gms | German Medical Science

GMS Current Topics in Otorhinolaryngology - Head and Neck Surgery

German Society of Oto-Rhino-Laryngology, Head and Neck Surgery (DGHNOKHC)

ISSN 1865-1011

Physiology and pathophysiology of respiratory mucosa of the nose and the paranasal sinuses

Review Article

Search Medline for

  • corresponding author Achim G. Beule - Department of Otorhinolaryngology, Head and Neck Surgery, University Greifswald, Germany

GMS Curr Top Otorhinolaryngol Head Neck Surg 2010;9:Doc07

doi: 10.3205/cto000071, urn:nbn:de:0183-cto0000719

Published: April 27, 2011

© 2011 Beule.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by-nc-nd/3.0/deed.en). You are free: to Share – to copy, distribute and transmit the work, provided the original author and source are credited.


Abstract

In this review, anatomy and physiology of the respiratory mucosa of nose and paranasal sinuses are summarized under the aspect of its clinical significance. Basics of endonasal cleaning including mucociliary clearance and nasal reflexes, as well as defence mechanisms are explained. Physiological wound healing, aspects of endonasal topical medical therapy and typical diagnostic procedures to evaluate the respiratory functions are presented. Finally, the pathophysiologies of different subtypes of non-allergic rhinitis are outlined together with treatment recommendations.

Keywords: physiology, pathophysiology, mucociliary clearance, non-allergic rhinitis, nasal cycle


1 Introduction

Respiratory mucosa of nose and paranasal sinuses are subject of daily routine of the ENT specialist. Restoration of the respiratory function is the aim of both conservative and surgical therapies. Therefore, the role of respiratory mucosa has been put into the centre of this paper. For ventilation aspects, it is referred to the article of Lindemann and Keck. From the perspective of clinical relevance, profound knowledge and innovative research results regarding anatomy (with embryology), physiological functions and their diagnostic tests are presented. Due to the amount of different pathophysiologies, only selected disorders are summarised with regard to their impaired function of the respiratory mucosa and therapy strategies are outlined. To facilitate coherent reading, references to the more surgically orientated papers are included.


2 Embryology

During the 6th week of embryological development, ectodermal thickened parts of the frontobasal prominence invaginate resulting in two symmetrical nasal placodes. Nasal pits divide the placodes (Fovea nasalis) into medial and lateral nasal processes [1]. This deepens, becoming the nasal sac – an ectodermally lined cavity, which is divided into two halves by the primary septum. Between weeks 6 and 7, a membrane separates the oral cavity from the primitive choanae. This resolves due to immigration of mesenchymal cells. The mesenchymal septum (secondary septum) grows caudally, merging with the palate and enlarging the embryonic structures of the nose. In weeks 7 and 8, first cartilaginous cells are identified, including parts of the turbinates. Starting in the 3rd month, mucosal lining migrates into the lateral nasal wall priming the development of the paranasal sinuses. By cell differentiation and maturation of the respiratory mucosa, detection of glands is possible after the 4th month of intrauterine development. During the 5th month of fetal development, ossification of the nose starts, including the previously cartilaginous areas of the turbinates. Influenced by various transcriptional factors, including Sox [2], cells differentiate further into the various cell types included in respiratory mucosa. For ciliogenesis and their pathophysiologies, several influencing factors play their part to a certain extent [3]. In animal models, the regulation of ciliogenesis is proven for FoxJ1, TTC25-GFP, Mig12-GFP, OFD 1, Ftm and Talpid3 [4], [5], [6]. Deletion of Talpid3, which works via the hedgehog signaling pathway, will inhibit the development of cilia completely [6], while disorders of the to hedgehog signaling pathway alone result in abnormal or degenerated cilia. A specific disruption of Talpid3 has been suggested as animal model for primary ciliary dyskinesia with polycystic kidneys [6]. For anchorage and orientation of cilia, further factors have been identified, such as FoxJ1 [4].


3 Anatomical and histological structures of the respiratory mucosa

Dorsally to the vestibulum nasi, which is lined with squamous epithelium, lies the nasal cavity. This is coated with 120 cm2 of pseudostratified columnar ciliated epithelium. The respiratory mucosa shows a thickness of 0.3–5 mm. Three (rarely four) turbinates protrude into the nasal cavity showing the thickest respiratory mucosa on their medial surface. In respiratory mucosa, several specific cell types can be identified. All cells are attached on the basal membrane. Basal cells lie on the membrane and show non contact with the epithelial surface. Their specific morphologic features are desmosomes for cell adhesion.

Columnar cells represent up to 70% of the epithelium and have 300–400 microvilli on their surface. The general principle of microvilli is the increase in surface area to retain moisture and to prevent drying of the surface [7]. Another 20–50% of epithelial cells are ciliated cell possessing 200–300 cilia on their surface, which are the morphological substrate of the mucociliary clearance. Cilia are 5 to 10 µm long and 250 nm thick and sheated in a plasma membrane. Within cilia, nine double tubules are arranged around two sheated central tubules. An inner sheath surrounds the central tubules. Outer pairs of microtubules (A and B-tubules) are connected to each other by nexin bridges and dynein arms and to the central pair by radial spokes. Bending of the cilia occurs as an ATP-consuming mechanism, called “sliding filament mechanism”. Existence of non-motile cilia is discussed in cells without central double tubules acting as sensory antenna. Cilia are covered by a 10–15 µm thick layer of mucus, filling also the spaces between cilia [8]. In respiratory mucosa, goblet cells and seromucous glands in the adjacent connective tissue are typically found. Goblet cells represent 5–15% of cells in the respiratory mucosa and produce secretions for the endonasal mucus together with the submucosal glands. Next to their form, existence of microvilli and a small opening, called stoma, are their characteristics. They differentiate out of non-ciliated cells [9]. In connective tissue, 20–80 anterior serous glands with 2–20 mm long ducts opening into small crypts are visible. Their importance is unclear [10]. About 90,000 seromucous glands possess ducts lined by cubical epithelium and are organised in two layers. Moreover, 20–50 intraepithelial mucous glands are detected around a central lumen. Their contribution to mucus production is regarded as small.

At the anterior septum, epithelium is underlined by a 1.5 mm thick tissue of convoluted vessels, called Locus Kieselbachii. In addition, venous plexus are seen in the area of the inferior turbinate and the nasal septum, acting as swell bodies during the nasal cycle.

Dimensional orientation for healthy respiratory mucosa in a biopsy indicate a subepithelial capillary network with a diameter of 0.025 mm, below a 1.6–10 µm thick basal membrane covered with a epithelium showing 40–100 µm height. Ciliated cells have a height of 15–20 µm and a width of 15 µm. Human cilia are about 6 µm long and 0.3 µm thick. Microvilli are 0.5–4 µm long and about 0.1 µm thick. These reference data are limited due to a high degree of variation depending on the endonasal localisation of cells.

As migrating cells, more T- than B-lymphocytes are detectable in subepithelial healthy tissue. The ratio of T-helper to regulatory T-cells (T-suppressor cells) is subepithelial 3:1 and in deeper layers 2:1 [11]. Natural killer cells are rare. 50% of IL-5-positive cells and 100% of IL-6 positive cells are mast cells [12].

3.1 Vascular supply

Branches of the maxillary artery are responsible for the arterial supply, including the sphenopalatine artery, the posterior lateral nasal artery and the infraorbital artery.

Venous blood drainage occurs along the facial vein after unification of the supratrochlear and supraorbital veins.

3.2 Lymphatic vessels

Superficial and deep lymphatic vessels (15–200 µm) can be demonstrated in respiratory mucosa, which lead in the middle nasal meatus to the natural ostium of the maxillary sinus. Their density decreases from top to bottom of the middle nasal meatus; their number is higher in the paranasal sinuses than the nose. Several connections are visible between lymphatic vessels and the vascular supply [13]. More lymphatic vessels begin at the nasal floor and the turbinates and converge in the medal inferior area of the turbinates. From this area, lymphatic vessels pass retropharyngeally and to the parapharyngeal space, reaching lymph nodes of both anatomic areas [14].

3.3 Innervation und regulation

Autonomic innervation takes place via the posterior ethmoidal nerve. Sympathetic innervation of the nasal mucosa is supplied via branches of the superior cervical ganglion passing along with the nerve of the pterygoid canal [15]. Hypothalamic stimulation will provoke a vasoconstriction [16]. Glands of the nasal mucosa, as well as the vessels, have a direct parasympathetic innervation leading to direct parasympathetic increase of nasal secretions via transsudation and exsudation [17].

Various cotransmitters were detected in nasal respiratory mucosa [18]. Parasympathetic neurons have mostly vasointestinal peptide (VIP) as cotransmitter to acetylcholine [19]. VIP is stimulating secretions (more serous than mucous) and vasodilatating at arterial and sinusoidal vessels [20]. Sympathetic neurons contain neuropeptide Y (NPY) as key cotransmitter to noradernaline and innervate predominantly arterioles and arteriovenous anastomoses. Liberation of NPY results in prolonged vasoconstriction, together with decongestion of venous sinusoidal vessels [21]. Substance P acts as Co-transmitter for Neurokinin A [22] and Calcitonin Gene-Related Peptide (CGRP) at arteries. The high number of cotransmitters is regarded as a possibility to adjust congestion of the nose delicately. Its disturbance may contribute to the aetiology of various, non-allergic rhinitis disorders [23].

An additional way of regulation is provided by endothelial nitric oxide synthase (NO-Synthetase; NOS; producing NO in the tissue) located in capillaries and arterioles near glands and nerve fibres around seromucous glands. As cotransmitter in parasympathetic nerve fibres, nitric oxide (NO) shows a neuromodulating and vasodilatating effect stimulating also seromucous glands [24]. As a consequence, both via nerve fibres and the endothelium, the degree of endonasal congestion can be altered [24].


4 Physiology of respiratory mucosa

Every day about 12,000 litre airflow is passing through the adult nose [25] and being hydrated (cf. article of Lindemann and Keck http://www.egms.de/en/journals/cto/2011-9/cto000072.shtml) and filtered. The nose is of extreme importance to protect the distal airways from deterious influence of gas, aerosol and pathogens. The nose and paranasal sinuses also act as area of voice resonance and produce nitric oxide (NO) for regulation of lower airways. Finally, the nose acts as chemosensoric organ responsible for smelling.

4.1 Cleaning function

The nasal passage filters 95% of particles with a diameter of more than 15 µm out of inspired air [25]. The cleaning function for pollen and dust of smaller dimensions is severely diminished but not abolished [26]. Liquids inhaled as aerosol will be eliminated from the upper airway if inspirited through the nose to 95% (mouth: 50%). Dosage of inhaled gas measured in the pulmonary alveoli will be diminished from 6–10% during mouth breathing to 0.9% during nasal breathing [27].

