5,637
Views
149
CrossRef citations to date
0
Altmetric
Special Focus Review

Particle size and pathogenicity in the respiratory tract

Pages 847-858 | Received 30 Sep 2013, Accepted 12 Nov 2013, Published online: 13 Nov 2013

Abstract

Particle size dictates where aerosolized pathogens deposit in the respiratory tract, thereafter the pathogens potential to cause disease is influenced by tissue tropism, clearance kinetics and the host immunological response. This interplay brings pathogens into contact with a range of tissues spanning the respiratory tract and associated anatomical structures. In animal models, differential deposition within the respiratory tract influences infection kinetics for numerous select agents. Greater numbers of pathogens are required to infect the upper (URT) compared with the lower respiratory tract (LRT), and in comparison the URT infections are protracted with reduced mortality. Pathogenesis in the URT is characterized by infection of the URT lymphoid tissues, cervical lymphadenopathy and septicemia, closely resembling reported human infections of the URT. The olfactory, gastrointestinal, and ophthalmic systems are also infected in a pathogen-dependent manner. The relevant literature is reviewed with respect to particle size and infection of the URT in animal models and humans.

Introduction: The Ubiquitous Nature of Bioaerosols

Bioaerosols are defined as a collection of particles suspended on a column of air derived from or incorporating material of biological origin. Individual particles within the bioaerosol can vary markedly in aerodynamic diameter from submicron (e.g., viruses) to many millimeters (e.g., pet dander). Importantly, bioaerosols can affect human health due to the presence of pathogens or allergens. The sources of bioaerosols containing pathogens are varied and individual aerosol particles may vary widely from submicron to several millimeters dependent on the method of aerosol generation and attachment to larger particulates, for example skin cells (); although only those <100 μm are within the inhalable range for humans, it should be noted that the initial particles will rapidly evaporate depending on the humidity of the local environment.Citation4,Citation33,Citation34 The final inhaled particle size is dependent on a number of factors including the solid organic content of the initial particle (including pathogens) and location of an individual to the aerosol source. It is evident that even pollen grains that represent comparatively larger bioaerosols can travel large distances in conducive meteorological conditions.Citation15,Citation35 Airborne dissemination whether outdoors or indoors is influenced by a number of inter-relating factors that impact on air mass movement, turbulence, and thermal convection including meteorology, vehicle/human activity, and ventilation that may far outweigh the terminal velocity of particles that are generally calculated in very still air.Citation4,Citation34 This review will concentrate on aerosols containing pathogens and the inter-relationship between factors governing particle size, deposition site, clearance, and inhalational infection.

Table 1. Sources of bioaerosol

Aerosol Transmission: Relation to Mechanism of Generation

The mechanism of generation influences the particle size of the resultant bioaerosol (), and these may be biotic (e.g., sneeze or pollen), or abiotic where the aerosol is produced by a non-living system (e.g., water cooling towers). Irrespective, all aerosols will be generated with an initial mass median aerodynamic diameter (MMAD) that will decrease with increased distance from the source due to evaporation and settling dependent on environmental parameters such as relative humidity and turbulence.

All mechanisms of human oro-nasal activity such as breathing, talking, laughing, coughing and sneezing produce particles within the inhalable range for humans of <1 to >100 μm (). Significant variation occurs between studies regarding number of particles expelled, size range of the particles and the number of pathogens incorporated within the particles, attributable to differences in methodology and human factors where standardization is difficult.Citation21,Citation26,Citation28,Citation36,Citation37,Citation39 Comparatively, coughing and sneezing produce greater quantities of particlesCitation28,Citation36,Citation37 that travel further due to the velocity of expulsion from the nose or mouth.Citation40 The majority of these particles reside in the inhalable fraction for humans (i.e., <100 μm) at 78.6–96.0% and 98.9% for coughing and sneezing respectively; while of this inhalable fraction, 7.1–46.7% and 18.8% produced by coughing and sneezing were less than 4 μm, evaporating to droplet nuclei and deposit in the bronchoalveolar region of the lung.Citation28,Citation36,Citation37

The situation in relation to deposition is more complex due to evaporation. Atmospheric relative humidity (RH) and temperature are generally lower than that of the body. Once the particles are in the atmosphere evaporation occurs at rates according to their original size and composition of the particle to reach equilibrium with atmospheric conditions. Hence, the aerosol produced is dynamic, changing with distance from the initial point of generation. Particles produced from sneezing and coughing will contain varying amounts of saliva and mucus comprising inorganic and organic ions plus glycoproteins.Citation33,Citation38 Many aerosol transmission modeling studies are based on assumptions e.g., settling velocities and/or evaporation parameters of pure water droplets within a vacuum and droplet distributions from healthy volunteers.Citation21,Citation22 It is probable that irrespective of composition due to the small droplet sizes originating from a cough or sneeze that evaporation will be rapid unless the presence of solutes greatly retards evaporation, for example, a 5 µm water droplet will evaporate within 0.8 s in 97% RH.Citation21 However, extrapolating to natural situations where the droplet composition will be very different and turbulence will exert a large effect on how rapidly particles deposit requires care. Indeed, one recent study demonstrated that infected individuals generated larger aerosol particles than healthy counterparts.Citation41 This could be attributable to differences in mucus (composition, quantity, and viscosity) produced during infection affecting evaporation and the location of the infection.Citation33,Citation42 The closer an individual is situated to an aerosol source then the greater the likelihood of large particles being inhaled prior to complete evaporation.

Similar principles can be extrapolated to any aerosol present in , simply the mechanism of generation, the solute type and concentrations (organic and inorganic) plus the surrounding environment will differ and therefore the processes of evaporation and dissemination will accordingly vary. Irrespective of whether aerosol particles are generated by abiotic or biotic processes (e.g., ), inhalation of particles into the warm humid respiratory tract will prompt rehydration and affect deposition due to particle growth.Citation43 It is evident that there is significant potential for URT deposition depending on how close an individual is to the source.

Initial Site of Deposition and Infection is Dependent on Particle Size

The respiratory tract is complex, comprising a collection of specialized organs, tissues, and cells ranging from the nares to the alveoli.Citation44 These tissues convey a range of physiological functions connected to breathing (i.e., air conditioning, air conductance, and gaseous exchange) and defense against foreign particulates (i.e., immune function and mucociliary or phagocytic clearance). After exposure to a bioaerosol containing pathogens, the initial site of deposition where infection may ensue is likely to be the respiratory epithelium. However, due to the interconnecting anatomical features and clearance mechanisms within the mammalian body, the ocular conjunctiva, olfactory epithelium, or URT immunological tissues may represent further sites where infection could initiate after inhalational exposure to a pathogenic aerosol (; ). Various viruses with dual tropism for both ocular and respiratory tissues utilize the nasolacrimal duct to produce URT infection (e.g., influenza virus, respiratory syncytial virus, and adenovirus).Citation45 Depending on clearance mechanisms and the sensitivity of the pathogen to stomach acidity, gastrointestinal (GI) tissues may represent a further portal. The enteric Norwalk-like virus has been demonstrated to be transmitted by aerosol, presumably via droplets produced during diarrhea and vomiting.Citation46,Citation47 Similarly, aerosol dissemination of Clostridium difficile spores has been observed in hospital wardsCitation48; however this has not been conclusively linked to infection via inhalation.

Table 2. Tissues that may represent the primary site of deposition, clearance and infection during exposure to a bioaerosol

Figure 1. Schematic of the interplay between deposition sites and clearance mechanisms in the respiratory tract. BALT, bronchial-associated lymphoid tissue; LALT, lung-associated lymphoid tissue; LNs, lymph nodes; NALT, nasal-associated lymphoid tissue.

Figure 1. Schematic of the interplay between deposition sites and clearance mechanisms in the respiratory tract. BALT, bronchial-associated lymphoid tissue; LALT, lung-associated lymphoid tissue; LNs, lymph nodes; NALT, nasal-associated lymphoid tissue.

