2023.

Edited by:

Jianbang Xiang, Sun Yat-sen University, China

Reviewed by:

Gaetano Settimo, Istituto Superiore di Sanità, Italy
Brian Crook, Health and Safety Laboratory, United Kingdom

Copyright © 2023 Chawla, Anand, Garg, Bhagat, Varmani, Bansal, McBain and Marwah. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Ruchi Gulati Marwah, ruchi.gulati@bcas.du.ac.in

ORCID: Andrew J. McBain, https://orcid.org/0000-0002-5255-5425

Ruchi Gulati Marwah, https://orcid.org/0000-0001-8945-7055

These authors have contributed equally to this work and share first authorship

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.