4.1.1 Sneeze reflex

Sneezing aims for elimination of particles from the nose. This reflex is most complexly coordinated affecting also the solitary nucleus [28]. Typically, sneezing will be provoked by foreign bodies in the anterior parts of the nose, which stimulate H1-receptors of trigeminal C-fibres [29]. After inspiration [30] ceases to allow glottic closure, sudden contraction of abdominal and breast muscles happens. After glottal opening, liquid drops or foreign bodies are tossed from the nose at velocities of 50 m/s [31]. This type of reflex can be triggered by light called photic sneeze reflex.

4.1.2 Nasolacrimal reflex

Nasolacrimal reflex results after chemical or mechanical stimulation of the nasal respiratory mucosa in increased lacrimal secretion. Afferent, C-fibre nociceptive neurons run together with the trigeminal nerve to the superior salivary nucleus, continuing via the geniculate ganglion, the large superficial petrosal nerve and through the pterygoid canal to the sphenopalatine ganglion. Cholinergic fibres reach together with the maxillary nerve the lacrimal gland [29]. Stimulation of one side leads also to a physiological, weaker reaction contralaterally [28] (for more information regarding reflexes: [32]).

4.2 Unspecific defense mechanism

Static and dynamic mechanisms (structure of the epithelium, configuration of endonasal airflow) and regulated physical and chemical mechanisms (structure and content of nasal mucus, mucociliary clearance, nasal cycle, plasma extravasation by NO [33]) assist in immune defense. Epithelial cells have a key position as a physical barrier and are the mainly responsible cells for maintaining the mucociliary transport. Respiratory mucosa of the nose is characterised by a high enzymatic activity, especially of the cytochrome P450 system [34]. NO, produced mainly by the mucosa of the sinuses and released at the surface, is discussed to have bactericidal effects in the airway [35].

4.2.1 Mucociliary clearance (MCC)

Mucociliary clearance is defined as cleaning of upper and lower airway by interaction of nasal mucus (about 200 g or 2 litre/day produced by the respiratory mucosa) [36] and ciliary beating. Number, structure and coordinated stroke of the cilia are as important as the biochemical, physical and chemical properties of the mucus. Nasal mucus has a weak, flexible, three-dimensional network formed from linear, hydrated mucin molecules. This is enhanced by disulfide bonds and secondary chemical connections between ions [37]. To prevent infections, mucus is slightly acid with a physiological pH-value of 5.5–6.5 and has a small capacity as chemical buffer. Via hydroxyl- (OH-) groups and oligosaccharide chains, negatively charged groups, the highly hydrated form of nasal mucus and the embedded network of linear and flexible glycoproteins, nasal mucus may form unspecific secondary connections, e.g. with pathogens or drugs. Viscoelasticity, adhesive and cohesive properties of nasal mucus are mainly determined by the glycoprotein compound [37]: They comprise a protein backbone covalently bound to oligosaccharide chains at a molecular mass of about 200 Kilodalton (kDa) and are responsible for the negative electric charge of the nasal mucus.

Optimal mucociliary clearance is achieved at 37° Celsius and 100% relative humidity (absolute: 44 mg/dm3). Nasal Mucus is about 10–15 µm thick [38] and has two layers: the lower, 6 µm thick liquid layer (also called: periciliary liquid) is covered by the more viscous gel phase. The gel phase is structured by embedded mucin. Height of the liquid layers has tremendous effect of the efficiency of the ciliary stroke [39]. Nasal mucus contains 90% water and glycoproteins as well as ions (cf. Table 1 [Tab. 1]). It is produced by submucosal, seromucous glands, goblet cells, transsudation of blood plasma, mucosal tissue fluid and tear fluid. Due to transsudate, most proteins detectable in serum may also be demonstrated in nasal secretions. In cases of local inflammation, the amount of transsudate and their respective proteins will increase.

Due to the coordinated, metachronous ciliary stroke (cf. Figure 1 [Fig. 1]), the mucus layer will be moved at a velocity of 2–25 mm/min [40]. In detail, control of the ciliary beat frequency is unknown. However, ciliary beat frequency will increase if cells are exposed to NO or a mechanical, calcium-mediated stimulus [41], whereas IL-13 will decrease the frequency [42]. In addition, intensive physical activity will decrease mucociliary clearance [43].

Particles bound to the mucus layer will be transported towards the pharynx passing the hiatus semilunaris. A second stream runs from the sphenoid sinus to the posterior ethmoid towards the choanae. Within a paranasal sinus, mucociliary clearance will always be orientated towards the primary natural opening [44], while accessory ostia are bypassed by the mucociliary clearance.

Next to water and electrolytes, immune globulin (Ig) G and A can be detected in high concentrations in nasal mucus. Secretoric Ig-A (up to 80% Ig-A1; among others against Coxsackie viruses and polio virus) is an obligatory ingredient and may provide up to 50% of the total protein of nasal secretions [45]. It is secreted into the tissue from plasma cells located near the basal membrane of the glands to bind and neutralise the antigen. Due to this effect, Ig-A is discussed as an important factor in pathogenic microbiological colonisation of respiratory mucosa.

Ig-G is synthesised in nasal submucosa and secreted after muscarinergic stimulation or exposition to histamine [46]. In physiology, Ig-M is not detectable, while Ig-E levels are below serum concentration. Moreover, lipids (e.g. surfactant [0.8%] and carboanhydrase [1%]) may be detected.

Application of sodium-chloride leads to increase of ciliary beat frequency [47] and thereby improvement of mucociliary clearance. According to a Cochrane review [48], salt concentration is of minor importance. On the other hand – depending on the pathophysiology – hypertonic solutions in chronic rhinosinusitis and isotonic salty solutions in acute sinusitis may be beneficial [49].

4.3 Humoral mechanisms

Unspecific substances for immune defense are localised in the epithelium and in nasal secretions. These include lysozyme (attacks peptidoglykans in the cell wall of gram positive bacteria), lactoferrin (inhibits bacterial growth) [50] and oligosaccharides (bind bacteria) [51]. Furthermore, neutrophil granulocytes produce proteases and hydrolases to destroy the cell membrane of bacteria and viruses. The 20 proteins of the complement system are detectable in blood and in tissue. They destroy foreign structures and marker pathogens for phagocytosis. The kinin-kallikrein system activates an inflammatory response after viral invasion by vasodilatation. Intracellular interferon impedes viral replication and inhibits ongoing viral infection.

4.4 Cellular mechanisms

Neutrophil granulocytes, monocytes and macrophages are cellular components of the host defense using phagocytosis in the subepithelial tissue of the nose and sinuses. Immigrated natural killer cells destroy infected cells. Under physiologic conditions, importance of cellular mechanisms is subordinated to unspecific components of immune defense. In case of pathogen invasion, multiple immunocompetent cells immigrate, and increase the importance of cellular defense. The specific immune system in nasal respiratory mucosa is part of the lymphatic system (mucosa-associated lymphatic tissue; MALT).

4.5 Nasal cycle

Nasal cycle is defined as spontaneous and reciprocal change of nasal congestion without change of the total nasal airflow. It was first described by R. Kayser in 1895 [52]. According to literature, nasal cycle can be detected in 70–90% of humans [53], [54].

The nasal cycle is regulated by the hypothalamus [55], with efferents running along the vidian nerve and showing an asymmetrical activity in controls [56]. Changes of volume in the erectile tissue of the septum, the inferior and middle turbinates and even in the paranasal sinuses can be demonstrated [57].

The working and resting phase can be easily distinguished. During the working phase the nasal passage is characterised by an increased hydraulic diameter, passage of nasal airflow and turbulence, while resistance is decreased. During resting phase resistance is raised, while hydraulic diameter, transnasal airflow and turbulence are diminished [53]. Mucociliary clearance is enhanced by factor 2.5 in a working phase in comparison to the contralateral resting phase [58]. Resting phase allows gathering of mucosal moisture [59], [60]. Mucosal cleaning and regeneration as well as improved host defense are discussed as advantageous during the evolution, eventually leading to the development of the nasal cycle [61].

If analysed more closely, contralateral working and resting phases with regular change ('classical type') can be differentiated from parallel working and resting phases ('in-concert type') and undefined phases ('irregular type'). The frequency of nasal cycle phase classified as classical type decreases with age [62]. There is a high amount of variation in the duration of phases reported (duration of phase: 1.6 h [54] 1–2.5 h [63]; 1.5–10 h [64]; 2.5 h [65], 2.9 h [66], 3–4 h [67]; 3–5 h [68]; 4.3 h [69]; 1–5–10 h [53]; 7 h [70]). A relationship to the individual duration of rapid eye movements (REM) in sleep has been demonstrated [71].

The change from one phase to another happens quite rapidly within a few minutes. This is helpful in distinguishing changes of nasal congestion caused by the nasal cycle from those due to change of body position [68], [72], [73], [74], or physical activity mediated though the adrenergic sympathetic system.

4.6 Microbiological colonisation

In the nose, a physiologic, multi-microbial colonisation exists. Specific articles are summarised in Table 2 [Tab. 2] and Table 3 [Tab. 3]. Normal microflora and pathogens are differentiated from each other by their ability to release a Th1/2-mediated immune reaction. Commensal bacteria may pass the epithelial barrier and will only evoke a minimal T-cell activation without effects on regulatory T-cells (CD25FoxP3) [75].

Paranasal sinuses are sterile under physiologic conditions [76]. In 50 cultures obtained during nasendoscopy of 25 healthy controls, 32 bacteria were detected such as Staphylococcus aureus, coagulase-negative Staphylococci, Corynebacteria and Propionibacterium acnes [77]. After sinus surgery, detection rate increased to 97% in patients with clinically normal endoscopic findings. Here, coagulase-negative Staphylococci (69%), Corynebacteria (25%), S. aureus (31%) and Pseudomonas aeruginosa (3%) can be identified [78]. Fungus is also part of normal microflora of nose and paranasal sinuses. While 20% of postnatal babies have a positive fungal culture in nasal secretions, this detection rate increased to 94% at the age of 4 months [79].

Methods of choice to detect colonisation are endoscopically guided aspirate specimens or sinus punctures. In comparison to the latter, accepted as gold standard, cultures from the middle nasal meatus showed a sensitivity of 80% and specificity of 100% with regard to their detection rate (positive predictive value: 100%; negative predictive value 78.6%). If analysed prospectively, results will match in 88.5%, suggesting that endoscopically guided aspiration will be sufficient while resulting in minor burden for the patient [80].

In 70% of people working in a hospital, S. aureus can be detected in a nasal swab. To facilitate defense of the respiratory mucosa, exposure to commensal bacteria has been evaluated. Medical doctors may terminate their colonisation with methicillin-sensitive S. aureus, if inoculated with Corynebacteria spp., also decreasing the risk of methicillin-resistant S. aureus infection [81].

4.7 Regeneration and wound healing

The detailed sequence of wound healing of the nasal respiratory mucosa is delicately regulated [82]. In the following, relevant procedures during course of wound healing with evident clinical aspects will be presented.

