Skip to main content

Introduction

The presence of biological agents in occupational environments is an important health and social problem. Micro- and macroorganisms and the structures and substances they produce may exert a harmful influence upon exposed individuals leading to numerous adverse health outcomes. Biological agents that are transmitted as bioaerosols are of the greatest epidemiological importance. On a global scale, at least several hundred million people are estimated to be exposed to these risks at work[1]. From a public health perspective, the costs of hazardous exposure to bioaerosols are significant reaching billions of dollars[2][3].

Definitions

Bioaerosols are defined as particles of microbial, plant, or animal origin and oftentimes are called “organic dust." They can include live or dead bacteria, fungi, viruses, allergens, bacterial endotoxins (components of cell membranes of Gram-negative bacteria), antigens (molecules that can induce an immune response), toxins (toxins produced by microorganisms), mycotoxins (toxins produced by fungi), glucans (components of cell walls of many molds), pollen, and plant fibers [4].

Actinomycetesvaried group of rod-shaped to filamentous Gram-positive bacteria
Aerodynamic diameterdiameter of a unit-density sphere having the same gravitational settling velocity as the particle being measured
Bioaerosolairborne particles of biological origin
Biological agentsmicroorganisms, including those which have been genetically modified, cell cultures and human endoparasites, which may be able to provoke any infection, allergy or toxicity[5]
cfu, colony forming unitsnumber of microbial cells, spores or their aggregates which grow on agar media as separate colonies
Endotoxinconstituent of the external membrane of Gram-negative bacteria (lipopolysaccharide)
Glucanswater-insoluble structural cell-wall components of most of the fungi and yeasts, some bacteria and plants
Gram-positive/Gram-negativeterms describing specific staining characteristics of bacteria. Compared to Gram-positive bacteria (which are stained purple), Gram-negative bacteria typically have an additional outer membrane in their cell walls (and are stained pink)
Half-lifetime interval required to reduce the rate of emission of particles in question (e.g. microbial particles) by a factor of two
Interceptioncollision with and deposition of a particle on an object when the particle passes within the distance of one particle radius of the object
Mycotoxinstoxic secondary metabolites produced by fungi
Propagulereproductive particle released by an organism, e.g. a fungal spore

 

Biological and physical features of bioaerosol particles

Sources and occurrence

Bioaerosols are ubiquitous in both indoor and outdoor environments. Their major outdoor sources are soil, natural and anthropogenic water reservoirs as well as living and dead plants. Indoor sources can be occupational or non-occupational. Occupational sources can be very productive and can create bioaerosol concentrations up to 1012 cfu/m3. Among non-occupational ones, the most important source is usually the presence of humans and their physical activities (e.g. breathing, talking, sneezing, coughing, scratching etc.). Bioaerosol particles are usually naturally present in the environment. The frequency of their appearance and occurrence depends on: climatic zones, continents, geographic regions, character of the biotope as well as the occurrence and distribution of specific groups of plants or animals.

All these factors contribute to spatial variations of bioaerosol concentrations. Moreover, the environmental and temporal variations have been found to produce both diurnal and annual cyclic changes in the quantity and quality of bioaerosol particles[6]. Time-dependent variations can range over several orders of magnitude as bioaerosol sources do not generate particles continuously. Microclimate parameters, i.e. temperature and relative humidity of the air, together with other physical parameters of the environment (e.g. ultraviolet radiation, oxygen forms and availability) contribute to the influx of airborne microorganisms. These factors also influence the survival of airborne microbes and affect their ability to colonise on surfaces after deposition[7].

Biological properties

Bioaerosol particles may be viable (i.e. able to reproduce or have metabolic activity) or non-viable (i.e. dead or unable to reproduce as whole cells or as their fragments). From human exposure point of view, both states are important. To be infectious, a biological agent must be viable, whereas allergenic and toxic properties can be preserved even after death of the agent. In the environment, viable bioaerosol particles, especially microorganisms, are not stable and undergo mutations (genotype changes), phenotype changes, evolution and relative selection (loss of current traits or gain of new traits by the succeeding generation). All these changes (when certain new attributes are acquired and/or other irretrievably lost) can lead to appearance of bioaerosol particles with “new" features, e.g. bacteria resistant to certain antibiotics or disinfectants, fungi resistant to fungicides, protozoans resistant to drugs.

Physical properties

Biological agents frequently pose a threat to individuals in form of bioaerosols. They mostly occur in the environment as particles. The shape and other characteristics of the particles determine their behaviour and dynamics in the air and their deposition on surfaces. Bioaerosols can penetrate into the human body through the nose, mouth and conjunctiva epithelium, bronchi and alveoli in the lung, as well as epidermis. In the human respiratory tract, the penetration depth and behaviour of bioaerosol particles depends on their sizes, shapes, densities, electrical charges, chemical composition and reactivity. Besides, the physiological factors such as airflow and breathing patterns influence the mechanism of particle deposition as well[8]. A basic particle property is its size characterised by the particle diameter. Bioaerosol particles cover a wide size range from about 0.02 μm up to hundreds of micrometres. Only a small fraction of particles are spherical. Therefore particles of irregular shapes are characterised by means of a particle equivalent diameter.

To describe bioaerosol particles in exposure or health studies, the aerodynamic diameter is usually used[9][8]. The aerodynamic particle sizes may significantly differ from their physical sizes. For example, the physical size of Pseudomonas fluorescens bacterial rods is between 0.7-0.8×1.5-3 μm and the corresponding aerodynamic diameter is 0.8 μm; for spherical spores of the fungus Aspergillus versicolor, the aerodynamic diameter of 0.8 μm is equivalent to a physical dimension of 2-3.5 μm.