The aerodynamic diameter of inhaled particles determine where within the respiratory tract pathogens incorporated within the particles deposit and interact with host tissues. A number of mechanisms determine deposition of particles within the respiratory tract including inertial impaction, Brownian diffusion, gravitational sedimentation, and electrostatic effects.Citation44 Small particles (<1–3 μm) diffuse deep into the lung tissue, depositing in the alveoli by a number of mechanisms including diffusion, sedimentation, and electrostatic effects. In contrast, larger particles (>8 μm) impact further up the respiratory airways due to greater inertion, depositing in a size-dependent manner from the nasal passages to the larger bronchioles. This relationship is extant across mammalian species albeit differences in respiratory anatomy and physiology dictate the penetration of a particular particle size into the respiratory tract.Citation49-Citation54 The size and body shape of the species determines the morphometry of the nasal cavity and respiratory airways influencing the size of the particles that may deposit in corresponding anatomical regions.Citation51,Citation55 Even within species, variables including age, body weight, breathing mode (oro-nasal/nasal), sex, strain (or ethnicity), activity (e.g., sleeping, exercise), and disease state (e.g., asthma, pneumonia, emphysema, and chronic obstructive pulmonary disorder) influence biometry or affect respiratory physiology and hence deposition profiles.Citation49,Citation56,Citation57 The propensity of humans to revert to oro-nasal breathing during exertion significantly increases the size of particles that may be inhaled into the respiratory tract due to the comparative size of the oral cavity and bypassing the filtration of the nasal cavity.Citation49 Differences in deposition can then affect clearance mechanisms and rates and ultimately infection kinetics for an inhaled pathogen ().

Clearance Mechanisms in the Respiratory Tract

Clearance kinetics are fundamental to determining the dose of deposited pathogens within the respiratory tract and ultimately systemically. A schematic of the interplay between deposition site, clearance mechanisms and pathogen dissemination from the respiratory tract is illustrated in . The nose effectively filters foreign particles that enter the nasal cavity in a manner dependent on particle size and air flow rate with filtration efficiency decreasing with particle size.Citation49 Once deposited, the speed of nasal clearance depends on the deposition location in the nasopharynx. In healthy humans, clearance from the ciliated anterior region is much more rapid than the non-ciliated poster region, ranging from 1.3 to 12.6 mm min−1.Citation58 These rates are comparable to reported tracheal and bronchial mucociliary rates that range from 0.8 to 12.4 mm min−1.Citation59,Citation60 Similarities exist in animal models however clearance rates are generally more rapid due to the decreased distances required to reach the larynx.Citation61

Both the nasal and tracheobronchial escalators comprise mucus that entraps deposited particulates and via the cumulative action of the cilia remove deposited material to the GI tract. Mucus composition is highly variable comprising glycoproteins (mucins), proteins, proteoglycans, and lipids. The quantities of these components present at any particular time govern the viscoelasticity, adhesiveness, and wettability of the mucus and can influence the size of particles emitted by coughing or sneezing.Citation41,Citation42,Citation62 The diversity of oligosaccharide chains present on respiratory mucins and proteoglycans adds complexity providing a mechanism for microbial interaction and clearance.Citation63 Factors such as underlying disease both infectious and non-infectious (e.g., cystic fibrosis, smoking, and diabetes) reduce mucociliary clearance rates,Citation58,Citation59,Citation62,Citation64 hence increasing residence time for deposited pathogens within the respiratory tract.

The pulmonary region is non-ciliated and clearance of foreign particulates is conducted by resident alveolar macrophages that phagocytose particles and transport to the local lung associated lymph nodes and play an important role in pulmonary immunity.Citation65-Citation67 Across species, numbers of alveolar macrophages increase in response to deposited foreign particles, however interspecies variability exists with respect to the rate of chemotaxis and phagocytosis.Citation68

Pathogenesis as a Function of Particle Size in Animal Models

Lung and URT lymphoid tissue infections

No studies exist investigating the effects of particle size on respiratory infection in humans. However, comparative studies in animal models have been conducted using experimental systems that can differentially deposit pathogens in the LRT and URT within small or large particle aerosols.Citation69-Citation77 The general theme running through studies investigating the effects of aerosol particle size on infectivity is that greater numbers of pathogens need to deposit in the URT to produce lethal infection compared with the LRT ( and ). This is likely a function of the mucociliary escalators present in the nasal cavity and tracheobronchial regions clearing material to the gastrointestinal tract. An increased time to death was observed in animals that inhaled large particles that in time-course studies could be related to differences in pathogenesis between infections initiating in the LRT and URT.Citation69-Citation77

Table 3. Influence of aerosol particle size on the respiratory lethal dose values for respiratory pathogens

Table 4. Effect of aerosol particle size and deposition site on infection kinetics and pathology for selected pathogens in animal models of infection

The inhalation of pathogens within small particle aerosols results in the typical disease profile associated with deposition in the alveolar region across all the rodent and NHP models used. Bacterial pathogens such as Francisella tularensis, Burkholderia pseudomallei, Brucella suis, and Yersinia pestis proliferate within the alveolar spaces or alveolar macrophages causing an influx of neutrophils that contribute to a massive cytokine storm resulting in edema and pneumonic consolidation characteristic of primary pneumonia. Eventually tissue destruction leads to dissemination to visceral organs, septic shock and death.Citation69-Citation72,Citation75,Citation76,Citation79-Citation82 B. pseudomallei is further characterized by the formation of abscesses throughout infected tissues and the potential to relapse after completion of antimicrobial therapy.Citation77 B. anthracis endospores are phagocytosed by alveolar macrophages and dendritic cells and trafficked to lung-associated lymph nodes where they germinate and replicate destroying the lymph node before eventually disseminating via the bloodstream resulting in septicemic shock and toxemia.Citation76,Citation83

Deposition of these pathogens within the nasal cavity produces a different disease profile with some similarities across the different bacterial pathogens. Degradation and ulceration of the nasal epithelium precedes infection of the URT lymphoid tissues such as the nasal-associated lymphoid tissue (NALT) and tonsils. Eventually cervical lymphadenitis is observed prior to dissemination to other tissues and septic shock.Citation69-Citation72,Citation75-Citation77 NALT is the rodent equivalent of the Waldemeyer rings in primates comprising the adenoids and tonsils. Lung infection is often observed despite primary deposition in the URT. There are two reasons for this, first a limitation of the aerosol device that produces a polydisperse aerosol comprising some 1–3 µm particles in addition to the larger aerosol particle sizes. These will deposit in the alveoli producing dual infection of the LRT and URT in organisms such as B. pseudomallei that have a very low infectious dose of a few bacteria.Citation77 Second, secondary pneumonia may occur very late in infection in mice that inhaled Y. pestis aerosolized within 12 µm particles resulting from hematogenous spread from the bloodstream into the alveolar spaces presumably after bacteraemic spread from the URT lymphoid tissues.Citation71,Citation75

Gastrointestinal, olfactory, and conjunctival infections from inhalational deposition

Intriguingly, differences in pathogenesis were observed between the bacterial pathogens upon deposition in the URT. In mice that inhaled 12 µm particles containing B. anthracis endospores GI pathology was observed with primary gastritis (17%), and activation and degeneration of GI lymphoid tissues such as the Peyer patches (72%) and mesenteric lymph nodes (67%).Citation76 This pathology was not observed in other species, and is perhaps related to a combined effect of mucociliary clearance of the endospores to the stomach and the hardiness of endospores to the harsh acidic environment of the stomach. Bacteria such as Y. pestis, F. tularensis, and B. pseudomallei that do not produce GI pathology after URT deposition are much more sensitive to low acidity.Citation84-Citation86 Mice that inhaled B. pseudomallei within 12 µm particles demonstrated tropism toward the olfactory epithelium with sequential infection and resultant inflammatory responses within the olfactory neurone and olfactory bulb (100%) culminating in brain abscessation (33%).Citation77 Similar observations were observed in an intranasal infection model.Citation87 Ocular infection characterized by severe conjunctivitis associated with “purulent discharge from the nose and eyes” was observed in 16% of rhesus macaques that inhaled F. tularensis within 12–24 µm particles. In contrast, inhalation of 1–8 µm particles did not produce this pathology.Citation69