Aim of wound healing is structural and functional repair of the tissue defect. Fundamental differences between partial and full-thickness defects (affecting all layers) of respiratory mucosa have been reported [83]: Preservation of the basal membrane results in rapid restoration of normal epithelial height [84], while its destruction will lead to a repair process taking weeks to months [84]. Deeper wounds will also result in scar formation, with its extent being quite variable [85]. Duration of wound healing of respiratory mucosa is estimated as 6 months, and sometimes even longer [86]. Physiological wound repair happens in different stages, which may temporarily overlap and macroscopic appearances show fluent transition. The stages are regulated by a variety of cytokines and growth factors.

4.7.1 Temporal sequence of wound healing

Based on the endoscopic aspect of the wound area, the stages of blood crusting (until day 10), obstructive lymphoedema (until day 30), and mesenchymal growth (up to three months) can be distinguished with subsequent scar formation [87].

Coagulation and haemostasis

The trauma leads to vessel injury and release of blood. Thrombocytes aggregate and release PDGF (Platelet-derived Growth Factor). In parallel, vasoconstriction occurs to diminish the injury of the vessel wall. Platelets activate the coagulation cascade to close the defect with a thrombus [88]. Subsequently, the thrombus dehydrates and resolves into crusts. Neutrophil granulocytes immigrate and release pro-inflammatory cytokines. This leads to vasodilatation and increase of vessel permeability with oedema formation. Platelets integrated into the fibrin net chemotactically attract macrophages [89], fibroblasts and endothelial cells to the defect [90]. Due to repeated microtraumata, this coagulation cascade may be activated on several occasions during the first one to two weeks [91].

Inflammatory phase

Next to neutrophil granulocytes, the thrombus is dissolved by immigrated monocytes and macrophages. Neutrophils release more proliferative cytokines (e.g. Transforming Growth Factor (TGF)-α, TGF β 1, IL-1). Cell migration shows a peak for neutrophil granulocytes after 2 days, for monocytes and macrophages after 4–5 days, and for lymphocytes after six days. Monocytes differentiate to tissue macrophages and eliminate microorganisms and debris. They also regulate enzymatic tissue disaggregation. Macrophages control the transition from the inflammatory to the proliferative phase by production of growth factors, such as tumor necrosis factor (TNF)-α [90].

Proliferative phase

The defect will be filled with granulation tissue. This happens by proliferation of connective tissue, with synthesis, degradation, and alteration of collagen, angiogenesis, production of glycosaminoglycans and epithelial coverage. In addition, osteoneogenesis can be observed [86], [92], [93]. This phase starts about three days after the initial trauma and lasts for at least 2–3 weeks. Fibroblasts produce among other substrates collagens, a main component of the extracellular matrix (ECM) in wounds detectable since day three after injury [94]. Collagen bundles are built by rearrangement of collagen fibril orientation. Due to better organisation of collagen, resistance of the wound against shear is increased. One year after injury, remodelling processes of collagen are still detectable. Due to the close molecular connections of fibroblasts, ECM and collagen synthesis, fibroblasts are regarded as main source of adhesions.

As early as 24 hours after epithelial injury [95], epithelial cells migrate concentrically at a velocity of 20 μm/h [92] into the defect [92], [96]. Until epithelial closure, depending on the defect’s size more than two weeks will pass [97]. In parallel, changes in ECM proteins, such as matrix metalloproteinases (MMPs) and release of more cytokines take place. MMP-9 (see below) is of special interest. Re-innervation of mucosal glands of the paranasal sinuses is detectable after 4–8 weeks [98].

Phase of remodelling

During this phase, granulation tissue is replaced by scar formation. “Airway remodelling” is frequently used as general term for continuous adaptation of mucosa during the course of a chronic inflammation (e.g. bronchial asthma or chronic sinusitis). In contrast, remodelling in wound healing is defined as a process of collagen degradation, conversion and synthesis and has a limited duration. Type-3 collagen is replaced by the more firm type-1 collagen. This process starts after 48–72 hours and is detectable by increased rate of mitosis and hyperplasia at the wound edges with increased activity of MMPs [95].

Continuous synthesis of firm collagen, its stabilization with cross-linking and the aggregation to thicker bundles as well as its coordinated orientation, increases the strength of the reconstruction.

Three months [99] to two years [100] takes “wound healing”, if the duration is measured as time interval from injury (e.g. during surgery) to accomplishment of the optimal and final functional result. Especially in adults [101], the regenerated tissue shows less serous glands [97] with increased number of excretory ducts [102]. Subepithelial stroma of the regenerated tissue is thicker and the number of vessels is increased [103]. Only slightly more than half of all patients show a normal mucociliary clearance 18 months after sinus surgery if evaluated with nuclear medicine techniques [104].

4.7.2 Relevance of matrix Metalloproteinase-9 (MMP-9) as indicator of wound healing

MMP-9 belongs to the family of matrix metalloproteinases, which may activate each other in-vitro synergistically. This matrix metalloproteinases is a type IV collagenase and is also called Gelatinase B [105]. After a stimulus, e.g. an injury, a 92 kDa proenzyme is released which can be activated by removal of the propeptide resulting in the active 82 kDa enzyme [106]. MMP-9 is detectable only during the first days after injury. Generally, increased levels of MMP-9 determined in wound secretion are associated with problems in reepitheliasation [107], [108].

With regard to sinus surgery, MMP-9 was presented as predictive marker for healing quality. High concentrations of MMP-9 go together with worse healing quality and distinct oedema [109].

4.7.3 Risk factors for deteriorated wound healing

Wound healing is above all, complicated by local disturbances like hypoxia [110] or low-nutrition. Systemic disorders (e.g. diabetes mellitus (see below), tumours, immune deficiencies [111], cystic fibrosis, primary ciliary dyskinesia, aspirin intolerance, bronchial asthma, gastroesophageal reflux, autoimmune diseases), cytotoxic or immunosuppressive medication, allergic disposition or active or passive smoking [112] may impede wound healing.

This may result in exceeding granulation tissue, which favours adhesion formation.

Eosinophilic histology was identified as risk factor for frontal ostium restenosis after type III drainage according to Draf [113]. Unfortunately, practical benefit of this finding is limited due to the high frequency of eosinophilic inflammation in chronic sinusitis.

4.8 Rhinologic functional diagnostics

4.8.1 Anterior rhinoscopy

During anterior rhinoscopy, blood vessel pattern and quality of nasal mucosa with colour and properties of secretion are evaluated. To facilitate better estimation of nasal congestion, inspection before and after decongestion is mandatory. Endonasal resistance is assessed based on configuration and extent of ventilated nasal areas. Bachmann’s test (introduction of a cotton-tipped swab in the upper nasal valve) as well as Cottle’s test, enlarge the area of the nasal isthmus. They are rated as too unreliable and observer dependent to diagnose isthmus stenosis. Visual judgement of the inflow areas has been recommended [114], but is also observer dependent.

4.8.2 Endoscopy und videoendoscopy

During standardised nasendoscopy, after inspection of the inferior nasal meatus, the middle nasal meatus with middle turbinate, the anterior wall of the sphenoid sinus, sphenoethmoidal recessus and the superior nasal meatus with olfactory groove are inspected. As videoendoscopy [115] digital documentation is possible for analysis by independent observers [116].

4.8.3 Clinical imaging

The clinical importance of imaging modalities such as computer tomography (CT) and magnetic resonance imaging (MRI) lies in the possibility to verify mucosal swelling of the sinuses [117]. Extent and type of sinusitis [118], presence of invasive processes such as neoplasia and invasive fungal sinusitis (MRI: [119]) or non-invasive fungal sinusitis [120] can be evaluated. A positive aspect is its high reliability up to a time interval of more than 120 days [121]].

The relative importance of sonography is limited to follow-up of acute sinusitis and special problems during pregnancy and childhood. Sensitivity and specificity of ultrasound of the sinuses is not sufficient for detection of mucosal swelling in clinical routine [122]. For the same reason, employment of conventional x-rays in chronic sinusitis is discouraged [117].

Innovative studies report a high resolution of optical coherence tomography with regard to nasal mucosa [123], suggesting this technique as most interesting alternative in future.

4.8.4 Diagnostic tests of mucociliary clearance in-vivo
Saccharin test

For this most commonly employed global assessment of mucociliary clearance, saccharin particles (usually a few crystals) are placed on the head of the inferior turbinate. Time intervall until the patient reports a sweet gustatoric sensation is measured. This technically easy and inexpensive test is lacking any side-effect. It may be employed in children, but is dependent on a normal sense of taste. A time interval from placement of saccharin to taste sensation of ≤30 min. is regarded as physiological; a time interval exceeding one hour as pathological [124]. Repeated measurements may vary by about 6 min. [125]. The long duration of this test is commonly regarded as disadvantagous. As modification, application of a solution has been suggested, shortening the transport time to about 10 min [126].

Nuclear medical tests

Using technetium 99, mucociliary clearance can be evaluated by means of nuclear medicine [127]. This sensitive method comes along with increased technical effort [128] and radioactive contamination, but is due to its reliability [129] still employed for special questions [130], [131].

Particle tests

Titanium dioxide (TiO2) is a 500 µm TiO2 particle, which can be placed on the mucosa to measure its velocity (in mm/min) in vivo. Standardised reference values have been reported for the nasal floor, employment is due to the small weight supposed to be possible on the whole area of respiratory mucosa. The short duration of this test is favourable [132]. In analogy to earlier studies using resin [128], [133], this test is believed to particularly evaluate the gel phase of the nasal mucus. General advantage of tests employing particles are: independence of the sense of taste, short duration of testing, and in some cases less technical effort and less costs.

4.8.5 In-vitro tests for parts of mucociliary clearance

Ciliary beat frequency (normal range: 7–12 Hz) can be determined after removal of ciliated cells (e.g. by nasal curettage or brushing [134]) using a phase-contrast microscope. An average value for at least four cells should be calculated when this frequently employed technique is reported [47], [135]. To increase reliability and allow for quicker analysis, a specially developed digital photometry has been reported [136].

If a hereditary disorder of cilia function is suspected, cell culture of ciliated cells has been recommended [137]. This time-consuming technique is only available in very few centres, but shows nearly ideal sensitivity and specificity [138].

Ingredients of nasal mucus can be evaluated in-vitro using ELISA [139], [140]. As well, the determination of certain cytokines is possible in nasal secretions [141]. However these techniques are not employed in clinical daily routine.

Mucosal biopsy may be subjected to conventional histology to aid in differential diagnoses and to assess inflammatory activity (based on the number of inflammatory cells). Furthermore, biopsies may facilitate more precise classification of chronic rhinitis (eosinophil/neutrophil inflammation). In addition, scanning electron microscopy can be used to visualise the ultrastructural surface of respiratory mucosa and nasal mucus. It is not appropriate for specific detection of biofilms [142].