Behaviour in the respiratory tract

Due to a high settling velocity, the particles with diameters bigger than 11 μm practically do not penetrate into the deeper respiratory tract[10]. A vast majority of bacterial and fungal airborne particles are below this size being present as separate cells, spores, their fragments or as aggregates of the solely biological or biological and dust particles[11]. If a particle is highly hygroscopic, its aerodynamic diameter may be significantly larger due to humidity of the air and moisture conditions e.g. in the lung[12].

  • Most particles greater than 10 μm, and up to 80% of particles between 5-10 μm, are trapped in the nasopharyngeal region due to inertial impaction and centrifugal condensation resulting from the anatomic formation of these parts of the respiratory tract where the air stream has the highest velocity. For particles with an elevated ratio between their length and diameter e.g. for fungal spore chains, these processes are assisted by the interception mechanism. Particles trapped in upper parts of the respiratory tract are usually removed within a few hours by the mucociliary system or as a result of expectoration.
  • Particles with diameters above 0.5 μm are deposited by sedimentation and impaction, which take place in bronchi, bronchioles, and alveoli, where the air velocity is low and the probability of deposition is directly proportional to the residence time. Particles deposited in the lower respiratory tract (bronchioles, alveolar ducts and alveoli, without ciliated epithelium) are removed much slower from several dozen to several hundred days[13].
  • Particles of less than 0.5 μm are mainly separated from the air stream and deposited almost solely by diffusion. This process depends inversely proportional on the particle diameter and is supported by electrostatic precipitation resulting from interaction between surface and particle charges[14][15].

The type of interactions between the bioaerosol particles and human cells depends on the place of their deposition and is conditioned by their retention time in the respiratory tract:

  • Inhaled particles with aerodynamic diameters equal or above 10 μm (e.g. pollens, fungal spores) cause eye or nose irritations.
  • Particles, whose sizes are between 5 and 10 μm (e.g. fungal spores, bioaerosol aggregates) may provoke asthmatic symptoms. Those deposited in bronchi require up to 24 hours to be expelled.
  • Particles below 5 μm (e.g. single fungal spores, bacterial cells, their fragments and aggregates with dust particles) may evoke allergic alveolitis type reactions[16][17].
  • Field studies show that microbial particles with diameters below 2.5 μm (i.e. actinomycetal spores, most indoor fungal spores) are the most dangerous for human health if they are inhaled. Having the ability to avoid numerous defence systems in the respiratory tract (e.g. ciliated epithelium, mucus, saliva, etc.), they can burden the body with high concentrations of biologically active molecules[10]. All of them are subsequently moved to the oral cavity and then directly or indirectly (i.e. through the alimentary tract) eliminated from the organism.

Stability and viability in the air

The time interval in which particles are airborne must be long enough to give them opportunity to be inhaled. In real situations when the air undergoes permanent turbulences, this factor can be characterised by the so-called ’half-life’. Briefly: despite the fact that bioaerosol particles are rarely spherical, their sedimentation velocity can be relatively precisely calculated. According to the Stokes’ law, it is dependent (among others) on the square of particle diameter[9]. Based on that, for large bioaerosol particles of about 100 μm, estimated half-life equals to a few seconds, whereas for particles with aerodynamic diameters of about 10 μm, 3-1 μm and 0.5 μm half-lives increase roughly to a few minutes, hours and days, respectively[18].

The air as a biotope does not support the survival of biological agents. Nevertheless, several studies show that numerous fine particles of biological origin are able to maintain their viability (simultaneously preserving their immunological reactivity) in the air much longer than bigger organisms. For example: survival time for the bacterium Legionella pneumophila is less than 15 minutes, for Escherichia coli and Streptococcus faecalis it is between 30-60 minutes, for other streptococci up to 48 hours, and for staphylococci about 3 days. Whereas for the influenza virus it may be extended up to 3 weeks and for spores of the fungi Aspergillus and Penicillium up to 12 years, mite allergens are decomposed within a few months[19][20][21]. From the inhalation exposure point of view, an interrelation between stability and viability of bioaerosol particles determines the possible health outcomes.

Bioaerosol sources at workplaces

In many occupational settings workers can be exposed to a wide spectrum of bioaerosols. The extent to which their health is at risk depends on the source strength of the biological contamination and the nature of the work activities, depending on the abilities to generate and subsequently disperse biological particles into the air as well as the appropriateness of the control measures put in place to protect workers from the bioaerosols[22].

Even if the specific activity does not involve a deliberate intention to work with or use a biological agent, specific circumstances can lead to unintended emission or multiplication of biological agents and may result in workers' exposure above a certain level recognised as ‘normal’ for such environment. Therefore incidental generation of biological agents has to be taken into account in the risk assessment.

Several examples of occupational activities where the exposure to bioaerosols is the most frequent and pronounced in the occupational environments are presented in Table 1. Dutkiewicz et al. (2007) identified exposure to biological agents for 151 occupational groups in 22 main branches of industry.[23][24][25][26][27][28]

Table 1 - Selected working environments, their biological agents and adverse health outcomes they may cause as bioaerosols
Table 1 - Selected working environments, their biological agents and adverse health outcomes they may cause as bioaerosols

Table 1 - Selected working environments, their biological agents and adverse health outcomes they may cause as bioaerosols

Health effects of bioaerosols

Bioaerosols of working environments may be:

  • agents that cause infectious and invasive diseases (e.g. viruses, bacteria, fungi);
  • allergens (of bacterial, fungal, plant and animal origin);
  • toxins and biological compounds with similar toxic effects (e.g. bacterial endotoxins, glucans);
  • carcinogens (e.g. fungal mycotoxins such as aflatoxins, ochratoxins or trichothecenes);
  • immunologically reactive fragments (submicro- and nano-metric particles of bacterial or fungal origin).