In contrast, for the encephalitic alphaviruses, the demarcation between LRT and URT infection is much less marked demonstrating the difficulties in employing rigid health-related demarcations to particle size penetration into the respiratory tract. Mice challenged with 1–3 µm particle aerosols containing VEE virus or after intranasal deposition demonstrated neuroinvasion via trans-synaptic spread through the olfactory or trigeminal neuronal pathway to the brain. The terminal stages of infection were characterized by multifocal necrotising encephalitis.Citation88-Citation92 Intranasal deposition resulted in the presence of higher viral titers in the nasal mucosa, NALT, and cervical lymph nodes.Citation90,Citation93 Similarly, intranasal challenge in the Guinea pig resulted in targeted infection of the olfactory bulb prior to viremia.Citation93 Similar pathogenesis has been observed in the Rhesus macaque, however, the virus localized in the olfactory bulb apparently not progressing to the brain.Citation94,Citation95 In eastern equine encephalitis virus (EEEV) the olfactory neuronal pathway is important for inhalational but not parenteral routes of infection in rodent models.Citation74,Citation92 Utilization of the olfactory neurone to cross the cribriform plate into the central nervous system has also been observed in URT infection models for Nipah virus, Japanese encephalitis virus, Hendra virus, herpes simplex virus, influenza virus (H5N1 subtype), Borna virus, Balamuthia mandrillaris, Naegleria fowleri, B. pseudomallei, Streptococcus pneumoniae, Listeria monocytogenes, and Neisseria meningitidis.Citation77,Citation87,Citation96-Citation105 Interestingly, recently differential phagocytosis of Escherichia coli and Burkholderia thailandensis was observed in vitro by olfactory sheathing and Schwann cells perhaps representing a mechanism for colonization, infection, or clearance.Citation155

These studies demonstrate that deposition site can profoundly influence infection kinetics and pathogenesis within inhalational animal models. Depending on the initial site of deposition and clearance kinetics, aerosolized pathogens may come into contact with a range of tissues and organs through which infection may occur: nasal mucosa, nasal and URT lymphoid tissues, olfactory epithelium, bronchoalveolar epithelium, gastrointestinal tract, and ocular conjunctiva.

Can CDC Select Agents Cause Infection of the URT in Humans?

In the United States, possession and transfer of biothreat agents is regulated under the Select Agent Program (SAP) administered by the Centers for Disease Control and Prevention (CDC). The pathogens and toxins controlled by the SAP are commonly referred to as select agents and represent a potential severe threat to the health and agriculture sectors. This review specifically deals with those select agents that are harmful to humans.

It is difficult to extrapolate animal studies to humans directly because time-course data are not available in humans and therefore the intricacies of pathogenesis cannot be related to respiratory tract deposition. The reported pathology in human cases of inhalational infections is generally from terminal cases or during treatment regimens where the infection is already at a progressed stage and it is difficult to be certain where the infection initiated. It is further complicated by the high mortality, often approaching 90%, for the LRT infections caused by these pathogens.

However, cases of human infections caused by select agents occur that originate in the URT and demonstrate similar pathology to that described in inhalational animal infection models of the URT. Patients present with febrile illness, tonsillitis, pharyngitis, and cervical lymphadenitis prior to septicemia (). These presentations are generally termed pharyngeal or oro-pharyngeal infections and are predominantly associated with consumption of contaminated meats or water for Y. pestis,Citation106-Citation110 B. anthracis,Citation111-Citation118 and F. tularensis.Citation119-Citation123 B. pseudomallei presents as a pharyngocervical infection predominantly in children, however there are adult cases.Citation124-Citation129

Table 5. Upper respiratory tract symptoms in bacterial select agents

Two respiratory forms of plague have been observed from outbreaks of bubonic and pneumonic plague, one that results in primary pneumonia and a second that results in tonsillitis and cervical lymphadenopathy in the absence of pneumonia.Citation130-Citation133 FlexnerCitation131 and CrowellCitation130 respectively stated that “of all the buccal structures, the tonsils seem to be most frequently the one attacked” and “tonsils have formed the portal of entrance for the bacilli and that the involvement of the cervical glands occurs secondarily through the lymph stream”. MeyerCitation133 further implied that “perhaps only the larger particles can lodge in the URT and give rise to tonsillar or septicemic plague”. The described URT pathology closely resembles that observed in NHPs, Guinea pigs, and mice ().Citation71,Citation75,Citation133 Interestingly, Guinea pigs suffering from pneumonic plague demonstrated 17% cross-infection to healthy control animals during cross-infection studies. The cross-infected animals all suffered from the URT infection.Citation75 It was implied that this may be a contributory factor toward the cessation of pneumonic plague outbreaks because the URT form is less infectious and has reduced mortality rates.Citation133 It is noteworthy that for B. anthracis,Citation117,Citation134 B. melitensis,Citation135 B. mallei,Citation136,Citation137 and Coxiella burnetiiCitation138 there are also bona fide cases of inhalational infections that lack LRT involvement and present with the aforementioned pathology indicative of primary infection via the URT. Furthermore, in humans, VEE, WEE, and EEE viruses can infect via the inhalational route and cause neurological sequelae similar to that observed in animal models.Citation139 However it is not known in humans whether the olfactory neurone plays a role in the pathogenesis of these viruses or B. pseudomallei.

These cases demonstrate that in humans these pathogens can produce infection via the URT and although mortality is reduced compared with the respective LRT infections, the rates are not insignificant if untreated (). Similar to the LRT infection, the URT infections caused by select agents are treatable by intervention with antimicrobial therapy, and prognosis is good if promptly administered. Furthermore, the window of opportunity is longer compared with the LRT infection due to the protracted course of infection. However, misdiagnosis is problematic resulting in incorrect treatment regimens because symptoms resemble common URT diseases. For example, similar URT pathogenesis characterized by adenotonsillar disease, pharyngitis and /or cervical adenitis has been observed for a range of common pathogens including respiratory viruses, measles virus, Staphylococcus aureus, Hemophilus influenzae, Streptococcus pyogenes, Streptococcus pneumoniae, and Mycobacterium spp.Citation140-Citation146

Gastrointestinal anthrax, different from the oropharyngeal form, is known to occur in humans due to consumption of contaminated meats with mortality rates reaching 29% if untreated. It presents typically as a febrile illness with severe abdominal pain, mesenteric lymphadenopathy, hemorrhagic ascites, hematemesis, and diarrhea.Citation115 Complications can include hemorrhagic gastritis.Citation147 Interestingly, a recent case of GI anthrax occurred after exposure to endospores aerosolized during drumming, the inference being that the deposited endospores were cleared from the respiratory tract to the GI tract.Citation148 This supports observations in mice where infection of the GI tract occurred in mice that inhaled 12 µm particles containing B. anthracis endospores with pathology similar to that described for humans.Citation76

Conjunctivitis was observed in NHPs exposed to 12–24 µm particles containing F. tularensis.Citation69 In humans, F. tularensis can cause a rare presentation known as oculoglandular syndrome characterized by conjunctivitis, ulceration and preauricular lymphadenitis,Citation149 although this appears to be a result of direct contact with fomites or vector-borne rather than airborne transmission.

Conclusions

It is difficult to predict the particle size that an individual may inhale from a bioaerosol because any source comprises a particle distribution () that changes with time and distance due to the local climate (e.g., meteorology, turbulent activity, ventilation, etc.) and removal from the air column. As such, from the view-point of inhalational infections the respiratory tract should be considered as a continuum with deposition occurring throughout and the probability of infection dependent on the interplay between respiratory physiology, regional dose, clearance kinetics, host-pathogen colonisation mechanisms and immunological response. Therefore, a range of tissues that have perhaps not been thought of as routes of inhalational challenge are brought into consideration including the URT lymphoid tissues, olfactory system, GI tract and potentially the ophthalmic system.

The site of deposition after an inhalational event can affect disease kinetics and pathogenesis, however, the deposition of respiratory pathogens in the lungs will generally result in the more rapid aggressive infection with higher mortality rates. However, evidence exists in both animal models and humans for a number of select agents for URT infections involving the URT lymphoid tissues as the initial foci followed by dissemination via the cervical lymph nodes and bacteremic spread. In addition to LRT presentations, other presentations have been observed in a pathogen specific manner including neurological infection via the olfactory system, GI tract infection, and conjunctivitis.

Increased understanding of the pathogenesis and immunology of infections resulting from inhalation and resultant clearance will aid in the development of vaccine candidates and antimicrobial regimens. Research into host-pathogen interactions and the immunology of URT infections is still in its infancy compared with LRT and systemic infections.Citation67,Citation80,Citation150 However, recent years have seen an increased understanding of host-pathogen and pathogen-pathogen interactions throughout the URT and the function of immunological tissues.Citation151-Citation153 In humans it is unknown whether the olfactory system is utilized as a direct pathway from the URT to the brain. However, it remains a potential route and only recently has the immune response within the olfactory system been researched for virusesCitation154 and the picture is even less clear for bacteria.