Using transmission electron microscopy (TEM), changes within the cilia going together with altered dynein structures are reliably detected. Gathering specimens is reliably accomplished by nasal brushing [134]. Costs and expenditure of time are unfavourable. Furthermore, TEM is highly specific in detection of biofilms but at the cost of immense work load [142].

Confocal laser scanning microscopy (CLSM) [51] is a less costly technique to detect biofilms with superior sensitivity and specificity. For CLSM, the specimen is incubated with fluorescent cell markers (e.g. Soty 9 and Propidium Iodide as differential staining for living and dead cells). Afterwards, it is inspected three dimensionally using a laser beam. Different wavelengths of the laser stimulate the cell markers, resulting in coloured light emission depending on the functional state of the cells. Analysing the surface of larger specimens can be accomplished in short time and with reasonable workload [51], [143].

4.8.6 Rhinomanometry (RMM)

Rhinomanometry (RMM) allows objective measurement of nasal airflow synchronous to nasal breathing based on the transnasal pressure difference. Airtight placement of a pressure hose at the nostril (e.g. by sticky tape) elongates the pressure column to the choanae and serves as reference value for the examined contralateral side [144]. Deformities of the nasal inflow area should be avoided, due to their associated measurement error. If measured in accordance to the recommended setting [144], [145] before and after decongestion, nasal obstruction can be rated at a pressure difference of 150 Pa as flow velocity or resistance (R=ratio of pressure difference and flow velocity; cf. Table 4 [Tab. 4]) [114]. One-sided reference values are discussed controversially in international literature [146]. As a consequence, total flow of >900 ml/sec at 150 Pa were reported as lower limit of normal [146]. In contrast to classical “anterior RMM“, “posterior RMM“ allows the measurement even in patients with septal perforation after placement of a pressure hose in the mouth, but is prone for measurement errors caused by positioning of the soft palate in relation to the back wall of the pharynx.

The relevance of RMM in rhinology has been documented [144], but is limited due to missing age-stratified reference values [147]. Measurement error may reach 15% [148] and normal rhinomanometry can be obtained in 25% of patients with obvious pathological findings [149].

4.8.7 Rhinoresistometry (RRM)

Employing fluid dynamics, rhinoresistometry is a refinement of rhinomanometry [144], [150]. Measurement technique and protocol are identical to anterior RMM. To sum up, nasal resistance can be classified at a flow velocity of 250 cm³/s according to the reference values indicated in Table 5 [Tab. 5]. In addition to the information availably by RMM, RRM aims at objectively detecting the aetiology of nasal obstruction. The value of RRM is still discussed controversially due to missing prospective studies.

As parameter for the inner width of the nasal cavity and endonasal friction, hydraulic diameter is indicated. After decongestion, the hydraulic diameter in a normal nose should measure ≥6 mm. This parameter can also be used for objectively evaluating changes in endonasal congestion. [151]. The comparison of a measured and a calculated graph allows objective assessment of inspiratory nasal wing collapse. Deviation of both graphs at flow velocities ≥500 cm³/s is regarded as pathological [152]. Endonasal turbulence is presented as a graph in relation to flow velocities and calculated as friction coefficient λ (normal after decongestion λ > 0.025) for both nasal sides separately.

Measurement process of RRM is reliable [152] and is aiming to objectively differentiate certain aetiologies of nasal obstruction (mucosal changes, cartilaginous/bony findings, pathological inspiratory nasal valve collapse, changes in endonasal turbulence). Results of rhinomanometry and rhinoresistometry correlate well [151].

4.8.8 Acoustic rhinometry (AR)

For acoustic rhinometry, sound is applied endonasally and based on its reflection the dimensions of the nasal cavity are calculated. Measurements should be performed under standardised circumstances [144] to secure its high reliability (measurement error: 2–4% [148], [153]). Volume and diameter of two typical locations, the internal isthmus (I-notch [mean cross-sectional area 1; MCA1] and the head of the inferior turbinate (Conchal notch [mean cross-sectional area 2; MCA2]) can be calculated [144], to geometrically determine extent and localisation of a stenosis. The nasal diffusor can be measured by the gain in diameter from I-notch to the the spacious area posterior to the C-notch [53]. Measurements distal to a severe stenosis are unreliable [154]. Reference values have been published for several groups of patients (Caucasian adults [155], [156], [157] and children [156], [158], [159]; Asian adults [155], [160] and children [161]).

4.8.9 Long-term rhinoflowmetry (LRM)

Oxygen cannulas are placed in the nostril near the floor of the nasal vestibule. Nasal airflow can be recorded for 24 hours via a portable device. Graphical presentation facilitates detailed insight into changes of nasal congestion over time including definition of the type of nasal cycle and measurement of the duration of working/resting phase [54], [64]. The aim is to objectively assess pathological mucosal swelling under conditions of daily living [64].

4.8.10 Measurement of nasal expiratory nitric oxide (NO)

Human cells produce nitric oxide (NO) from arginine by the enzymes NO-synthases [162]. Nitric oxide is produced in the nasal cavity [162], but first and foremost in the paranasal sinuses [162]. It has anti-bacterial, as well as anti-viral properties and regulates mucociliary clearance via control of nasal secretions [163], [164]. Measurement of nasal NO [in parts per million; ppm] is influenced by lower airways [165], local and systemic diseases and their therapy [166], compliance [167], nasal cycle [168], anatomy [169] and technical aspects [170]. Valid measurement protocols have been reported [171] for compliant patients [169] and infants [172]. Reliable reference values do not exist due to high interindividual variation [173].

Using nasal NO measurement, patients with chronic rhinosinusitis with polyps can be distinguished from those without polyps or healthy controls [174] and treatment effects can be monitored [175]. Most significant differences to a normal population can be obtained in patients with primary ciliary dyskinesia or cystic fibrosis [176], suggesting nasal NO as non-invasive screening tool for these diseases [177].

4.9 Special aspects to functions of respiratory mucosa

4.9.1 Particularities of children

Children in Western industrial nations suffer from 6–8 viral upper airway infections per year [178] with 5–13% developing bacterial sinusitis [179], [180]. This frequency, together with the common problems of adenoid hyperplasia favouring nasal blockage and susceptibility to local inflammation, complicate effective diagnosis of chronic rhinitis. No epidemiological data is available for non-allergic chronic rhinitis in children. In newborn and infants, a connatal hypothyroidism, a (usually autoimmune) thyreoiditis [181], or an infection with Chlamydia (usually acquired during birth) may results in symptoms of nasal obstruction [182]. Children, whose mother is heroin addicted, show nasal complaints often due to structural changes of cilia [183]. In children below the age of 12, a sweat test is mandatory to exclude cystic fibrosis in case endonasal polypoid tissue [184]. If simultaneously chronic bronchitis, sinusitis and otitis are present, a primary ciliary dyskinesia (immotile ciliary syndrome) or Kartagener syndrome should be taken into consideration [184].

In infants suffering from recurrent infections, (humoral) immunodeficiencies should be taken into account. With an incidence of 1:700 Ig-A deficiencies due to a B-cell-defect are most often diagnosed [185]. Besides, there exists evidence that in children gastroesophageal reflux is an confounding factor for recurrent upper airway infections [186]. During puberty, sexually transmitted diseases [187] (e.g. syphilis [188]) and hormonal fluctuations may result in rhinologic functional disturbances [189].

For diagnostics in and follow-up of paediatric patients, acoustic rhinometry is recommended. Growth-induced increase of airflow or decrease of resistance measured with rhinomanometry can be estimated by comparison with body weight development. Using this technique, pathophysiological effects can be distinguished from growth-induced ones.

4.9.2 Particularities of geriatric patients

During aging, nasal mucosa shows signs of atrophy with decrease of goblet cells and thickening of the basal membrane [190]. This repair is mirrored in increased expression of caspase-3, an apoptotic marker [191]. In parallel, elasticity of nasal mucosa decreases, by part due to a decreasing level of oestrogen. Accordingly, in postmenopausal women, substitution using a nasal spray has been recommended [192], [193].

Results on the effect of age on mucociliary clearance are contradicting: While some studies showed no effect on mucociliary clearance [194], [195], others were able to demonstrate a decrease of ciliary beat frequency with increase of saccharin transit time as equivalent of deterioration of mucociliary clearance in old age [196]. Viscosity of nasal secretion increases and causes post nasal drip with consecutive repeated clearing of the throat.

As structural change of respiratory mucosa in old age, single instead of double tubules were detected using transmission electron microscopy. In rhinoscopy, a wide nasal cavity with quite dry nasal mucosa lining is typically seen. Increased turbulence despite objectively free nasal passages may result in a feeling of nasal obstruction (paradoxical nasal blockage). Accordingly, endonasal volume and areas of MCA1 and MCA2 are enlarged [197]. Prevalence of endonasal polyps increases with age [198]. Also, primary atrophic rhinitis is associated with aging [199]. Due to the frequent medication with multiple drugs, rhinitis medicamentosa has a high relevance as differential diagnosis [199], [200]. As therapy, topical steroids (sometimes under addition of propylene glycole), expectorant drugs to stimulate exocrine glands and nasal douching with saline are recommended [184].

4.9.3 Influence of diabetes mellitus on the respiratory function

In diabetics, upper airway infections do not occur more frequently than in healthy subjects [201]. There exists controversial discussion whether the diabetic nose is more frequently colonised by pathogens, such as S. aureus [202].

Much clearer is the situation in diabetics undergoing dialysis: These have in 53.4% (non-diabetics: 18.6%) a nasal colonisation with methicillin-sensible and in 19% with methicillin-resistant S. aureus (MRSA; non-diabetics: 6.0%). In addition, if a patient has a central-venous catheter, his risk of endonasal colonisation of S. aureus is 35 times increased in comparison to patients with arterio-venous fistula [203].

A polymorphism of vitamine D receptor in diabetics has also been associated with colonisation of S. aureus [204]. Moreover, concentration of the bactericidal nasal NO is decreased in insulin dependent diabetes mellitus [166]. In parallel, diabetics have an increased resistance measured by RMM. This increase is more pronounced before than after decongestion, indicating a chronic state of increased congestion.

Chronic congestion is also favoured by mucociliary clearance, the latter being reduced in diabetics [205], [206] due to alcalised pH-values of nasal secretions [207]. Insulin-dependent diabetes mellitus and diabetes lasting for more than ten years result in further deterioration of mucociliary clearance [207].

Diabetics have more frequent clinical signs of a dry nose [208] and are at a higher risk to suffer from non-invasive [209] or invasive [210] fungal sinusitis [211]. A topical moistening by employment of nasal douching and oil has been recommended as therapy or prophylaxis, respectively [208].

4.9.4 Nasal respiratory mucosa as area of drug application

If topical drug application is discussed, it is of special importance to distinguish pharmacokinetics and pharmacodynamics. The pharmacokinetic science describes effects of the human body on the drug, while pharmacodynamics aims at studying the effect of the drug on the human body. In the following, special pharmacokinetic aspects are elaborated on.