Until quite recent occupational biohazards were commonly assimilated with infectious agents. This is still reflected in the current EU legislation. In the majority of recently published scientific papers, however, an integrative concept which takes into account an airborne transport of harmful biological agents as the most common route for their dispersion has been accepted[29].

Bioaerosols can be responsible for many adverse health effects from allergic reactions, through infections, to toxic reactions and other non-specific symptoms referred to e.g. as ‘sick building syndrome’ (SBS) or ‘mucous membrane syndrome’ (MMS, commonly results in a dry cough, irritation of the eyes, nose and throat and may represent an irritant effect not involving immune responses or mediators). In most situations the observed outcomes are a result of mixed exposure to toxins and allergens[23].

Exposure assessment and risk management

The report ‘Expert forecast on Emerging Biological Risks related to Occupational Safety and Health’[25] concluded that difficulties to conduct an assessment of biological risks, the lack of proper information, inadequate training or poor knowledge, together with poor maintenance of the working environment and inadequate emergency response plans to biological risks result in both poor risk management and poor prevention practices. In particular, the fact that companies and local authorities are not fully aware of and underestimate the risk of exposure to biological agents hampers an effective approach to the problem.

Figure 1: Algorithm of action to assess exposure to bioaerosols
Figure 1: Algorithm of action to assess exposure to bioaerosols

A simple algorithm of action in exposure assessment (Figure 1) should always be applied to eliminate the above mentioned shortcomings[30].

When investigating bioaerosols as a possible source of workplace exposure and health risk, an integrated assessment should also include examining heating, ventilation air conditioning systems, checking for water infiltration and moisture, evaluating microbial contamination in metal working fluids, and evaluating possible internal and external sources of bioaerosols. Measuring bioaerosols includes the measurement of viable (culturable and non-culturable) and nonviable bioaerosols in indoor settings (e.g., industrial, office, education, and residential buildings), industrial facilities (e.g., biotechnology, waste treatment, manufacturing, textile, and food processing), and outdoor environments (e.g., farms) [31].

A clear description of the dose-response relationship for the majority of biological agents cannot be established. One of the reasons seems to be the inadequacy of analytical methods applied in bioaerosol exposure assessment. Spatial and temporal variations usually do not permit measurement of the maximum possible concentration of the microbial agent.

  • Traditional methods of bioaerosol sampling and analysis focus on evaluation of viable microbial spores and vegetative cells, omitting the role of non-viable propagules and small fragments of their structural elements. To be infectious, bioaerosol particles must be viable; however, non-viable aerosol particles of biological origin can still be immunologically reactive and cause numerous adverse outcomes like allergies or toxic responses. This fact should be taken into account when exposure assessment is performed. The most commonly applied measurement procedures, based on bioaerosol viability at the level of 1-25%, significantly underestimate the real exposure[32][33], as they are a measure of only a small part of bioaerosol particles.
  • The duration of sampling is usually short and, due to environmental, spatial and temporal variations, does not adequately represent the real degree of environmental contamination.
  • Moreover, results can be biased by the time of measurement, i.e. microbial propagule release into the air may be sporadic, irregular, dependent on physical factors, sensitive to the specific environmental conditions (e.g. bioaerosol may not be well mixed in the air) and may not correspond well with the sampling time.

All of these may result in collection of unrepresentative samples. Hence, the exposure evaluation should contain, as its immanent part, the source identification and, if it is possible, reliable measure of microorganisms.

Source identification is often carried out by surface sampling using different techniques (e.g. transparent sticky tape, swab sampling or contact plates). All these methods can identify the source but cannot evaluate the magnitude of emission of microbial propagules into the surrounding air. Hence, the surface sampling is crude, expected to yield a poor surrogate of airborne concentrations and its results should be interpreted with caution[26].

Several studies have shown that personal exposure to particles is usually higher than the concentration measured simultaneously with stationary samplers. In epidemiological studies which relied on personal sampling of microorganisms, exposure-response associations were found almost twice as often as in studies using stationary sampling. As stationary sampling appears to underestimate the bioaerosol exposure, personal sampling is thus the method of choice[34].

At present, there are no worldwide distinct regulations to standardise bioaerosol sampling in occupational environments. Hence, the guidelines contained in European standards (Table 2) published by CEN TC 137 describing the methods of workplace control can be applied for this purpose[1]. They simultaneously provide a base for development of more detailed guidelines and/or instructions for bioaerosol assessment and control in individual countries as it was done for example by the Institute for Occupational Safety and Health of the German Social Accident Insurance (IFA)[35].

Table 2 - European standards related to bioaerosol measurement at workplaces

Standard referenceTitleScope
EN 13098:2019Workplace exposure - Measurement of airborne microorganisms and microbial compounds - General requirementsThe standard - specifies general requirements for the measurement of microorganisms and microbial compounds - provides guidelines for the assessment of workplace exposure to airborne microorganisms including the determination of total number and culturable number of microorganisms and microbial compounds in the workplace atmosphere. - does not apply to the measurement of viruses.
EN 14031:2003Workplace atmospheres - Determination of airborne endotoxinsThe standard provides - guidelines for the assessment of workplace exposure to airborne bacterial endotoxins. - methods for sampling, transportation, and storage of samples and determination of endotoxins.
EN 14583:2004Workplace atmospheres - Volumetric bioaerosol sampling devices - Requirements and test methodsThis standard specifies requirements and test methods to determine the performance of volumetric sampling devices used to assess bioaerosols in the workplace.