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

Acknowledgments

The author recognizes the contribution of Ministry of Defense funding for aspects of work covered by this review. © Crown copyright 2013. Published with the permission of the Defence Science and Technology Laboratory on behalf of the Controller of HMSO.

10.4161/viru.27172

References

  • Jewett DL, Heinsohn P, Bennett C, Rosen A, Neuilly C. Blood-containing aerosols generated by surgical techniques: a possible infectious hazard. Am Ind Hyg Assoc J 1992; 53:228 - 31; http://dx.doi.org/10.1080/15298669291359564; PMID: 1529914
  • Szymańska J. Dental bioaerosol as an occupational hazard in a dentist’s workplace. Ann Agric Environ Med 2007; 14:203 - 7; PMID: 18247451
  • Greene VW, Vesley D, Bond RG, Michaelsen GS. Microbiological contamination of hospital air. I. Quantitative studies. Appl Microbiol 1962; 10:561 - 6; PMID: 13950172
  • Tang JW, Li Y, Eames I, Chan PKS, Ridgway GL. Factors involved in the aerosol transmission of infection and control of ventilation in healthcare premises. J Hosp Infect 2006; 64:100 - 14; http://dx.doi.org/10.1016/j.jhin.2006.05.022; PMID: 16916564
  • Roberts K, Hathway A, Fletcher LA, Beggs CB, Elliott MW, Sleigh PA. Bioaerosol production on a respiratory ward. Indoor Built Environ 2006; 15:35 - 40; http://dx.doi.org/10.1177/1420326X06062562
  • Rothman T, Ledbetter JO. Droplet size of cooling tower fog. Environ Lett 1975; 10:191 - 203; http://dx.doi.org/10.1080/00139307509435821; PMID: 1240052
  • Bausum HT, Schaub SA, Kenyon KF, Small MJ. Comparison of coliphage and bacterial aerosols at a wastewater spray irrigation site. Appl Environ Microbiol 1982; 43:28 - 38; PMID: 7055376
  • Lee S-A, Adhikari A, Grinshpun SA, McKay R, Shukla R, Reponen T. Personal exposure to airborne dust and microorganisms in agricultural environments. J Occup Environ Hyg 2006; 3:118 - 30; http://dx.doi.org/10.1080/15459620500524607; PMID: 16484176
  • Olsen KN, Lund M, Skov J, Christensen LS, Hoorfar J. Detection of Campylobacter bacteria in air samples for continuous real-time monitoring of Campylobacter colonization in broiler flocks. Appl Environ Microbiol 2009; 75:2074 - 8; http://dx.doi.org/10.1128/AEM.02182-08; PMID: 19201953
  • Rathburn CB Jr., Dukes JC. A comparison of the mortality of caged adult mosquitoes to the size, number and volume of ULV spray droplets sampled in an open and a vegetated area. J Am Mosq Control Assoc 1989; 5:173 - 5; PMID: 2746204
  • Mount GA. A critical review of ultralow-volume aerosols of insecticide applied with vehicle-mounted generators for adult mosquito control. J Am Mosq Control Assoc 1998; 14:305 - 34; PMID: 9813829
  • Kim JS, Je YH. Milling effect on the control efficacy of spray-dried Bacillus thuringiensis technical powder against diamondback moths. Pest Manag Sci 2012; 68:321 - 3; http://dx.doi.org/10.1002/ps.2330; PMID: 22413132
  • Rosas-García NM. Laboratory and field tests of spray-dried and granular formulations of a Bacillus thuringiensis strain with insecticidal activity against the sugarcane borer. Pest Manag Sci 2006; 62:855 - 61; http://dx.doi.org/10.1002/ps.1245; PMID: 16786544
  • Teschke K, Chow Y, Bartlett K, Ross A, van Netten C. Spatial and temporal distribution of airborne Bacillus thuringiensis var. kurstaki during an aerial spray program for gypsy moth eradication. Environ Health Perspect 2001; 109:47 - 54; http://dx.doi.org/10.1289/ehp.0110947; PMID: 11171524
  • DiGiovanni F, Kevan PG. Factors affecting pollen dynamics and its importance to pollen contamination: a review. Can J Res 1991; 21:1155 - 70; http://dx.doi.org/10.1139/x91-163
  • Brown JKM, Hovmøller MS. Aerial dispersal of pathogens on the global and continental scales and its impact on plant disease. Science 2002; 297:537 - 41; http://dx.doi.org/10.1126/science.1072678; PMID: 12142520
  • Brandl H, Bachofen R, Bischoff M. Generation of bioaerosols during manual mail unpacking and sorting. J Appl Microbiol 2005; 99:1099 - 107; http://dx.doi.org/10.1111/j.1365-2672.2005.02700.x; PMID: 16238740
  • Barrabeig I, Rovira A, Garcia M, Oliva JM, Vilamala A, Ferrer MD, Sabrià M, Domínguez A. Outbreak of Legionnaires’ disease associated with a supermarket mist machine. Epidemiol Infect 2010; 138:1823 - 8; http://dx.doi.org/10.1017/S0950268810000841; PMID: 20392306
  • Baylor ER, Peters V, Baylor MB. Water-to-air transfer of virus. Science 1977; 197:763 - 4; http://dx.doi.org/10.1126/science.329413; PMID: 329413
  • Baron PA, Willeke K. Respirable droplets from whirlpools: measurements of size distribution and estimation of disease potential. Environ Res 1986; 39:8 - 18; http://dx.doi.org/10.1016/S0013-9351(86)80003-2; PMID: 3943512
  • Morawska L, Johnson GR, Ristovski ZD, Hargreaves M, Mengersen K, Corbett S, et al. Size distribution and sites of origin of droplets expelled from the human respiratory tract during expiratory activities. Aero Sci 2009; 40:256 - 69; http://dx.doi.org/10.1016/j.jaerosci.2008.11.002
  • Chao CYH, Wan MP, Morawska L, Johnson GR, Ritovski ZD, Hargreaves M, et al. Characterization of expiration air jets and droplet size distributions immediately at the mouth opening. Aerosol Sci 2009; 40:122 - 33; http://dx.doi.org/10.1016/j.jaerosci.2008.10.003
  • Xie X, Li Y, Sun H, Liu L. Exhaled droplets due to talking and coughing. J R Soc Interface 2009; 6:Suppl 6 S703 - 14; http://dx.doi.org/10.1098/rsif.2009.0388.focus; PMID: 19812073
  • Lai KM, Bottomley C, McNerney R. Propagation of respiratory aerosols by the vuvuzela. PLoS One 2011; 6:e20086; http://dx.doi.org/10.1371/journal.pone.0020086; PMID: 21629778
  • Yang S, Lee GWM, Chen C-M, Wu C-C, Yu K-P. The size and concentration of droplets generated by coughing in human subjects. J Aerosol Med 2007; 20:484 - 94; http://dx.doi.org/10.1089/jam.2007.0610; PMID: 18158720
  • Fennelly KP, Martyny JW, Fulton KE, Orme IM, Cave DM, Heifets LB. Cough-generated aerosols of Mycobacterium tuberculosis: a new method to study infectiousness. Am J Respir Crit Care Med 2004; 169:604 - 9; http://dx.doi.org/10.1164/rccm.200308-1101OC; PMID: 14656754
  • Lindsley WG, Blachere FM, Thewlis RE, Vishnu A, Davis KA, Cao G, Palmer JE, Clark KE, Fisher MA, Khakoo R, et al. Measurements of airborne influenza virus in aerosol particles from human coughs. PLoS One 2010; 5:e15100; http://dx.doi.org/10.1371/journal.pone.0015100; PMID: 21152051
  • Duguid JP. The numbers and the sites of origin of the droplets expelled during expiratory activities. Edinb Med J 1945; 52:385 - 401; PMID: 21009905
  • Jennison MW. Atomizing of mouth and nose secretions into the air as revealed by high-speed photography. Aerobiol 1942; 17:106 - 28