Distribution of drugs in the nose and the paranasal sinuses

Use of nasal spray transports only a small part of the dosage to the paranasal sinuses, while the biggest part is deposited in the inferior meatus and on the head of the inferior turbinate [212]. For optimal application with regard to the frontal sinus, Vertex-to floor position has been recommended [213]. Effectivity of dexamethasone nasal drops in this position has been demonstrated with prevention of frontal ostium restenosis [214]. Unfortunately, compliance to employ this position is limited by existing co-morbidities (e.g. diseases of the cervical spine). In contrast, inhalation devices using alternating pressure waves reach the paranasal sinuses reliably within three minutes near the ostium [215].

Another application form is nasal douching being well-tolerated by 80% of patients. To rinse a paranasal sinus effectively, an ostium with a diameter of 3.95 mm is necessary [216]. Postoperatively, nasal douche is a reliable mode of application [215], [217]. To rinse the frontal sinus, a maximal forward bending (“Vertex to floor-position“) with maximal volume of the employed squeeze bottle is recommended. Generally, nasal douching should be performed with high volume and low pressure [218].

An additional innovative method of topical therapy is the employment of drug-releasing stents [219]. Releasing dexamethasone, granulation formation [96] and osteoneogenese [93] are decreased significantly, without impeding reepithelialisation. While stents may provoke a local foreign body reaction with biofilm formation, they can reliably provide topical application to inaccessible areas of the upper airways.

Aspects regarding resorption and absorption of medication in nose and sinuses

Mucociliary clearance limits the dosage of topically applied drugs (cf. Figure 2 [Fig. 2]). Particle transport of MCC restricts application duration of topical medication to an estimated 20–30 min. [220], [221]. In addition, mechanical activity of cilia destroys drug molecules [37], [220].

Moreover, nasal mucus is a barrier to diffusion [222], [223]. Molecular size and electric charge are decisive for the diffusion capacity of a drug [222], [224]. With regards to this background, microemulsions were developed to increase resorption into nasal mucus [225].

Mucins of the gel phase conjugate drugs and change their absorption properties. This may be taken advantage of by supply of lipophil pro-forms of the active substances [226]. Alternatively, medication influences mucin secretion [227] and it viscosity [228]. For drugs interacting with glycoproteins of nasal mucus [229], this property may limit drug permeation [230]. Especially charged particles with the capacity to build hydrogen bonds and lipophil particles are affected [231].

Nasal mucosa is well permeable for hydrophilic substances and lipophilic ones below a molecular mass of 1000 Da [232], [233], because the nasal membrane has pores with dimensions of 3.9 to 8.4 Å. Transcellular transport is possible for lipophil substances, while charged hydrophilic drugs are mainly transported paracellularly [234]. Paracellular transport happens via tight junctions (TJ) or the zonula occludens (ZO) [234]. Connections between goblet cells or goblet cells and ciliated cells are rated as loose [234], [235]. Besides, they show a high variation of permeability [236]. Many substances are inactivated by the high enzymatic activity (local ”first-pass effect“) of nasal mucosa during the process of absorption [234]. Carrier systems (e.g. chitosan [135], [237]), are employed to enhance nasal bioavailability. In parallel, a strategy has recently been developed to intraoperatively employ resorbable, medication-releasing application systems to increase the applicated dosage. Possiblities of technical refinement of modified systems are limited, as mucosal function should not be harmed.

Nasal drugs have to be well tolerated, exert beneficial effect on ciliary function and if applicable should show good dispensing properties (if applicated as nasal spray) and/or a high degree of adhesion (for better resorption).


5 Chronic rhinitis – selected disorders

In the following section, selected typical forms of chronic rhinitis will be discussed. Allergic rhinitis is differentiated from non-allergic forms. 57% of patients with chronic rhinitis suffer from a non-allergic form [238]. Classification of non-allergic rhinitis is still controversially discussed and quite complex [184]. Some non-allergic, non-infectious forms of rhinitis with a pathophysiology closely connected to nasal respiratory mucosa are summarised here. With regards to different non-allergic forms, it is referred to two elaborative and recent publications [184], [239].

5.1 Rhinitis medicamentosa

Rhinitis medicamentosa, also called drug-induced rhinitis, is defined as pathological congestion or the nasal mucosa with nasal blockage, nasal drainage (anterior rhinorrhea and post nasal drip) and optional sneezing. It occurs as a direct consequence of drug intake, with the classical disorder developing after prolonged use of decongesting nasal drops [240]. Due to its high prevalence, rhinitis medicamentosa is of special significance among the various subtypes of non-allergic rhinitis. Incidence is estimated as 6–9% in patients with nasal obstruction [241], [242] and 1–7% [241] of unselected patients in an ENT office.

Reported drugs that may possibly cause rhinitis medicamentosa (cf. Table 6 [Tab. 6]) belong to various groups of drugs.

In 1931, this disorder was first reported by Fox [243]. As pathophysiology, respiratory mucosa becomes accustomed to the drug, with the latter resulting in minor [244] and shorter [245] drug-induced effect (hyporeactivity). It remains unknown whether this is a consequence of tissue hypoxia with reactive hyperaemia due to vasodilatation [246], a negative presynaptic feedback mechanism at α-2-receptors [247], and/or change of vasomotoric tonus by increased activity of cholinesterase [248], [249]. Based on structural deformities, increased permeability [250] and secretion of glands [251] were reported.

The risk to suffer from rhinitis medicamentosa is gender independent and it most commonly affects middle aged adults. The postulated increased susceptibility of patients with chronic inflammation of nasal respiratory mucosa [252] failed to be proven [253]. The risk to acquire this disease dramatically increases if topical sympathomimetics are applied three times a day for more than 10 days [245], [253], [254] or night-time application exceeds 4 weeks [245], [255].

Decongestants used to have benzalkonium chloride as a preservative. This quartery ammonium compound is capable of destroying cell walls of microorganisms. In-vitro, toxic effects on cilia [252], deterioration of granulocyte chemotaxis and phagocytosis [256] and of neutrophil defense [257] have been reported. Because benzalkonium chloride aids development of drug-induced rhinitis [253], [258], aggravates its symptoms [241], [259] and may provoke rhinitis medicamentosa if used alone [260], only a few formulations in Europe are still available containing benzalkonium chloride [240].

Structural changes in rhinitis medicamentosa include damage and loss of cilia [261], metaplasia [261] and epithelial oedema [262], tears in the basal membrane [261] and openings in subepithelial endothel [250], [261] together with vasodilatation [241]. Moreover, hyperplasia of goblet cells [251], [262] and infiltration of inflammatory cells [263] were demonstrated probably as equivalent of reparative changes with increased expression of epithelial growth factor [263]. As complication, septal perforation may occur [264]. If not adequately treated, this disease leads to secondary atrophic rhinitis up to frank clinical picture of ozaena [265].

Careful history taking with questions regarding last topical and complete current medication is the key element to diagnose this disorder. For therapy planning, detailed history of previously unsuccessful therapeutic trials is helpful. Clinical examination should be performed in the early morning to provide the typical clinical picture with congested nasal mucosa in nasendoscopy after last topical medication on the previous evening [241]. Nasendoscopy is also beneficial to detect structural changes, which eventually may have led to the use of nasal drops. For apparative diagnostics, anterior rhinomanometry, acoustic rhinometry and measurement of mucociliary function have been recommended [266].

For successful therapy, identification and elimination of the causative drug is decisive. The patient has to be informed of the underlying pathophysiology to improve compliance (“counselling”) [267]. Sudden stop of intake leads to a rebound-swelling. In severe cases, separate cessation of application on both nasal sides is mandatory. Topical application of steroids (in severe cases starting before cessation of the causative agent) is able to effectively diminish rebound-swelling [268], [269]. Onset of decongestive effect is expected 4–7 days after start of steroid use [262]. A surgical procedure like a turbinoplasty may ease withdrawal for the patient (reported success rate of 88% for laser turbinoplasty [270]), but will not cure the disease. After repeated turbinoplasty due to prolonged abuse of decongestants, a surgically caused nasal dryness may develop. Therefore a three month time interval between cessation of abuse and surgical correction has been recommended [241].

The necessity to diagnose and possibly treat the underlying pathology, initially leading to abuse of decongestants, is not to be underestimated [241]. The underlying disorder for topical medication resulting in rhinitis medicamentosa is in 30% of cases an episode of upper respiratory tract infection, in 22% a non-allergic, non-infectious rhinitis, in 16% allergic rhinitis, in 30% structural deformities (such as septal deviation, endonasal polyps, posttraumatic pathologies), and in 3% hormonal changes [271].

For prevention, avoidance of benzalkonium chloride is mandatory [259]. Patients have to be aware of their persistent susceptibility to develop relapse of rhinitis medicamentosa even one year after onset of successful treatment [272]. Furthermore, control visits every three months have been recommended. During infection, short-term use (up to 3 days) of nasal decongestants at lower concentrations may present a preventive strategy.

5.2 Atrophic rhinitis, anterior rhinitis sicca and ozaena

Atrophic rhinopathy is characterised as a chronic inflammation of unclear aetiology, leading to atrophy and decay of the entire endonasal mucosa and the underlying bone.

A primary form, foremost occurring in developing countries and showing spontaneous development, is distinguished from a secondary form after rhinologic surgical procedures (especially: turbinoplasty), trauma, infection, granulomatous inflammation, or radiation [273]. For the primary form, six times more females than males are affected in developing countries. Patients live 1.5 times more often in rural areas than in urban ones [274]. A genetic component has been reported [275] with incidence decreasing in recent years [276]. A vetenerian type of primary atrophic rhinitis is often seen in pigs, resulting in development of multiple veterinary vaccines against the causative pasteurella multocida toxin [277].

Endonasal crust formation (usually on the middle turbinate), epithelial metaplasia with loss of cilia and superinfection of nasal mucosa (with detection of Klebsiella ocaena in 65–100% [274], [278]) are visible. In addition, vasodilatation [274] and changes in bone formation including osteitis of the turbinate bony may occur [278]. As a consequence, MCA1 is enlarged [279]. A classical smell develops and polyp formation, chronic dacryocystitis [280], scar-induced stenosis of the choanae [281], or pharyngitis sicca [282] are possible. The endpoint of atrophic rhinitis is called ozaena ('stink-polyp'). The patient does not realise his odour, but complaints about anosmia [283].

Histologically, increased activity of caspace-3 in the epithelium and subepithelial glands has been demonstrated as indicator for apoptosis [284]. Absolute quantity and chemical quality of phospholipids (with surfactant) is decreased [285]. Therefore, a lack of surfactant has been suspected as aetiopathogenesis of atrophic rhinitis [285]. By application of angiogenetic inhibitors, a similar disorder can be provoked in a recently reported animal model [286].