 

Threshold limit values for occupational bioaerosols

On the global scale, there is a lack of acceptable occupational exposure limits (or threshold limit values) for bioaerosols. The main reason for this is the lack of well-documented and epidemiologically proven dose-response relationships between the exposure to specific biological agents and adverse health effects caused by their exact dose(s). Moreover, the sensitivity to each organism is individual and the strength of immunological reaction to the specific agent(s) is usually not the same in everybody. Despite progress in development of aerosol sampling techniques and analytical methods over the last two decades, the worldwide scientific database on bioaerosols is still insufficient to quantitatively and qualitatively characterise them. If threshold limit or reference values are established, they are usually connected with the clinical picture of the specific disease caused by the agent, taking into consideration its presence in a certain element of the environment only. Nevertheless, despite these limitations, the reference values expressed in numbers are applied to facilitate interpretation of the measurement data. 

Figure 2: Interdependence of key factors important for the elaboration of threshold limit values for bioaerosols
Figure 2: Interdependence of key factors important for the elaboration of threshold limit values for bioaerosols

An appropriate elaboration strategy of threshold limit values for bioaerosols takes into consideration the research method and several environmental, source, quantitative and qualitative criteria, which play a key role in such a process. Figure 2 illustrates their interdependencies[1].

From the point of view of clinical outcomes caused by exposure to bioaerosols, the ideal situation would be as follows:

  1. A hygienic standard is created based on the relationship between exposure, type and concentration of biological agents and their health effects.
  2. The relationship would be epidemiologically proven.
  3. Experimental evidence proving this relationship would become known.
  4. The relationship would be clinically proven, i.e. medical evidence that a specific agent was responsible for certain specific health effects exists.

Today, this ideal situation does not exist for any biological agent. Therefore, due to the lack of possibilities to determine the relationship between the dose of biological agent(s) and health outcome in a precise way, the creation of a threshold limit value based on the "environmental philosophy" seems to be a reasonable alternative for the “clinical" approach[1].

Reference values derived from multiple biological agent concentration measurements, should enable an evaluation of the quality of environment, as well as the determination of “what is typical or acceptable" and “what is atypical or inacceptable" for a specific type of environment (or its specific area).

Guidance values have been proposed as maximum levels for culturable fungi and bacteria, (colony-forming units (cfu)/m3) but they differ between countries depending on the outcomes of research studies. Examples of such guidance values that can be found in the scientific literature for bioaerosols in occupational environments are:

  • for total number of bacteria: ≤1.0×103-7.0×103 cfu/m3 for non-industrial workplaces and ≤7.5×102-1.0×107 cfu/m3 for manufacturing and industrial premises; for pathogenic microorganisms, there is no safety level (the threshold limit should be 0 cfu/m3);
  • for Gram-negative bacteria: 1.0×103-2.0×104 cfu/m3 for manufacturing and industrial premises;
  • for bacterial endotoxin: 0.005-0.2 µg/m3for productive and industrial premises;
  • for fungi: 1.0×101-1.0×104 cfu/m3 for non-industrial workplaces and <1.0×102-1.0×107 cfu/m3 for manufacturing and industrial premises; for the pathogenic microorganisms, there is no safety level (the threshold limit should be 0 cfu/m3).

In special environments like hospital rooms or clean rooms during operation, the guidance values for airborne microbial contaminants should be within the range of 1.0×100-4.0×103 cfu/m3 and <1.0×100-1.0×103 cfu/m3, respectively. In case of indoor spaces where a high air quality is required (e.g. clean rooms in action), not only the guidance values for biological agents in the air are applied but recommended limits for microbial contamination of surfaces are used as well. A wide review of guidance values for occupational bioaerosols can be found in the publications by Brandys & Brandys (2007)[36] or Górny et al. (2011)[1].

Legislative aspects

In Europe, Directive 2000/54/EC on the protection of workers against health and safety risks related to exposure to biological agents lays down the principles for the management of biological agents related risks and assigns to employers the duty of assessing the risks posed by biological agents in the occupational environment[5]. The directive defines employers’ obligations regarding, for example: methods of risk determination, assessment, elimination and reduction of risks, rules of hygiene and individual protection, training of workers and records of exposures, accidents and incidents. The classification of biological agents (known to infect humans) into four risk groups, containment measures and levels as well as containment for industrial processes are specified in the annexes. Directive 2000/54/EC gives minimum requirements aiming at the reduction of health risks from biological agents in the workplace. Within the European Union, the Directive provisions have been introduced to the legal systems of individual Member States by suitable legal instruments (e.g. in Poland, it is an Ordinance of the Ministry of Health) or Codes of Practice and guidelines. It is therefore important to refer to the relevant national regulations related to biological agents at workplaces.

Prevention and control measures

Directive 2000/54/EC obliges employers to protect workers against risks to their health and safety, including the prevention of such risks, arising or likely to arise from exposure to biological agents at work. This requires assessing the risks of exposure to biological agents based on information regarding the classification of biological agents, recommendations of competent authorities, information on diseases which may be contracted as a result of work, potential allergenic or toxigenic effects as a result of work, and knowledge of a work-related disease from which a worker is found to be suffering[5]. Based on the risk assessment as well as the nature and possible extent of associated adverse effects, the next step is to design measures to prevent or control these hazards[37].

The risk assessement and the prevention measures have to be based on reliable information on the properties and risks of the bioaerosols. National and sectoral guidance is available in a number of countries and it may provide information on the biological agents encountered in some jobs or professions. For example, the GESTIS Biological Agents Database contains data on almost 15,000 biological agents. For each agent information is provided on the basic technical, organisational and personal protective measures during activities in laboratories, biotechnology and keeping of experimental animals. It also contains task sheets which provide advice for taking adequate prevention and control measures.