  • Zhou Y, Benson JM, Irvin C, Irshad H, Cheng Y-S. Particle size distribution and inhalation dose of shower water under selected operating conditions. Inhal Toxicol 2007; 19:333 - 42; http://dx.doi.org/10.1080/08958370601144241; PMID: 17365038
  • Blatny JM, Fossum H, Ho J, Tutkun M, Skogan G, Andreassen O, Fykse EM, Waagen V, Reif BA. Dispersion of Legionella-containing aerosols from a biological treatment plant, Norway. Front Biosci (Elite Ed) 2011; 3:1300 - 9; PMID: 21622136
  • Ruiz A, Bulmer GS. Particle size of airborn Cryptococcus neoformans in a tower. Appl Environ Microbiol 1981; 41:1225 - 9; PMID: 7020595
  • Gralton J, Tovey E, McLaws M-L, Rawlinson WD. The role of particle size in aerosolised pathogen transmission: a review. J Infect 2011; 62:1 - 13; http://dx.doi.org/10.1016/j.jinf.2010.11.010; PMID: 21094184
  • Fernstrom A, Goldblatt M. Aerobiology and its role in the transmission of infectious diseases. J Pathog 2013; 2013:493960; http://dx.doi.org/10.1155/2013/493960; PMID: 23365758
  • Williams CG. Long-distance pine pollen still germinates after meso-scale dispersal. Am J Bot 2010; 97:846 - 55; http://dx.doi.org/10.3732/ajb.0900255; PMID: 21622450
  • Loudon RG, Roberts RM. Droplet expulsion from the respiratory tract. Am Rev Respir Dis 1967; 95:435 - 42; PMID: 6018703
  • Loudon RG, Roberts RM. Relation between the airborne diameters of respiratory droplets and the diameter of the stains left after recovery. Nature 1967; 213:95 - 6; http://dx.doi.org/10.1038/213095a0
  • Nicas M, Nazaroff WW, Hubbard A. Toward understanding the risk of secondary airborne infection: emission of respirable pathogens. J Occup Environ Hyg 2005; 2:143 - 54; http://dx.doi.org/10.1080/15459620590918466; PMID: 15764538
  • Papineni RS, Rosenthal FS. The size distribution of droplets in the exhaled breath of healthy human subjects. J Aerosol Med 1997; 10:105 - 16; http://dx.doi.org/10.1089/jam.1997.10.105; PMID: 10168531
  • Zhou B, Zhang Z, Li X. Numerical study of the transport of droplets or particles generated by respiratory system indoors. Build Environ 2005; 40:1032 - 9; http://dx.doi.org/10.1016/j.buildenv.2004.09.018
  • Hersen G, Moularat S, Robine E, Ghin E, Corbet S, Vabret A. Impact of health on particle size of exhaled respiratory aerosols: case-control study. Clean 2008; 36:572 - 7
  • Girod S, Zahm J-M, Plotkowski C, Beck G, Puchelle E. Role of the physiochemical properties of mucus in the protection of the respiratory epithelium. Eur Respir J 1992; 5:477 - 87; PMID: 1563506
  • Finlay WH, Stapleton KW, Chan HK, Zuberbuhler P, Gonda I. Regional deposition of inhaled hygroscopic aerosols: in vivo SPECT compared with mathematical modeling. J Appl Physiol (1985) 1996; 81:374 - 83; PMID: 8828688
  • Wang C-S. Inhaled Particles. Interface Sci Tech 2005; 5:1 - 187
  • Belser JA, Rota PA, Tumpey TM. Ocular tropism of respiratory viruses. Microbiol Mol Biol Rev 2013; 77:144 - 56; http://dx.doi.org/10.1128/MMBR.00058-12; PMID: 23471620
  • Caul EO. Small round structured viruses: airborne transmission and hospital control. Lancet 1994; 343:1240 - 2; http://dx.doi.org/10.1016/S0140-6736(94)92146-6; PMID: 7910270
  • Marks PJ, Vipond IB, Carlisle D, Deakin D, Fey RE, Caul EO. Evidence for airborne transmission of Norwalk-like virus (NLV) in a hotel restaurant. Epidemiol Infect 2000; 124:481 - 7; http://dx.doi.org/10.1017/S0950268899003805; PMID: 10982072
  • Roberts K, Smith CF, Snelling AM, Kerr KG, Banfield KR, Sleigh PA, Beggs CB. Aerial dissemination of Clostridium difficile spores. BMC Infect Dis 2008; 8:7; http://dx.doi.org/10.1186/1471-2334-8-7; PMID: 18218089
  • ICRP. Annals of the ICRP – Human respiratory tract model for radiological protection. In. Smith H, (ed), New York. Pergamon. 1994.
  • Heyder J, Gebhart J, Rudolf G, Schiller CF, Stahlhofen W. Deposition of particles in the human respiratory tract in the size range 0.005-15 µm. J Aerosol Sci 1986; 5:811 - 25; http://dx.doi.org/10.1016/0021-8502(86)90035-2
  • Raabe OG, Al-Bayati MA, Teague SV, Rasolt A. Regional deposition of inhaled monodisperse coarse and fine aerosol particles in small laboratory animals. Ann Occup Hyg 1988; 3:53 - 63; http://dx.doi.org/10.1093/annhyg/32.inhaled_particles_VI.53
  • Schlesinger RB. Comparative deposition of inhaled aerosols in experimental animals and humans: a review. J Toxicol Environ Health 1985; 15:197 - 214; http://dx.doi.org/10.1080/15287398509530647; PMID: 3892021
  • Ménache MG, Miller FJ, Raabe OG. Particle inhalability curves for humans and small laboratory animals. Ann Occup Hyg 1995; 39:317 - 28; PMID: 7793751
  • Kuehl PJ, Anderson TL, Candelaria G, Gershman B, Harlin K, Hesterman JY, Holmes T, Hoppin J, Lackas C, Norenberg JP, et al. Regional particle size dependent deposition of inhaled aerosols in rats and mice. Inhal Toxicol 2012; 24:27 - 35; http://dx.doi.org/10.3109/08958378.2011.632787; PMID: 22145784
  • Schreider JP, Raabe OG. Anatomy of the nasal-pharyngeal airway of experimental animals. Anat Rec 1981; 200:195 - 205; http://dx.doi.org/10.1002/ar.1092000208; PMID: 7270920
  • Oldham MJ, Phalen RF. Dosimetry implications of upper tracheobronchial airway anatomy in two mouse varieties. Anat Rec 2002; 268:59 - 65; http://dx.doi.org/10.1002/ar.10134; PMID: 12209565
  • Schulz H, Johner C, Eder G, Ziesenis A, Reitmeier P, Heyder J, Balling R. Respiratory mechanics in mice: strain and sex specific differences. Acta Physiol Scand 2002; 174:367 - 75; http://dx.doi.org/10.1046/j.1365-201x.2002.00955.x; PMID: 11942924
  • Puchelle E, Aug F, Zahm JM, Bertrand A. Comparison of nasal and bronchial mucociliary clearance in young non-smokers. Clin Sci (Lond) 1982; 62:13 - 6; PMID: 7056029
  • Yeates DB, Aspin N, Levison H, Jones MT, Bryan AC. Mucociliary tracheal transport rates in man. J Appl Physiol 1975; 39:487 - 95; PMID: 240802
  • Foster WM, Langenback E, Bergofsky EH. Measurement of tracheal and bronchial mucus velocities in man: relation to lung clearance. J Appl Physiol Respir Environ Exerc Physiol 1980; 48:965 - 71; PMID: 7380708
  • Lippmann M, Schlesinger RB. Interspecies comparisons of particle deposition and mucociliary clearance in tracheobronchial airways. J Toxicol Environ Health 1984; 13:441 - 69; http://dx.doi.org/10.1080/15287398409530509; PMID: 6376822
  • Beule AG. Physiology and pathophysiology of respiratory mucosa of the nose and the paranasal sinuses. GMS Current Topics in Otorhinology –. Head Neck Surg 2010; 9:1 - 24
  • Lamblin G, Aubert JP, Perini JM, Klein A, Porchet N, Degand P, Roussel P. Human respiratory mucins. Eur Respir J 1992; 5:247 - 56; PMID: 1559590