Anterior rhinitis sicca is not easily distinguished from atrophic rhinitis, but affects only the mucosa of the anterior septum with elsewhere in the nasal cavity presenting normally configurated mucosa. In case of perichondritis, ulcer formation up to septal perforation may occur. Patients report of nasal dryness, itchiness and crust formation. Aetiology is unknown. Physical, chemical (e.g. snuff, cocaine), and mechanical irritation (digital manipulation) are discussed.

Based on a progressively deteriorated mucociliary clearance, differentiation of rhinitis sicca to atrophic rhinitis is possible. For diagnosis, endoscopy, histological and microbiological examination, imaging and allergy tests are mandatory [273], because 60% have a sinusitis [274] and 85% an allergic disposition. Endonasal resistance if below normal, which is called ”paradoxical nasal obstruction“ in case of subjective nasal blockage [273]. At a sensitivity of 95% (specificity: 77%) an atrophic rhinitis is diagnosed, if the patient’s suffers from chronic sinusitis for more than six months and shows an additional two out of five features. These include: nasal bleeding, anosmia, purulent rhinorrhea, chronic upper airway infection, and having had two or more sinus surgeries [287].

Next to an infectious aetiology (due to Klebsiella ozaenae and Bacillus foetidus [273], [282]), environmental, endocrinologic and allergic-immunologic causes have been suspected [274], [282]. Atrophic rhinitis occurs more commonly near desserts [282] in patients with a relatively wide nasal cavity [288]. In parallel, surgical procedures narrowing the nose (up to its closure) cure the disease. From this background, disturbance of endonasal airflow was estimated as underlying pathophysiology. Accordingly, atrophic rhinitis was called a consequence of nasal cycle deterioration [289], [290].

With regards to treatment, no prospective controlled studies have been reported [291]. Empiric conservative treatment consists of local application of dexpanthenol and oil to resolve the crusts and to stimulate the regeneration of nasal mucosa. Sesame and soja oil improve ciliary beat frequency in contrast to essential oils [292]. Additional nasal douching, occasionally under supplementation of glucose (up to 25%), sole inhalation and sleeping next to an opened window have been recommended [293]. Antibiotics (rifampicin: [294]) have been used in case of purulent secretion [295] or as adjunct to surgical procedures (aminoglycoside: [296]), without being able to cure the disease [273]. Even closure of the nostrils e.g. by an obturator, has been suggested [297].

5.3 Vasomotor rhinitis

Vasomotor rhinitis (also called: hyperreflectoric rhinitis) is a disorder diagnosed by exclusion. Among non-allergic rhinitis, this subgroup is most common, affecting 60% of patients [298]. Aetiology is unclear with both episodical and perennial clinical manifestations occurring. A cholinergic course, going together with increased nasal secretion ('Runner'), can be distinguished from a clinical course characterised by nasal obstruction ('Blocker') mediated by nociceptive nerve fibre.

Reported trigger mechanisms include cold air [299], change of temperature, food (spices, alcohol), physical activity, smoke, dust and automobile exhaust.

Histologically, a strong expression of NO-synthetase is detected in subepithelial smooth muscle cells of the cavernous sinus [300]. Moreover, intense staining with 3-Nitrotyrosin, a product of NO-metabolism [300]. SP, CGRP, and VIP are increased, while IL-12 is diminished. Eosinophilic findings contradict the presence of vasomotor rhinitis [291].

A missing change of endonasal resistance after application of oxymetazoline supports diagnostically the hypothesis of etiologic changed autonomic innervation [301] and is helpful for posing the diagnosis. Measurements of heart frequency show a relative, parasympathetic hyperreactivity [302]. Based on standardised questionnaires, void of symptoms in spring and in presence of cats, missing allergic diseases of the parents, symptoms associated with perfumes and scents and age of first manifestation show a predicative value for suffering from vasomotor rhinitis [303].

Topical antihistamines [304], glucocorticoids [305] and ipatropium bromide are very effective in treating vasomotor rhinitis [298]. Interestingly, in a subgroup with complaints triggered by wind and change of temperature, Fluticasone treatment showed no significant benefit [306]. Locally injected botulinum toxine (I0 international units) improves subjective complaints of patients with vasomotor rhinitis for eight weeks [307]. Vidian neurectomy is reported to achieve five year lasting cessation of symptoms in 80% of patients. Acupuncture used in a placebo-controlled phase III study, showed a positive effect on a nasal symptom score [308].


6 Conclusion

Respiratory mucosa of the nose and paranasal sinuses exerts various functions, with their complex morphological and functional correlations becoming increasingly well understood.

Derangement of the respiratory function is one of the most frequent disorders of human beings. Due to available treatments with antibiotics and effective topical steroids, frequency of disorders is changing. Previously less common subtypes of chronic rhinitis gain increased importance. To encourage detailed pathophysiological understanding, selected disorders have been presented with their treatment options.

Early diagnostics and differentiated treatment can prevent continuous loss of function and development of consecutive disorders. This may help to avoid increased treatment costs. In this way, diagnosis and therapy of respiratory function gain next to their inherent individual importance increased relevance from the perspective of society.