The most effective prevention strategy is ensuring a healthy work environment without the risk of exposure to bioaerosols. While it is not usually feasible to completely eliminate the risk posed by biological agents, its reduction to the lowest possible level can be achieved by implementing prevention measures based on the S.T.O.P. principle. This prevention strategy is based on four types of measures, i.e. substitution, technical, organisational and personal, that should be taken in a hierarchical order and as part of one holistic approach[38][39][40].

Examples of prevention measures regarding occupational bioaerosol exposure for each type of measure (in hierarchical order) are as follows:

  1. Substitution: elimination or (if not possible) substitution of biological agents/processes with others less ones;
  2. Technical measures: minimisation of the release of bioaerosols (e.g. fast delivery, short storage times or immediate processing of critical materials), guarding/fencing/shielding of machinery and other equipment to reduce bioaerosol immissions, avoidance of manual processing, controlled atmosphere in workplaces with air filtration or air conditioning;
  3. Organisational control measures based on systems and procures by designing suitable systems of work and maintaining plant and equipment in safe and hygienic conditions (e.g. cleaning of workplaces must be considered an integral part of operations and it should be carried out properly in order to minimise dust generation, workplaces should be designed with easy-to-clean surfaces, separate storage of leisure and work clothing; regular cleaning and changing of work and protective clothes, provision of facilities to wash hands (or take a shower) when leaving the workplace, avoidance of eating, drinking or smoking at the workplace, provision of clean and separated storage facilities for food and drinks); Other organisational measures such as isolation of workstations (e.g. automatically closing doors, sluice), restriction of entrance to areas with high bioaerosol levels to an operational minimum number of workers, workers’ information and training to promote safe work practices, medical surveillance (preventive medical check-ups and vaccinations, monitoring of exposure and record-keeping); proper labelling, safe storage, procedures for safe transfer, handling, use and disposal of agents being used and of waste.
  4. Personal measures: respiratory protection including respirators with clean filtered air supply, personal protective equipment (clothes, gloves and goggles; however, such equipment should be used as the last possible prevention measure only when eliminating or reducing the level of risk to an acceptable level is not possible).

New trends in bioaerosol measurement and control

Microbial source strength concept

The first attempt which tries to eliminate all the weak points of the data gathered by traditional bioaerosol sampling is the microbial source strength concept[41]. Source strength in this context means the ability of microbial sources for maximal emission of particles into the air under the most favourable release conditions. This novel idea assumes a dynamic description of the aerosolisation process of microbial particles from their source (e.g. quantification of the particle emission rate from a microbiologically contaminated surface of building material). According to this concept, the aerosolisation potential is not only limited to the emission from the source of fungal or bacterial spores or vegetative cells, but also includes the role of fragments of microbial colony structures. It allows the assessment of the maximal potential exposure irrespective of the viability of aerosolised microbial particles or temporal and spatial variations of propagule emission[42].

Monoclonal antibodies in bioaerosol analyses

Among the new techniques, that are currently being developed to analyse bioaerosol samples are a species-specific detection method based on immunoassays using monoclonal antibodies (mAbs) or phage display reagents in combination with enzyme-linked immunosorbent assays (ELISA) or immunofluorescent techniques. A successful implementation of sensitive and specific enzyme immunoassays for fungal aerosols in air quality research might only be achieved if highly specific antibodies are used, and if sample processing and sample analysis are standardised. This will increase the accuracy and precision of monitoring data and further our understanding of the temporal and spatial dynamics of fungal contamination. Standardisation will also enhance direct comparisons of monitoring results obtained in different laboratories. It is especially important in epidemiological studies characterising causal relationships between adverse health effects and fungal exposure[43][44].

Health protection by electric charges and microwaves

Microorganisms in the airborne state carry electric charges. Their amount depends on the microbial species and the dispersion method, and can be more than 10,000 elementary electric charges[45]. Elaboration of new antistatic materials and subsequent production of special clothes (e.g. medical aprons) may substantially reduce the number of infections caused by airborne microbial agents. As it was shown by Allen & Henshaw (2006), new antistatic apron materials may constitute an improved barrier to the spread of microorganisms exhibiting a 38% reduction in bacteria attracted onto their surface compared with the white plastic aprons currently in use[46]. The ability to carry electrical charges on airborne microorganisms can also be used to eliminate them from the contaminated environment. To reduce human exposure to bioaerosol particles, the unipolar air ionisation may be successfully applied. As it was shown, a significant percentage (up to 92%) of airborne viable bacteria could be inactivated by the ion emission[47]. Moreover, such process, if performed within the human breathing zone, may considerably enhance the efficiency of filtering facepiece respirators or even surgical masks. The experiments showed that for a particle size range of ~0.04-1.3 μm, air ionisation in the vicinity of a manikin with a respiratory mask enables the particle penetration efficiency reduction up to about 3,000-fold[48].

Due to both high penetration efficiency and sterilisation effectiveness, the microwave technique is utilised for non-invasive cleaning and could help in effective protection against microbial contamination. The experiments revealed that microwave radiation can decrease both fungal and actinomycetal spore viability when growing on building materials. Such radiation can also deprive microbial spores of their cytotoxic properties. The effectiveness of this process, however, strongly depends on the degree of surface hydration and on the combination of power density and time of exposure to microwaves[49][50].