  • Regnis JA, Robinson M, Bailey DL, Cook P, Hooper P, Chan H-K, Gonda I, Bautovich G, Bye PT. Mucociliary clearance in patients with cystic fibrosis and in normal subjects. Am J Respir Crit Care Med 1994; 150:66 - 71; http://dx.doi.org/10.1164/ajrccm.150.1.8025774; PMID: 8025774
  • Sherman MP, Ganz T. Host defense in pulmonary alveoli. Annu Rev Physiol 1992; 54:331 - 50; http://dx.doi.org/10.1146/annurev.ph.54.030192.001555; PMID: 1562178
  • Yoshida M, Whitsett JA. Interactions between pulmonary surfactant and alveolar macrophages in the pathogenesis of lung disease. Cell Mol Biol (Noisy-le-grand) 2004; 50 Online Pub:OL639 - 48; PMID: 15579257
  • Eddens T, Kolls JK. Host defenses against bacterial lower respiratory tract infection. Curr Opin Immunol 2012; 24:424 - 30; http://dx.doi.org/10.1016/j.coi.2012.07.005; PMID: 22841348
  • Warheit DB, Hartsky MA. Role of alveolar macrophage chemotaxis and phagocytosis in pulmonary clearance responses to inhaled particles: comparisons among rodent species. Microsc Res Tech 1993; 26:412 - 22; http://dx.doi.org/10.1002/jemt.1070260509; PMID: 8286787
  • Day WC, Berendt RF. Experimental tularemia in Macaca mulatta: relationship of aerosol particle size to the infectivity of airborne Pasteurella tularensis. Infect Immun 1972; 5:77 - 82; PMID: 4632469
  • Druett HA, Henderson DW, Packman LP, Peacock S. Studies on respiratory infection. I. The influence of particle size on respiratory infection with anthrax spores. J Hyg (Lond) 1953; 51:359 - 71; http://dx.doi.org/10.1017/S0022172400015795; PMID: 13096744
  • Druett HA, Robinson JM, Henderson DW, Packman L, Peacock S. Studies on respiratory infection. II. The influence of aerosol particle size on infection of the guinea-pig with Pasteurella pestis.. J Hyg (Lond) 1956; 54:37 - 48; http://dx.doi.org/10.1017/S0022172400044284; PMID: 13319689
  • Druett HA, Henderson DW, Peacock S. Studies on respiratory infection. III. Experiments with Brucella suis.. J Hyg (Lond) 1956; 54:49 - 57; http://dx.doi.org/10.1017/S0022172400044296; PMID: 13319690
  • Fothergill LD. Biological warfare – Nature and consequences. Tex Med 1964; 60:8 - 14; PMID: 14115387
  • Roy CJ, Reed DS, Wilhelmsen CL, Hartings J, Norris S, Steele KE. Pathogenesis of aerosolized Eastern Equine Encephalitis virus infection in guinea pigs. Virol J 2009; 6:170 - 83; http://dx.doi.org/10.1186/1743-422X-6-170; PMID: 19852817
  • Thomas RJ, Webber D, Collinge A, Stagg AJ, Bailey SC, Nunez A, Gates A, Jayasekera PN, Taylor RR, Eley S, et al. Different pathologies but equal levels of responsiveness to the recombinant F1 and V antigen vaccine and ciprofloxacin in a murine model of plague caused by small- and large-particle aerosols. Infect Immun 2009; 77:1315 - 23; http://dx.doi.org/10.1128/IAI.01473-08; PMID: 19188359
  • Thomas RJ, Davies C, Nunez A, Hibbs S, Flick-Smith H, Eastaugh L, Smither S, Gates A, Oyston P, Atkins T, et al. Influence of particle size on the pathology and efficacy of vaccination in a murine model of inhalational anthrax. J Med Microbiol 2010; 59:1415 - 27; http://dx.doi.org/10.1099/jmm.0.024117-0; PMID: 20798216
  • Thomas RJ, Davies C, Nunez A, Hibbs S, Eastaugh L, Harding S, et al. Particle-size dependent effects in the Balb/c murine model of inhalational melioidosis. Front Cell Infect Microbiol 2012; 2:1 - 12; http://dx.doi.org/10.3389/fcimb.2012.00101; PMID: 22919593
  • Roy CJ, Hale M, Hartings JM, Pitt L, Duniho S. Impact of inhalation exposure modality and particle size on the respiratory deposition of ricin in BALB/c mice. Inhal Toxicol 2003; 15:619 - 38; PMID: 12692733
  • Agar SL, Sha J, Foltz SM, Erova TE, Walberg KG, Parham TE, Baze WB, Suarez G, Peterson JW, Chopra AK. Characterization of a mouse model of plague after aerosolization of Yersinia pestis CO92. Microbiology 2008; 154:1939 - 48; http://dx.doi.org/10.1099/mic.0.2008/017335-0; PMID: 18599822
  • Weiner ZP, Glomski IJ. Updating perspectives on the initiation of Bacillus anthracis growth and dissemination through its host. Infect Immun 2012; 80:1626 - 33; http://dx.doi.org/10.1128/IAI.06061-11; PMID: 22354031
  • West TE, Myers ND, Liggitt HD, Skerrett SJ. Murine pulmonary infection and inflammation induced by inhalation of Burkholderia pseudomallei.. Int J Exp Pathol 2012; 93:421 - 8; PMID: 23136994
  • Henning LN, Miller SM, Pak DH, Lindsay A, Fisher DA, Barnewall RE, Briscoe CM, Anderson MS, Warren RL. Pathophysiology of the rhesus macaque model for inhalational brucellosis. Infect Immun 2012; 80:298 - 310; http://dx.doi.org/10.1128/IAI.05878-11; PMID: 22064715
  • Twenhafel NA. Pathology of inhalational anthrax animal models. Vet Pathol 2010; 47:819 - 30; http://dx.doi.org/10.1177/0300985810378112; PMID: 20656900
  • Butler T, Fu YS, Furman L, Almeida C, Almeida A. Experimental Yersinia pestis infection in rodents after intragastric inoculation and ingestion of bacteria. Infect Immun 1982; 36:1160 - 7; PMID: 7095845
  • KuoLee R, Zhao X, Austin J, Harris G, Conlan JW, Chen W. Mouse model of oral infection with virulent type A Francisella tularensis. Infect Immun 2007; 75:1651 - 60; http://dx.doi.org/10.1128/IAI.01834-06; PMID: 17242058
  • West TE, Myers ND, Limmathurotsakul D, Liggitt HD, Chantratita N, Peacock SJ, Skerrett SJ. Pathogenicity of high-dose enteral inoculation of Burkholderia pseudomallei to mice. Am J Trop Med Hyg 2010; 83:1066 - 9; http://dx.doi.org/10.4269/ajtmh.2010.10-0306; PMID: 21036839
  • Owen SJ, Batzloff M, Chehrehasa F, Meedeniya A, Casart Y, Logue CA, Hirst RG, Peak IR, Mackay-Sim A, Beacham IR. Nasal-associated lymphoid tissue and olfactory epithelium as portals of entry for Burkholderia pseudomallei in murine melioidosis. J Infect Dis 2009; 199:1761 - 70; http://dx.doi.org/10.1086/599210; PMID: 19456230
  • Charles PC, Walters E, Margolis F, Johnston RE. Mechanism of neuroinvasion of Venezuelan equine encephalitis virus in the mouse. Virology 1995; 208:662 - 71; http://dx.doi.org/10.1006/viro.1995.1197; PMID: 7747437
  • Ryzhikov AB, Ryabchikova EI, Sergeev AN, Tkacheva NV. Spread of Venezuelan equine encephalitis virus in mice olfactory tract. Arch Virol 1995; 140:2243 - 54; http://dx.doi.org/10.1007/BF01323243; PMID: 8572944
  • Vogel P, Abplanalp D, Kell W, Ibrahim MS, Downs MB, Pratt WD, Davis KJ. Venezuelan equine encephalitis in BALB/c mice: kinetic analysis of central nervous system infection following aerosol or subcutaneous inoculation. Arch Pathol Lab Med 1996; 120:164 - 72; PMID: 8712896
  • Steele KE, Davis KJ, Stephan K, Kell W, Vogel P, Hart MK. Comparative neurovirulence and tissue tropism of wild-type and attenuated strains of Venezuelan equine encephalitis virus administered by aerosol in C3H/HeN and BALB/c mice. Vet Pathol 1998; 35:386 - 97; http://dx.doi.org/10.1177/030098589803500508; PMID: 9754544