References

1.
Lang J. Klinische Anatomie der Nase, Nasenhöhle und Nebenhöhlen: Grundlagen für Diagnostik und Operation. Stuttgart, New York: Thieme; 1988.
2.
Park KS, Wells JM, Zorn AM, Wert SE, Whitsett JA. Sox17 influences the differentiation of respiratory epithelial cells. Dev Biol. 2006; 294: 192-202. DOI: 10.1016/j.ydbio.2006.02.038 External link
3.
Badano JL, Mitsuma N, Beales PL, Katsanis N. The ciliopathies: an emerging class of human genetic disorders. Annu Rev Genomics Hum Genet. 2006; 7: 125-148. DOI: 10.1146/annurev.genom.7.080505.115610 External link
4.
Huang T, You Y, Spoor MS, Richer EJ, Kudva VV, Paige RC, Seiler MP, Liebler JM, Zabner J, Plopper CG,Brody SL. Foxj1 is required for apical localization of ezrin in airway epithelial cells. J Cell Sci. 2003; 116: 4935-4945. DOI: 10.1242/jcs.00830 External link
5.
Hayes JM, Kim SK, Abitua PB, Park TJ, Herrington ER, Kitayama A, Grow MW, Ueno N, Wallingford JB. Identification of novel ciliogenesis factors using a new in vivo model for mucociliary epithelial development. Dev Biol. 2007; 312: 115-130. DOI: 10.1016/j.ydbio.2007.09.031 External link
6.
Yin Y, Bangs F, Paton I, Prescott A, Joames J, Davey M, Whitley P, Genikhovich G, Technau U, Burt D,Tickle C. The Talpid3 gene (KIAA0586) encodes a centrosomal protein. Development. 2009; 136: 655-664. DOI: 10.1242/dev.028464 External link
7.
Mygind N, Pedersen M, Nielsen M. Morphology of the upper airway epithelium. In: Proctor D, Andersen I, eds. The Nose. Amsterdam: Elsevier; 1982.
8.
Widdicombe JH, Bastacky SJ, Wu DX, Lee CY. Regulation of depth and composition of airway surface liquid. Eur Respir J. 1997; 10: 2892-2897. DOI: 10.1183/09031936.97.10122892 External link
9.
Plopper CG, Nishio SJ, Alley JL, Kass P, Hyde DM. The role of the nonciliated bronchiolar epithelial (Clara) cell as the progenitor cell during bronchiolar epithelial differentiation in the perinatal rabbit lung. Am J Respir Cell Mol Biol. 1992; 7: 606-613.
10.
Proctor D, Andersen I. The Nose. Amsterdam: Elsevier; 1982.
11.
Winther B, Innes DJ Jr, Mills SE, Mygind N, Zito D, Hayden FG. Lymphocyte subsets in normal airway mucosa of the human nose. Arch Otolaryngol Head Neck Surg. 1987; 113: 59-62.
12.
Bradding P, Feather IH, Wilson S, Bardin PG, Heusser CH, Holgate ST, Howarth PH. Immunolocalization of cytokines in the nasal mucosa of normal and perennial rhinitic subjects. The mast cell as a source of IL-4, IL-5, and IL-6 in human allergic mucosal inflammation. J Immunol. 1993; 151: 3853-3865.
13.
Hosemann W, Kühnel T, Burchard AK, Werner JA. Histochemical detection of lymphatic drainage pathways in the middle nasal meatus. Rhinology. 1998; 36: 50-54.
14.
Pan WR, Suami H, Corlett RJ, Ashton MW. Lymphatic drainage of the nasal fossae and nasopharynx: preliminary anatomical and radiological study with clinical implications. Head Neck. 2009; 31: 52-57. DOI: 10.1002/hed.20926 External link
15.
Malcomson K. The vasomotor activities of the nasal mucous membrane. J Laryngol Otol. 1959; 73: 73-98. DOI: 10.1017/S0022215100054980 External link
16.
Fowler E. Unilateral vasomotor rhinitis due to interference with the cervical sympathetic system. Arch Otolaryngol. 1943; 37: 710-712.
17.
Golding-Wood P. The surgery of nasal allergy. Int Rhinol. 1963; 188-193.
18.
Baraniuk JN. Neuropeptides. Am J Rhinol. 1998; 12: 9-16. DOI: 10.2500/105065898782103025 External link
19.
Figueroa JM, Mansilla E, Suburo AM. Innervation of nasal turbinate blood vessels in rhinitic and nonrhinitic children. Am J Respir Crit Care Med. 1998; 157: 1959-1966.
20.
Knipping S. Untersuchungen zur Regulation der seromukösen Drüsen der respiratorischen Nasenschleimhaut des Menschen. Halle (Saale): Martin-Luther-Universität; 2004.
21.
Baraniuk JN, Silver PB, Kaliner MA, Barnes PJ. Neuropeptide Y is a vasoconstrictor in human nasal mucosa. J Appl Physiol. 1992; 73: 1867-1872.
22.
Baraniuk JN, Lundgren JD, Okayama M, Goff J, Mullol J, Merida M, Shelhamer JH, Kaliner MA. Substance P and neurokinin A in human nasal mucosa. Am J Respir Cell Mol Biol. 1991; 4: 228-236.
23.
Baraniuk JN, Merck SJ. Neuroregulation of human nasal mucosa. Ann N Y Acad Sci. 2009; 1170: 604-609. DOI: 10.1111/j.1749-6632.2009.04481.x External link
24.
Riederer A, Knipping S, Toleti B. Überlegungen zur Dynamik des Schwellgewebes der unteren Nasenmuschel des Menschen. Laryngorhinootologie. 2002; 81: 469-475. DOI: 10.1055/s-2002-33292 External link
25.
Cole P. Nasal and oral airflow resistors. Site, function, and assessment. Arch Otolaryngol Head Neck Surg. 1992; 118: 790-793.
26.
Gudziol H, Blau B, Stadeler M. Untersuchungen zur nasalen Depositionseffizienz für Weizenmehl- und Maisstärke-Staub. Laryngorhinootologie. 2009; 88: 398-404. DOI: 10.1055/s-0028-1119410 External link
27.
Zhou Y, Benson JM, Irvin C, Irshad H, Cheng YS. Particle size distribution and inhalation dose of shower water under selected operating conditions. Inhal Toxicol. 2007; 19: 333-342. DOI: 10.1080/08958370601144241 External link
28.
Lacroix JS, Potter EK. Nasonasal reflex mechanisms in anaesthetized dogs. Acta Otolaryngol. 1999; 119: 249-256. DOI: 10.1080/00016489950181765 External link
29.
Baraniuk JN, Kim D. Nasonasal reflexes, the nasal cycle, and sneeze. Curr Allergy Asthma Rep. 2007; 7: 105-111. DOI: 10.1007/s11882-007-0007-1 External link
30.
Macron JM, Wallois F, Duron B. Influence of vagal afferents in the sneeze reflex in cats. Neurosci Lett. 1994; 177: 79-82. DOI: 10.1016/0304-3940(94)90049-3 External link
31.
Xie X, Li Y, Chwang AT, Ho PL, Seto WH. How far droplets can move in indoor environments--revisiting the Wells evaporation-falling curve. Indoor Air. 2007; 17: 211-225. DOI: 10.1111/j.1600-0668.2007.00469.x External link
32.
Baraniuk JN, Merck SJ. Nasal reflexes: implications for exercise, breathing, and sex. Curr Allergy Asthma Rep. 2008; 8: 147-153. DOI: 10.1007/s11882-008-0025-7 External link
33.
Riederer A, Held B, Mack B. Immunhistochemische Untersuchungen zur Verteilung der konstitutiven Stickoxidsynthase in Gefässendothelien der Nasenschleimhaut des Menschen. Laryngorhinootologie. 1999; 78: 373-377. DOI: 10.1055/s-2007-996889 External link
34.
Sarkar MA. Drug metabolism in the nasal mucosa. Pharm Res. 1992; 9: 1-9. DOI: 10.1023/A:1018911206646 External link
35.
Hoehn T, Huebner J, Paboura E, Krause M, Leititis JU. Effect of therapeutic concentrations of nitric oxide on bacterial growth in vitro. Crit Care Med. 1998; 26: 1857-1862.
36.
Saloga J, Klimek L, Buhl R, Knop J. Allergologie-Handbuch: Grundlagen und klinische Praxis. Stuttgart: Schattauer; 2006.
37.
Quraishi MS, Jones NS, Mason J. The rheology of nasal mucus: a review. Clin Otolaryngol Allied Sci. 1998; 23: 403-413. DOI: 10.1046/j.1365-2273.1998.00172.x External link
38.
Wilson WR, Allansmith MR. Rapid, atraumatic method for obtaining nasal mucus samples. Ann Otol Rhinol Laryngol. 1976; 85: 391-393.
39.
Rahmoune H, Shephard KL. State of airway surface liquid on guinea pig trachea. J Appl Physiol. 1995; 78: 2020-2024.
40.
Gizurarson S. Animal models for intranasal drug delivery studies. A review article. Acta Pharm Nord. 1990; 2: 105-122.
41.
Alberty J, Stoll W, Rudack C. The effect of endogenous nitric oxide on mechanical ciliostimulation of human nasal mucosa. Clin Exp Allergy. 2006; 36: 1254-1259. DOI: 10.1111/j.1365-2222.2006.02563.x External link
42.
Laoukili J, Perret E, Willems T, Minty A, Parthoens E, Houcine O, Coste A, Jorissen M, Marano F, Caput D,Tournier F. IL-13 alters mucociliary differentiation and ciliary beating of human respiratory epithelial cells. J Clin Invest. 2001; 108: 1817-1824.
43.
Muns G, Singer P, Wolf F, Rubinstein I. Impaired nasal mucociliary clearance in long-distance runners. Int J Sports Med. 1995; 16: 209-213. DOI: 10.1055/s-2007-972993 External link
44.
Messerklinger W. Über die Sekretströmung auf der Schleimhaut der oberen Luftwege. Z Laryngol Rhinol Otol. 1951; 30: 302-308.
45.
Swart SJ, van der Baan S, Steenbergen JJ, Nauta JJ, van Kamp GJ, Biewenga J. Immunoglobulin concentrations in nasal secretions differ between patients with an IgE-mediated rhinopathy and a non-IgE-mediated rhinopathy. J Allergy Clin Immunol. 1991; 88: 612-619. DOI: 10.1016/0091-6749(91)90155-H External link
46.
Meredith SD, Raphael GD, Baraniuk JN, Banks SM, Kaliner MA. The pathophysiology of rhinitis. III. The control of IgG secretion. J Allergy Clin Immunol. 1989; 84: 920-930. DOI: 10.1016/0091-6749(89)90390-4 External link
47.
Wabnitz DA, Wormald PJ. A blinded, randomized, controlled study on the effect of buffered 0.9% and 3% sodium chloride intranasal sprays on ciliary beat frequency. Laryngoscope. 2005; 115: 803-805. DOI: 10.1097/01.MLG.0000157284.93280.F5 External link
48.
Harvey R, Hannan SA, Badia L, Scadding G. Nasal saline irrigations for the symptoms of chronic rhinosinusitis. Cochrane Database Syst Rev. 2007; CD006394. DOI: 10.1002/14651858.CD006394.pub2 External link
49.
Ural A, Oktemer TK, Kizil Y, Ileri F, Uslu S. Impact of isotonic and hypertonic saline solutions on mucociliary activity in various nasal pathologies: clinical study. J Laryngol Otol. 2009; 123: 517-521. DOI: 10.1017/S0022215108003964 External link
50.
Psaltis AJ, Wormald PJ, Ha KR, Tan LW. Reduced levels of lactoferrin in biofilm-associated chronic rhinosinusitis. Laryngoscope. 2008; 118: 895-901. DOI: 10.1097/MLG.0b013e31816381d4 External link
51.
Psaltis AJ, Ha KR, Beule AG, Tan LW, Wormald PJ. Confocal scanning laser microscopy evidence of biofilms in patients with chronic rhinosinusitis. Laryngoscope. 2007; 117: 1302-1306. DOI: 10.1097/MLG.0b013e31806009b0 External link
52.
Kayser R. Die exakte Messung der Luftdurchgängigkeit der Nase. Arch Laryngol Rhinol (Berl). 1895; 3: 101-120.
53.
Lang C, Grutzenmacher S, Mlynski B, Plontke S, Mlynski G. Investigating the nasal cycle using endoscopy, rhinoresistometry, and acoustic rhinometry. Laryngoscope. 2003; 113: 284-289. DOI: 10.1097/00005537-200302000-00016 External link
54.
Ohki M, Ogoshi T, Yuasa T, Kawano K, Kawano M. Extended observation of the nasal cycle using a portable rhinoflowmeter. J Otolaryngol. 2005; 34: 346-349. DOI: 10.2310/7070.2005.34509 External link
55.
Bamford OS, Eccles R. The central reciprocal control of nasal vasomotor oscillations. Pflügers Arch. 1982; 394: 139-143. DOI: 10.1007/BF00582915 External link
56.
Eccles R. Sympathetic control of nasal erectile tissue. Eur J Respir Dis Suppl. 1983; 128: 150-154.
57.
Kennedy DW, Zinreich SJ, Kumar AJ, Rosenbaum AE, Johns ME. Physiologic mucosal changes within the nose and ethmoid sinus: imaging of the nasal cycle by MRI. Laryngoscope. 1988; 98: 928-933.
58.
Soane RJ, Carney AS, Jones NS, Frier M, Perkins AC, Davis SS, Illum L. The effect of the nasal cycle on mucociliary clearance. Clin Otolaryngol. 2001; 26: 9-15. DOI: 10.1046/j.1365-2273.2001.00423.x External link
59.
Keck T, Leiacker R, Meixner D, Kuhnemann S, Rettinger G. Erwärmung der Atemluft in der Nase. HNO. 2001; 49: 36-40. DOI: 10.1007/s001060050705 External link