Conclusion

Bioaerosols are ubiquitous in nature, however some human activities such as animal farming, waste management, the food industry, healthcare and biotechnology, may substantially influence them both quantitatively and qualitatively. The presence of bioaerosols can lead to various human diseases covering respiratory symptoms, allergic reactions, non-infectious diseases as well as infectious diseases. However, understanding the impacts of bioaerosol exposure remains difficult as there is little data available on the exposure-response relationship. There is need for standardised protocols and methods on exposure assessment of bioaerosols to support effective risk management strategies in the workplace. Novel techniques using electric charges of microorganisms or microwaves may help reduce infection potential.

References

[1] Dutkiewicz, J., ‘Biological agents’, In: Koradecka, D. (Ed.), ‘Handbook of Occupational Safety and Health’, CRC Press, Boca Raton, 2010, pp. 385-400.

[2] Cox, C. S. & Wathes, C. M., ‘Bioaerosols handbook’, Lewis Publishers/CRC Press, Inc., Boca Raton, 1995.

[3] Crook, B., ‘Difficulty of assessing biological risks in the workplace’. Seminar ‘Occupational biological risks: Facing up to the challenges’ organised by EU-OSHA, Brussels, 5-6 June 2007. Available at: https://osha.europa.eu/en/tools-and-resources/seminars/occupational-risks-biological-agents-facing-challenges

[4] Rogoff, M., Planning of Solid Waste Recycling Facilities and Programs, Solid Waste Recycling and Processing, 2014, pp. 43-111

[5] Directive 2000/54/EC of the European Parliament and of the Council of 18 September 2000 on the protection of workers from risks related to exposure to biological agents at work (seventh individual directive within the meaning of Article 16 of Directive 89/391/EEC). Available at: [13]

[6] Lighthart, B. & Stetzenbach, L. D.‚ Distribution of microbial bioaerosol’. In: Lighthart, B. & Mohr, A. J. (Eds.), ‘Atmospheric microbial aerosols: theory and applications’. Chapman and Hall, Inc., New York, 1994, pp. 68-98.

[7] Stetzenbach, L., ‘Introduction to aerobiology’. In: Hurst C. J. (Ed.), ‘Manual of environmental microbiology’. ASM Press, Washington, 1997, pp. 619-628.

[8] WHO – World Health Organisation, ‘Guidelines for concentration and exposure-response measurement of fine and ultra-fine particulate matter for use in epidemiological studies’, 2002, pp. 39-65. Available at: http://whqlibdoc.who.int/hq/2002/a76621.pdf

[9] Baron, P. A. & Willeke, K. (Eds.), ‘Aerosol measurement: principles, techniques, and applications’. John Wiley & Sons, Inc., New York, 2001.

[10] Spengler, J. & Wilson, R. ’Emission, dispersion, and concentration of particles’. In: Wilson, R. & Spengler, J.D. (Eds.), ‘Particles in our air: concentrations and health effects’, Harvard University Press, Cambridge, 1996, pp. 41-62.

[11] Górny, R. L., Dutkiewicz, J. & Krysińska-Traczyk, E., 1999, Size distribution of bacterial and fungal bioaerosols in indoor air. Annals of Agricultural and Environmental Medicine, Vol. 6, No. 2, 1999, pp. 105-113.

[12] Reponen, T., Grinshpun, S. A., Conwell, K. L., Wiest, J. & Anderson, M., ‘Aerodynamic versus physical size of spores: measurement and implication on respiratory deposition, Grana, Vol. 40, 2001, pp. 119-125.

[13] Clarke, S., ‘Physical defenses’, In: Brewis, R. A. L., Gibson, G. J. & Geddes, D. M. (Eds.), ‘Respiratory medicine’, Bailliere Tindall, London, 1990.

[14] Lippmann, M., ‘Respiratory tract deposition and clearance of aerosols’. In: Lee, S. D., Schneider, T., Grant, L. D. & Verkerk, P. J. (Eds.), ‘Risk Assessment and Control Strategies’. Lewis Publishers, Chelsea, 1986, pp. 43-57.

[15] Utell, M. & Samet, J. ‘Airborne particles and respiratory disease: clinical and pathogenetic considerations’. In: Wilson, R. & Spengler, J.D. (Eds.), ‘Particles in our air: concentrations and health effects’, Harvard University Press, Cambridge, 1996, pp. 169-188.

[16] Owen, M. K., Ensor, D. S. & Sparks, L. E., ’Airborne particle sizes and sources found in indoor air’, Atmospheric Environment, Vol. 26, 1992, pp. 2149-2162.

[17] Horner, W. E., Helbling, A., Salvaggio, J. E. & Lehrer, S. B., ’Fungal allergens’, Clinical Microbiology Reviews, Vol. 8, 1995, pp. 161-179.

[18] Martinez, K. F., ’Anthrax: environmental sampling’, In: ‘Mold, spores, and remediation workshop’. ACGIH Worldwide, Cincinnati, 2002.

[19] Burge, H. A., (Ed.), ‘Bioaerosols’, Lewis Publishers/CRC Press, Inc., Boca Raton, 1995.

[20] Flannigan, B., ‘Biological particles in the air of indoor environments’, In: Johanning, E. & Yang, C. S. (Eds.): Fungi and bacteria in indoor air environment, Proceedings of the International Conference at Saratoga Springs, New York, 1994, pp. 21-29.

[21] Yang, C. S., ‘Understanding the biology of fungi found indoors’, In: Johanning, E. & Yang, C. S. (Eds.), Fungi and bacteria in indoor air environment, Proceedings of the International Conference at Saratoga Springs, New York, 1994, pp. 131-137.

[22] Crook, B. & Swan, J. R. M., ‘Bacteria and other bioaerosols in industrial workplaces’, In: Flannigan, B., Samson, R. A. & Miller, J. D. (Eds.), ‘Microorganisms in home and indoor work environments’, CRC Press, Boca Raton, 2001, pp. 69-82.