  • Steele KE, Twenhafel NA. REVIEW PAPER: pathology of animal models of alphavirus encephalitis. Vet Pathol 2010; 47:790 - 805; http://dx.doi.org/10.1177/0300985810372508; PMID: 20551475
  • Danes L, Rychterová V, Kliment V, Hrusková J. Penetration of Venezuelan equine encephalomyelitis virus into the brain of guinea pigs and rabbits after intranasal infection. Acta Virol 1973; 17:138 - 46; PMID: 4144277
  • Danes L, Kufner J, Hrusková J, Rychterová V. The role of the olfactory route on infection of the respiratory tract with Venezuelan equine encephalomyelitis virus in normal and operated Macaca rhesus monkeys. I. Results of virological examination. Acta Virol 1973; 17:50 - 6; PMID: 4405396
  • Danes L, Rychterová V, Kufner J, Hrusková J. The role of the olfactory route on infection of the respiratory tract with Venezuelan equine encephalomyelitis virus in normal and operated Macaca rhesus monkeys. II. Results of histological examination. Acta Virol 1973; 17:57 - 60; PMID: 4405397
  • Jarolim KL, McCosh JK, Howard MJ, John DT. A light microscopy study of the migration of Naegleria fowleri from the nasal submucosa to the central nervous system during the early stage of primary amebic meningoencephalitis in mice. J Parasitol 2000; 86:50 - 5; PMID: 10701563
  • Jin Y, Dons L, Kristensson K, Rottenberg ME. Neural route of cerebral Listeria monocytogenes murine infection: role of immune response mechanisms in controlling bacterial neuroinvasion. Infect Immun 2001; 69:1093 - 100; http://dx.doi.org/10.1128/IAI.69.2.1093-1100.2001; PMID: 11160006
  • Kiderlen AF, Laube U. Balamuthia mandrillaris, an opportunistic agent of granulomatous amebic encephalitis, infects the brain via the olfactory nerve pathway. Parasitol Res 2004; 94:49 - 52; http://dx.doi.org/10.1007/s00436-004-1163-z; PMID: 15338290
  • van Ginkel FW, McGhee JR, Watt JM, Campos-Torres A, Parish LA, Briles DE. Pneumococcal carriage results in ganglioside-mediated olfactory tissue infection. Proc Natl Acad Sci U S A 2003; 100:14363 - 7; http://dx.doi.org/10.1073/pnas.2235844100; PMID: 14610280
  • Yamada M, Nakamura K, Yoshii M, Kaku Y, Narita M. Brain lesions induced by experimental intranasal infection of Japanese encephalitis virus in piglets. J Comp Pathol 2009; 141:156 - 62; http://dx.doi.org/10.1016/j.jcpa.2009.04.006; PMID: 19523649
  • Mori I, Nishiyama Y, Yokochi T, Kimura Y. Olfactory transmission of neurotropic viruses. J Neurovirol 2005; 11:129 - 37; http://dx.doi.org/10.1080/13550280590922793; PMID: 16036791
  • Sjölinder H, Jonsson A-B. Olfactory nerve--a novel invasion route of Neisseria meningitidis to reach the meninges. PLoS One 2010; 5:e14034; http://dx.doi.org/10.1371/journal.pone.0014034; PMID: 21124975
  • Dups J, Middleton D, Yamada M, Monaghan P, Long F, Robinson R, Marsh GA, Wang LF. A new model for Hendra virus encephalitis in the mouse. PLoS One 2012; 7:e40308; http://dx.doi.org/10.1371/journal.pone.0040308; PMID: 22808132
  • Munster VJ, Prescott JB, Bushmaker T, Long D, Rosenke R, Thomas T, Scott D, Fischer ER, Feldmann H, de Wit E. Rapid Nipah virus entry into the central nervous system of hamsters via the olfactory route. Sci Rep 2012; 2:736; http://dx.doi.org/10.1038/srep00736; PMID: 23071900
  • Schrauwen EJ, Herfst S, Leijten LM, van Run P, Bestebroer TM, Linster M, et al. The multibasic cleavage site in H5N1 virus is critical for systemic spread along the olfactory and hematogenous routes in ferrets. J Virol 2012; 86:3795 - 808; http://dx.doi.org/10.1128/JVI.06828-11; PMID: 22258251
  • Marshall JD Jr., Quy DV, Gibson FL. Asymptomatic pharyngeal plague infection in Vietnam. Am J Trop Med Hyg 1967; 16:175 - 7; PMID: 6021740
  • Conrad FG, LeCocq FR, Krain R. A recent epidemic of plague in Vietnam. Arch Intern Med 1968; 122:193 - 8; http://dx.doi.org/10.1001/archinte.1968.00300080001001; PMID: 5691443
  • Arbaji A, Kharabsheh S, Al-Azab S, Al-Kayed M, Amr ZS, Abu Baker M, Chu MC. A 12-case outbreak of pharyngeal plague following the consumption of camel meat, in north-eastern Jordan. Ann Trop Med Parasitol 2005; 99:789 - 93; http://dx.doi.org/10.1179/136485905X65161; PMID: 16297292
  • Bin Saeed AA, Al-Hamdan NA, Fontaine RE. Plague from eating raw camel liver. Emerg Infect Dis 2005; 11:1456 - 7; http://dx.doi.org/10.3201/eid1109.050081; PMID: 16229781
  • Leslie T, Whitehouse CA, Yingst S, Baldwin C, Kakar F, Mofleh J, Hami AS, Mustafa L, Omar F, Ayazi E, et al. Outbreak of gastroenteritis caused by Yersinia pestis in Afghanistan. Epidemiol Infect 2011; 139:728 - 35; http://dx.doi.org/10.1017/S0950268810001792; PMID: 20663260
  • Sirisanthana T, Navachareon N, Tharavichitkul P, Sirisanthana V, Brown AE. Outbreak of oral-oropharyngeal anthrax: an unusual manifestation of human infection with Bacillus anthracis.. Am J Trop Med Hyg 1984; 33:144 - 50; PMID: 6696173
  • Doğanay M, Almaç A, Hanağasi R. Primary throat anthrax. A report of six cases. Scand J Infect Dis 1986; 18:415 - 9; http://dx.doi.org/10.3109/00365548609032357; PMID: 3775269
  • Navacharoen N, Sirisanthana T, Navacharoen W, Ruckphaopunt K. Oropharyngeal anthrax. J Laryngol Otol 1985; 99:1293 - 5; http://dx.doi.org/10.1017/S002221510009856X; PMID: 3934300
  • Sirisanthana T, Brown AE. Anthrax of the gastrointestinal tract. Emerg Infect Dis 2002; 8:649 - 51; http://dx.doi.org/10.3201/eid0807.020062; PMID: 12095428
  • Beatty ME, Ashford DA, Griffin PM, Tauxe RV, Sobel J. Gastrointestinal anthrax: review of the literature. Arch Intern Med 2003; 163:2527 - 31; http://dx.doi.org/10.1001/archinte.163.20.2527; PMID: 14609791
  • Babamahmoodi F, Aghabarari F, Arjmand A, Ashrafi GH. Three rare cases of anthrax arising from the same source. J Infect 2006; 53:175 - 9; http://dx.doi.org/10.1016/j.jinf.2005.12.018; PMID: 16376990
  • Holty JE, Kim RY, Bravata DM. Anthrax: a systematic review of atypical presentations. Ann Emerg Med 2006; 48:200 - 11; http://dx.doi.org/10.1016/j.annemergmed.2005.11.035; PMID: 16857469
  • Ozdemir H, Demirdag K, Ozturk T, Kocakoc E. Anthrax of the gastrointestinal tract and oropharynx: CT findings. Emerg Radiol 2010; 17:161 - 4; http://dx.doi.org/10.1007/s10140-009-0821-y; PMID: 19499256
  • Arikan OK, Koç C, Bozdoğan O. Tularemia presenting as tonsillopharyngitis and cervical lymphadenitis: a case report and review of the literature. Eur Arch Otorhinolaryngol 2003; 260:298 - 300; http://dx.doi.org/10.1007/s00405-002-0565-8; PMID: 12883950
  • Kandemir B, Erayman I, Bitirgen M, Aribas ET, Guler S. Tularemia presenting with tonsillopharyngitis and cervical lymphadenitis: report of two cases. Scand J Infect Dis 2007; 39:620 - 2; http://dx.doi.org/10.1080/00365540601105814; PMID: 17577830