[23] Douwes, J., Thorne, P., Pearce, N. & Heederik, D., ‘Bioaerosol health effects and exposure assessment: progress and prospects’, Annals of Occupational Hygiene, Vol. 47, No. 3, 2003, pp. 187-200.

[24] Dutkiewicz, J., Śpiewak, R., Jabłoński, L. & Szymanska, J. ‘Occupational biohazards. Classification, exposed workers, measurements, prevention’, Ad Punctum, Lublin, 2007.

[25] EU-OSHA – European Agency for Safety and Health at Work, ‘Expert forecast on emerging biological risks related to occupational safety and health’, European risk observatory report, Office for Official Publications of the European Communities, Luxembourg, 2007, pp. 87-88. Available at: http://osha.europa.eu/en/publications/reports/7606488

[26] IOM – Institute of Medicine, ‘Damp indoor spaces and health’, National Academies Press, Washington, 2004.

[27] Macher, J. (Ed.), ‘Bioaerosols: assessment and control’, American Conference of Governmental Industrial Hygienists, Cincinnati, 1999.

[28] WHO – World Health Organisation, ‘Guidelines for indoor air quality: dampness and mould’, WHO Regional Office for Europe, Copenhagen, 2009. Available at: http://www.euro.who.int/__data/assets/pdf_file/0017/43325/E92645.pdf

[29] Sullivan, J. B. Jr & Krieger, G. R. (Eds.), ‘Clinical and environmental health and toxic exposures’, Lippincott Williams & Wilkins, Philadelphia, 2001.

[30] Górny, R. L., Cyprowski, M., Ławniczek-Wałczyk, A., Gołofit-Szymczak, M. & Zapór, L., ‘Biohazards in the indoor environment – a role for threshold limit values in exposure assessment’, In: Dudzińska, M. R. (Ed.), ‘Management of indoor air quality’, CRC Press/Balkema, Taylor & Francis Group, Leiden, 2011, pp. 1-20

[31] Lindsley, W., Green, B., Blachere, F., Martin, S., Law, B., Jensen, P., Schafer, M., Sampling and characterization of bioaerosols, Chapter 5, NIOSH Manual of Analytical Methods, 2017

[32] Flannigan, B., ‘Membrane filtration sampling for detection of small numbers of culturable microorganisms in indoor air’, In: Raw, G., Aizlewood, C. & Warren, P. (Eds.), Proceedings of Indoor Air, Vol. 4, 1999, pp. 948-949.

[33] Karlsson, K. & Malmberg, P., ‘Characterization of exposure to molds and actinomycetes in agricultural dusts by scanning electron microscopy, fluorescence microscopy and the culture methods’, Scand J Work Environ Health, Vol. 15, 1989, pp. 353-359.

[34] Thorne, P. S., Duchaine, C., Douwes, J., Eduard, W., Górny, R., Jacobs, R., Reponen, T., Schierl, R. & Szponar, B., ‘Working group report 4: exposure assessment for biological agents’, American Journal of Industrial Medicine, Vol. 46, 2004, pp. 419-422.

[35] Albrecht, A., Kiel, K. & Kolk, A. ‘Strategies and methods for investigation of airborne biological agents from work environments in Germany’, International Journal of Occupational Safety and Ergonomics (JOSE), Vol. 13, No. 2, 2007, pp. 201-213.

[36] Brandys, R. C., Brandys, G. M., ‘Worldwide exposure standards for mold and bacteria. 7th edition’, Hinsdale: OEHCS, Inc., Publications Division, 2007.

[37] Ferrari Goelzer, B. I., ‘Occupational hygiene. Goals, definitions and general information’, In: Stellman, J. M., ‘Encyclopaedia of occupational health and safety’, International Labour Organisation, Geneva, Vol. I, part IV, Chapter 30, 1998. Available at: https://www.iloencyclopaedia.org/part-iv-66769/occupational-hygiene-47504/item/570-goals-definitions-and-general-information

[38] Boyle, T., ‘Health and safety: risk management’, IOSH Services Ltd; Leicester, 2008.

[39] ECHA – European Chemicals Agency, ’Guidance on information requirements and chemical safety assessment. Chapter R.14: Occupational exposure estimation, 2016. Available at: https://echa.europa.eu/documents/10162/13632/information_requirements_r14_en.pdf/bb14b581-f7ef-4587-a171-17bf4b332378

[40] EU-OSHA - European Agency for Safety and Health at Work, Seminar ‘Occupational risks from biological agents: Facing up the challenges’, organised by EU-OSHA in Brussels on 5-6 June 2007. Online summary available at http://osha.europa.eu/en/seminars/occupational-risks-from-biological-agents-facing-up-the-challenges

[41] Grinshpun, S. A., Górny, R. L., Reponen, T., Willeke, K., Trakumas, S., Hall, P. & Dietrich, D.F., ‘New method for assessment of potential spore aerosolization from contaminated surfaces’, Proceedings of 6th International Aerosol Conference, Taipei, Taiwan, September 8-13, Vol. 2, 2002, pp. 767-768.

[42] Górny, R. L., ‘Filamentous microorganisms and their fragments in indoor air - a review’, Annals of Agricultural and Environmental Medicine, Vol. 11, No. 2, 2004, pp. 185-197.