  • Chitadze N, Kuchuloria T, Clark DV, Tsertsvadze E, Chokheli M, Tsertsvadze N, Trapaidze N, Lane A, Bakanidze L, Tsanava S, et al. Water-borne outbreak of oropharyngeal and glandular tularemia in Georgia: investigation and follow-up. Infection 2009; 37:514 - 21; http://dx.doi.org/10.1007/s15010-009-8193-5; PMID: 19826763
  • Dlugaiczyk J, Harrer T, Zwerina J, Traxdorf M, Schwarz S, Splettstoesser W, Geissdörfer W, Schoerner C. Oropharyngeal tularemia--a differential diagnosis of tonsillopharyngitis and cervical lymphadenitis. Wien Klin Wochenschr 2010; 122:110 - 4; http://dx.doi.org/10.1007/s00508-009-1274-8; PMID: 20213378
  • Uyar M, Cengiz B, Unlü M, Celebi B, Kılıç S, Eryılmaz A. [Evaluation of the oropharyngeal tularemia cases admitted to our hospital from the provinces of Central Anatolia]. Mikrobiyol Bul 2011; 45:58 - 66; PMID: 21341160
  • Pongrithsukda V, Simakachorn N, Pimda J. Childhood melioidosis in northeastern Thailand. Southeast Asian J Trop Med Public Health 1988; 19:309 - 16; PMID: 3227408
  • Lumbiganon P, Viengnondha S. Clinical manifestations of melioidosis in children. Pediatr Infect Dis J 1995; 14:136 - 40; http://dx.doi.org/10.1097/00006454-199502000-00010; PMID: 7746696
  • Tan NG, Sethi DS. An unusual case of sorethroat: nasopharyngeal melioidosis. Singapore Med J 1997; 38:223 - 5; PMID: 9259605
  • Lim WK, Gurdeep GS, Norain K. Melioidosis of the head and neck. Med J Malaysia 2001; 56:471 - 7; PMID: 12014768
  • How H-S, Ng K-H, Yeo H-B, Tee H-P, Shah A. Pediatric melioidosis in Pahang, Malaysia. J Microbiol Immunol Infect 2005; 38:314 - 9; PMID: 16211138
  • Brent AJ, Matthews PC, Dance DA, Pitt TL, Handy R. Misdiagnosing melioidosis. Emerg Infect Dis 2007; 13:349 - 51; http://dx.doi.org/10.3201/eid1302.061290; PMID: 17479916
  • Crowell BC. Pathologic anatomy of bubonic plague. Philipp J Sci 1915; 10B:249 - 308
  • Flexner S. The pathology of bubonic plague. Am J Med Sci 1901; 122:396 - 416
  • Wu L-T, Chun WH, Pollitzer R. Clinical observations upon the second Manchurian plague epidemic, 1920-1921. Nat Med J China 1922; 8:225 - 55
  • Meyer KF. Pneumonic plague. Bacteriol Rev 1961; 25:249 - 61; PMID: 14473142
  • Krauss F. Anthrax appearing primarily in the nose. Arch Otolaryngol 1943; 37:238 - 41; http://dx.doi.org/10.1001/archotol.1943.00670030247008
  • Zachou K, Papamichalis PA, Dalekos GN. Severe pharyngitis in stockbreeders: an unusual presentation of brucellosis. Occup Med (Lond) 2008; 58:305 - 7; http://dx.doi.org/10.1093/occmed/kqn020; PMID: 18397911
  • Howe C, Miller WR. Human glanders; report of six cases. Ann Intern Med 1947; 26:93 - 115; http://dx.doi.org/10.7326/0003-4819-26-1-93; PMID: 20278465
  • Gregory BC, Waag DM. Glanders. In. Lenhart MK, Lounsbury DE, Martin JW, ed(s). Medical Aspects of Biological Warfare. New York. Borden Institute. 2007:121-46.
  • Pappas G, Giannoutsos C, Christou L, Tsianos E. Coxiella burnetii: an unusual ENT pathogen. Am J Otolaryngol 2004; 25:263 - 5; http://dx.doi.org/10.1016/j.amjoto.2004.01.006; PMID: 15239034
  • Steele KE, Reed DS, Glass PJ, Hart MK, Ludwig GV, Pratt WD, et al. Alphavirus encephalitides. In: Lenhart, MK, Lounsbury, DE, Martin, JW, ed(s). Medical Aspects of Biological Warfare. New York. Borden Institute. 2007:241-70.
  • Lau SK, Kwan S, Lee J, Wei WI. Source of tubercle bacilli in cervical lymph nodes: a prospective study. J Laryngol Otol 1991; 105:558 - 61; http://dx.doi.org/10.1017/S0022215100116603; PMID: 1908506
  • Al-Serhani AM. Mycobacterial infection of the head and neck: presentation and diagnosis. Laryngoscope 2001; 111:2012 - 6; http://dx.doi.org/10.1097/00005537-200111000-00027; PMID: 11801988
  • Brook I. Current management of upper respiratory tract and head and neck infections. Eur Arch Otorhinolaryngol 2009; 266:315 - 23; http://dx.doi.org/10.1007/s00405-008-0849-8; PMID: 18985371
  • Zautner AE. Adenotonsillar disease. Recent Pat Inflamm Allergy Drug Discov 2012; 6:121 - 9; http://dx.doi.org/10.2174/187221312800166877; PMID: 22452646
  • Proenca-Modena JL, Pereira Valera FC, Jacob MG, Buzatto GP, Saturno TH, Lopes L, Souza JM, Escremim Paula F, Silva ML, Carenzi LR, et al. High rates of detection of respiratory viruses in tonsillar tissues from children with chronic adenotonsillar disease. PLoS One 2012; 7:e42136; http://dx.doi.org/10.1371/journal.pone.0042136; PMID: 22870291
  • Ludlow M, de Vries RD, Lemon K, McQuaid S, Millar E, van Amerongen G, Yüksel S, Verburgh RJ, Osterhaus AD, de Swart RL, et al. Infection of lymphoid tissues in the macaque upper respiratory tract contributes to the emergence of transmissible measles virus. J Gen Virol 2013; 94:1933 - 44; http://dx.doi.org/10.1099/vir.0.054650-0; PMID: 23784446
  • Munck K, Mandpe AH. Mycobacterial infections of the head and neck. Otolaryngol Clin North Am 2003; 36:569 - 76; http://dx.doi.org/10.1016/S0030-6665(03)00032-X; PMID: 14567053
  • Dutz W, Saidi F, Kohout E. Gastric anthrax with massive ascites. Gut 1970; 11:352 - 4; http://dx.doi.org/10.1136/gut.11.4.352; PMID: 5428857
  • Mayo L, Dionne-Odom J, Talbot EA, Adamski C, Bean C, Daly ER, et al, Centers for Disease Control and Prevention (CDC). Gastrointestinal anthrax after an animal-hide drumming event - New Hampshire and Massachusetts, 2009. MMWR Morb Mortal Wkly Rep 2010; 59:872 - 7; PMID: 20651643
  • Halperin SA, Gast T, Ferrieri P. Oculoglandular syndrome caused by Francisella tularensis.. Clin Pediatr (Phila) 1985; 24:520 - 2; http://dx.doi.org/10.1177/000992288502400909; PMID: 4017402
  • Li B, Yang R. Interaction between Yersinia pestis and the host immune system. Infect Immun 2008; 76:1804 - 11; http://dx.doi.org/10.1128/IAI.01517-07; PMID: 18250178
  • Brandtzaeg P. Immune functions of nasopharyngeal lymphoid tissue. Adv Otorhinolaryngol 2011; 72:20 - 4; http://dx.doi.org/10.1159/000324588; PMID: 21865681
  • Bosch AA, Biesbroek G, Trzcinski K, Sanders EAM, Bogaert D. Viral and bacterial interactions in the upper respiratory tract. PLoS Pathog 2013; 9:e1003057; http://dx.doi.org/10.1371/journal.ppat.1003057; PMID: 23326226
  • Mina MJ, Klugman KP. Pathogen replication, host inflammation, and disease in the upper respiratory tract. Infect Immun 2013; 81:625 - 8; http://dx.doi.org/10.1128/IAI.01460-12; PMID: 23319561
  • Kalinke U, Bechmann I, Detje CN. Host strategies against virus entry via the olfactory system. Virulence 2011; 2:367 - 70; http://dx.doi.org/10.4161/viru.2.4.16138; PMID: 21758005
  • Panni P, Ferguson IA, Beacham I, Mackay-Sim A, Ekberg JAK, St John JA. Phagocytosis of bacteria by olfactory ensheathing cells and Schwann cells. Neurosci Lett 2013; 539:65 - 70; http://dx.doi.org/10.1016/j.neulet.2013.01.052; PMID: 23415759