[43] Schmechel, D., Górny, R. L., Simpson, J. P., Reponen, T., Grinshpun, S. A. & Lewis, D. M., ‘Limitations of monoclonal antibodies for monitoring of fungal aerosols using Penicillium brevicompactum as a model fungus’, Journal of Immunological Methods, Vol. 283, 2003, pp. 235-245

[44] Schmechel, D., Górny, R. L., Simpson, J. P., Reponen, T., Grinshpun, S. A., Beezhold, D. & Lewis, D. M., ‘The potentials and limitations of monoclonal antibody-based monitoring techniques for fungal bioaerosols’, Workshop on Methods of bioaerosol detection, Karlsruhe, Germany, 8-9 July, 2004.

[45] Mainelis, G., Willeke, K., Baron, P., Grinshpun, S. A., Reponen, T., Górny, R. L. & Trakumas, S., ‘Electrical charges on airborne microorganisms’, Journal of Aerosol Science, Vol. 32, 2001, pp. 1087-1110.

[46] Allen, J. E. & Henshaw, D. L., ‘Antistatic nurses’ aprons may contribute to a reduction of hospital infection in isolation wards’, 2006, pp. 1-9.

[47] Grinshpun, S. A., Adhikari, A., Lee, B. U., Trunov, M., Mainelis, G., Yermakov, M. & Reponen, T., ‘Indoor air pollution control through ionization’, In: Brebbia, C. A., ‘Air pollution XII’, WIT Press, Southampton, 2004, pp. 689-704.

[48] Lee, B. U., Yermakov, M. & Grinshpun, S. A., ‘Filtering efficiency of N95- and R95-type facepiece respirators, dust-mist facepiece respirators, and surgical masks operating in unipolarly ionized indoor air environments’, Aerosol and Air Quality, Vol. 5, No 1, 2005, pp. 25-38.

[49] Górny, R. L., Mainelis, G., Wlazło, A., Niesler, A., Lis, D. O., Marzec, S., Siwińska, E., Łudzeń-Izbińska, B., Harkawy, A. & Kasznia-Kocot, J., ‘Viability of fungal and actinomycetal spores after microwave radiation of building materials’, Annals of Agricultural and Environmental Medicine, Vol. 14, No. 2, 2007, pp. 313-324.

[50] Górny, R. L., Wlazło, A., Kapka, L., Niesler, A., Ławniczek-Wałczyk, A., Kozłowska, A., Lis, D. O. & Gołofit-Szymczak, M., Change in fungal and actinomycetal spore cytotoxicity after exposure to microwave radiation, 9th International Congress on Aerobiology, Buenos Aires, Argentina, 23-27 August 2010, pp. 58-59.

Further reading

EU-OSHA - European Agency for Safety and Health at Work 2003, ‘Factsheet 41 – Biological agents’. Available at http://osha.europa.eu/en/publications/factsheets/41/view

EU-OSHA – European Agency for Safety and Health at Work 2011, ‘Factsheet 100 - Legionella and legionnaires’ disease: European policies and good practices’. Available at http://osha.europa.eu/en/publications/factsheets/100/view

EU-OSHA - European Agency for Safety and Health at Work, E-fact 53: Risk assessment for biological agents, Available at: https://osha.europa.eu/en/publications/e-fact-53-risk-assessment-biological-agents/view

EU-OSHA - European Agency for Safety and Health at Work, Biological agents and work-related diseases: results of a literature review, expert survey and analysis of monitoring systems, Available at: https://osha.europa.eu/en/publications/biological-agents-and-work-related-diseases-results-literature-review-expert-survey-and/view

EU-OSHA - European Agency for Safety and Health at Work, Exposure to biological agents and related health problems in animal-related occupations, Available at: https://osha.europa.eu/en/publications/exposure-biological-agents-and-related-health-problems-animal-related-occupations/view

EU-OSHA - European Agency for Safety and Health at Work, Exposure to biological agents and related health problems in arable farming, Available at: https://osha.europa.eu/en/publications/exposure-biological-agents-and-related-health-problems-arable-farming/view

EU-OSHA - European Agency for Safety and Health at Work, Exposure to biological agents and related health problems for healthcare workers, Available at: https://osha.europa.eu/en/publications/exposure-biological-agents-and-related-health-problems-healthcare-workers/view

EU-OSHA - European Agency for Safety and Health at Work, Exposure to biological agents and related health effects in the waste management and wastewater treatment sectors, Available at: https://osha.europa.eu/en/publications/exposure-biological-agents-and-related-health-effects-waste-management-and-wastewater/view

EU-OSHA - European Agency for Safety and Health at Work, Biological agents and associated work-related diseases in occupations that involve travelling and contact with travellers, Available at: https://osha.europa.eu/en/publications/biological-agents-and-associated-work-related-diseases-occupations-involve-travelling/view

BAuA - German Institute for Occupational Safety and Health, Measurement of Airborne Bacteria.

Flannigan, B., Samson, R. A. & Miller, J. D. (Eds.), ‘Microorganisms in home and indoor work environments’, CRC Press, Boca Raton, 2011.

Hung, L.-L., Miller, J. D. & Dillon, K. (Eds.), ‘Field guide for the determination of biological contaminants in environmental samples’, AIHA, Fairfax, 2005.

Kolk A., ‘Managing biological hazards in the workplace ’, in Magazine 6, Dangerous Substances Handle with care, EU-OSHA – European Agency for Safety and Health at Work 2003, p. 31

Prezant, B., Weekes, D. M. & Miller, J. D., ‘Recognition, evaluation, and control of indoor mold’, AIHA, Fairfax, 2008.

Waites, M. J., Morgan, N. L., Rockey, J. S. & Higton, G. (Eds.), ‘Industrial microbiology: an introduction’, Blackwell Science Ltd., Oxford, 2001.

Select theme

Contributor

Karla Van den Broek

Prevent, Belgium
Klaus Kuhl

Rafal L. Gorny