60.
Keck T. Untersuchungen zur Konditionierung der Atemluft in der Nase. Laryngorhinootologie. 2003; 82: 289-290. DOI: 10.1055/s-2003-38939 External link
61.
Eccles R. A role for the nasal cycle in respiratory defence. Eur Respir J. 1996; 9: 371-376. DOI: 10.1183/09031936.96.09020371 External link
62.
Mirza N, Kroger H, Doty RL. Influence of age on the 'nasal cycle'. Laryngoscope. 1997; 107: 62-66. DOI: 10.1097/00005537-199701000-00014 External link
63.
Eccles R. The central rhythm of the nasal cycle. Acta Otolaryngol. 1978; 86: 464-468.
64.
Grützenmacher S, Lang C, Mlynski R, Mlynski B, Mlynski G. Long-term rhinoflowmetry: a new method for functional rhinologic diagnostics. Am J Rhinol. 2005; 19: 53-57.
65.
Heetderks R. Observations on the reaction of normal nasal mucous membrane. Amer J Med Sci. 1927; 174: 231-242. DOI: 10.1097/00000441-192708000-00008 External link
66.
Hasegawa M, Kern EB. The human nasal cycle. Mayo Clin Proc. 1977; 52: 28-34.
67.
Fisher EW, Scadding GK, Lund VJ. The role of acoustic rhinometry in studying the nasal cycle. Rhinology. 1993; 31: 57-61.
68.
Cole P, Haight JS. Posture and the nasal cycle. Ann Otol Rhinol Laryngol. 1986; 95: 233-237.
69.
Gilbert AN. Biological Rhythmicity of nasal airway patency: a re-examination of the 'nasal cycle'. Acta Otolaryngol. 1987; 104: 180-186. DOI: 10.3109/00016488709109065 External link
70.
Shilenkova VV, Kozlov VS. A nasal cycle in healthy children. Vestn Otorinolaringol. 2008; 11-16.
71.
Atanasov AT, Dimov PD. Nasal and sleep cycle-possible synchronization during night sleep. Med Hypotheses. 2003; 61: 275-277. DOI: 10.1016/S0306-9877(03)00169-5 External link
72.
Cole P, Haight JS. Mechanisms of nasal obstruction in sleep. Laryngoscope. 1984; 94: 1557-1559. DOI: 10.1288/00005537-198412000-00004 External link
73.
Haight JS, Cole P. Is the nasal cycle an artifact? The role of asymmetrical postures. Laryngoscope. 1989; 99: 538-541s. DOI: 10.1288/00005537-198905000-00013 External link
74.
Eccles R. The effects of posture on nasal resistance to airflow. Rhinology. 1986; 24: 75.
75.
Costalonga M, Cleary P, Fischer L, Zhao Z. Intranasal bacteria induce Th1 but not Treg or Th2. Mucosal Immunology. 2009; 2: 85-95. DOI: 10.1038/mi.2008.67 External link
76.
Strutz J, Mann W. Praxis der HNO-Heilkunde, Kopf- und Halschirurgie. Stuttgart: Thieme; 2001. pp. 1071.
77.
Nadel DM, Lanza DC, Kennedy DW. Endoscopically guided sinus cultures in normal subjects. Am J Rhinol. 1999; 13: 87-90. DOI: 10.2500/105065899782106742 External link
78.
Al-Shemari H, Abou-Hamad W, Libman M, Desrosiers M. Bacteriology of the sinus cavities of asymptomatic individuals after endoscopic sinus surgery. J Otolaryngol. 2007; 36: 43-48. DOI: 10.2310/7070.2006.0019 External link
79.
Lackner A, Freudenschuss K, Buzina W, Stammberger H, Panzitt T, Schosteritsch S, Braun H. Ab wann sind Pilze im Nasensekret des Menschen nachweisbar? Laryngorhinootologie. 2004; 83: 117-121. DOI: 10.1055/s-2004-814208 External link
80.
Joniau S, Vlaminck S, Van Landuyt H, Kuhweide R, Dick C. Microbiology of sinus puncture versus middle meatal aspiration in acute bacterial maxillary sinusitis. Am J Rhinol. 2005; 19: 135-140.
81.
Uehara Y, Nakama H, Agematsu K, Uchida M, Kawakami Y, Abdul Fattah AS, Maruchi N. Bacterial interference among nasal inhabitants: eradication of Staphylococcus aureus from nasal cavities by artificial implantation of Corynebacterium sp. J Hosp Infect. 2000; 44: 127-133. DOI: 10.1053/jhin.1999.0680 External link
82.
Beule AG, Hosemann W. Wundheilung und postoperative Behandlung nach Nasennebenhöhlenoperationen. HNO. 2009; 57: 763-771. DOI: 10.1007/s00106-009-1938-8 External link
83.
Shaw CK, Cowin A, Wormald PJ. A study of the normal temporal healing pattern and the mucociliary transport after endoscopic partial and full-thickness removal of nasal mucosa in sheep. Immunol Cell Biol. 2001; 79: 145-148. DOI: 10.1046/j.1440-1711.2001.00982.x External link
84.
Watelet JB, Bachert C, Gevaert P, Van Cauwenberge P. Wound healing of the nasal and paranasal mucosa: a review. Am J Rhinol. 2002; 16: 77-84.
85.
Eloy P, Watelet JB, Donckier J, Gustin T, Gaudon IP, Collet S, Rombaux P, Gillard C, Bertrand B. Endoscopic and microscopic paraseptal transsphenoidal approach to the sella turcica. Rhinology. 2005; 43: 271-276.
86.
Rajapaksa SP, Ananda A, Cain TM, Oates L, Wormald PJ. Frontal ostium neo-osteogenesis and restenosis after modified endoscopic Lothrop procedure in an animal model. Clin Otolaryngol Allied Sci. 2004; 29: 386-388. DOI: 10.1111/j.1365-2273.2004.00824.x External link
87.
Hosemann W, Wigand ME, Göde U, Langer F, Dunker I. Normal wound healing of the paranasal sinuses: clinical and experimental investigations. Eur Arch Otorhinolaryngol. 1991; 248: 390-394. DOI: 10.1007/BF01463560 External link
88.
Mauro T. Natural course of wound repair versus impaired healing in chronic cutaneous ulcers. In: Shai A, Maibach H, eds. Wound healing and ulcers of the skin. Diagnosis and therapy - the practical approach. Berlin, Heidelberg, New York: Springer; 2005. pp. 7-17.
89.
Watelet JB, Gevaert P, Bachert C, Holtappels G, van Cauwenberge P. Secretion of TGF-betal, TGF-beta2, EGF and PDGF into nasal fluid after sinus surgery. Eur Arch Otorhinolaryngol. 2002; 259: 234-238. DOI: 10.1007/s00405-002-0448-z External link
90.
Singer AJ, Clark RA. Cutaneous wound healing. N Engl J Med. 1999; 341: 738-746. DOI: 10.1056/NEJM199909023411006 External link
91.
Weber R, Keerl R. Einsatz moderner Bilddatenverarbeitung in der klinisch-rhinologischen Forschung. Eur Arch Otorhinolaryngol. 1996; Suppl 1: 271-296.
92.
Hosemann W, Göde U, Länger F, Röckelein G, Wigand ME. Experimentelle Untersuchungen zur Wundheilung in den Nasennebenhöhlen. I. Ein Modell respiratorische Wunden in der Kaninchen-Kieferhöhle. HNO. 1991; 39: 8-12.
93.
Beule AG, Steinmeier E, Kaftan H, Biebler KE, Göpferich A, Wolf E, Hosemann W. Effects of a dexamethasone-releasing stent on osteoneogenesis in a rabbit model. Am J Rhinol Allergy. 2009; 23: 433-436. DOI: 10.2500/ajra.2009.23.3331 External link
94.
Kurkinen M, Vaheri A, Roberts P, Stenmam S. Sequential appearance of fibronectin and collagen in experimental granulation tissue. Lab Invest. 1980; 43: 47-51.
95.
Goslen J. Wound healing for the dermatologic surgeon. J Dermatol Surg Oncol. 1988; 14: 959-972.
96.
Beule AG, Scharf C, Biebler KE, Göpferich A, Steinmeier E, Wolf E, Hosemann W, Kaftan H. Effects of topically applied dexamethasone on mucosal wound healing using a drug-releasing stent. Laryngoscope. 2008; 118: 2073-2077. DOI: 10.1097/MLG.0b013e3181820896 External link
97.
Forsgren K, Stierna P, Kumlien J, Carlsoo B. Regeneration of maxillary sinus mucosa following surgical removal. Experimental study in rabbits. Ann Otol Rhinol Laryngol. 1993; 102: 459-466.
98.
Forsgren K, Jung YG, Stierna P, Rivero C. Regeneration of nerve fibres in the maxillary sinus mucosa after experimental surgery. An immunocytochemical double-labelling study in the rabbit. Acta Otolaryngol. 1999; 119: 486-491. DOI: 10.1080/00016489950181044 External link
99.
Hosemann W, Dunker I, Göde U, Wigand ME. Experimentelle Untersuchungen zur Wundheilung in den Nasennebenhöhlen. III. Endoskopie und Histologie des Operationsgebietes nach einer endonasalen Siebbeinausräumung. HNO. 1991; 39: 111-115.
100.
Moriyama H, Yanagi K, Ohtori N, Asai K, Fukami M. Healing process of sinus mucosa after endoscopic sinus surgery. Am J Rhinology. 1996; 10: 61-66. DOI: 10.2500/105065896781795067 External link
101.
Huang HM, Cheng JJ, Liu CM, Lin KN. Mucosal healing and mucociliary transport change after endoscopic sinus surgery in children with chronic maxillary sinusitis. Int J Pediatr Otorhinolaryngol. 2006; 70: 1361-1367. DOI: 10.1016/j.ijporl.2006.01.016 External link
102.
Inanli S, Tutkun A, Batman C, Okar I, Uneri C, Sehitoglu MA. The effect of endoscopic sinus surgery on mucociliary activity and healing of maxillary sinus mucosa. Rhinology. 2000; 38: 120-123.
103.
Forsgren K, Otori N, Stierna P, Kumlien J. Microvasculature, blood flow, and vasoreactivity in rabbit sinus mucosa after surgery. Laryngoscope. 1999; 109: 562-568. DOI: 10.1097/00005537-199904000-00008 External link
104.
Behrbohm H, Sydow K. Nuklearmedizinische Untersuchungen zum Reparationsverhalten der Kieferhöhlenschleimhaut nach FES. HNO. 1991; 39: 173-176.
105.
Vu T, Werb Z. Gelatinase B: Structure, Regulation, and Function, in matrix metalloproteinases. San Diego: Academic Press; 1998. pp. 115-148.
106.
Van den Steen P, Dubois B, Nelissen I, Rudd P, Dwek R, Opdenakker G. Biochemistry and molecular biology of gelatinase B or matrix metalloproteinase-9 (MMP-9). Crit Rev Biochem Mol Biol. 2002; 37: 375-536. DOI: 10.1080/10409230290771546 External link
107.
Fini ME, Parks WC, Rinehart WB, Girard MT, Matsubara M, Cook JR, West-Mays JA, Sadow PM, Burgeson RE, Jeffrey JJ, Raizman MB, Krueger RR,Zieske JD. Role of matrix metalloproteinases in failure to re-epithelialize after corneal injury. Am J Pathol. 1996; 149: 1287-1302.
108.
Mohan R, Chintala SK, Jung JC, Villar WV, McCabe F, Russo LA, Lee Y, McCarthy BE, Wollenberg KR, Jester JV, Wang M, Welgus HG, Shipley JM, Senior RM,Fini ME. Matrix metalloproteinase gelatinase B (MMP-9) coordinates and effects epithelial regeneration. J Biol Chem. 2002; 277: 2065-2072. DOI: 10.1074/jbc.M107611200 External link
109.
Watelet JB, Claeys C, Van Cauwenberge P, Bachert C. Predictive and monitoring value of matrix metalloproteinase-9 for healing quality after sinus surgery. Wound Repair Regen. 2004; 12: 412-418. DOI: 10.1111/j.1067-1927.2004.012411.x External link
110.
Sun D, Matsune S, Ohori J, Fukuiwa T, Ushikai M, Kurono Y. TNF-alpha and endotoxin increase hypoxia-induced VEGF production by cultured human nasal fibroblasts in synergistic fashion. Auris Nasus Larynx. 2005; 32: 243-249. DOI: 10.1016/j.anl.2005.01.004 External link
111.
Urschel S, Kayikci L, Wintergerst U, Notheis G, Jansson A, Belohradsky BH. Common variable immunodeficiency disorders in children: delayed diagnosis despite typical clinical presentation. J Pediatr. 2009; 154: 888-894. DOI: 10.1016/j.jpeds.2008.12.020 External link
112.
Atef A, Zeid IA, Qotb M, El Rab EG. Effect of passive smoking on ciliary regeneration of nasal mucosa after functional endoscopic sinus surgery in children. J Laryngol Otol. 2009; 123: 75-79. DOI: 10.1017/S0022215108003678 External link
113.
Tran KN, Beule AG, Singal D, Wormald PJ. Frontal ostium restenosis after the endoscopic modified Lothrop procedure. Laryngoscope. 2007; 117: 1457-1462. DOI: 10.1097/MLG.0b013e31806865be External link
114.
Mlynski G, Beule A. Diagnostik der respiratorischen Funktion der Nase. HNO. 2008; 56: 81-99. DOI: 10.1007/s00106-007-1655-0 External link
115.
Yanagisawa E. The use of video in ENT endoscopy: its value in teaching. Ear Nose Throat J. 1994; 73: 754-763.
116.
Athanasiadis T, Beule A, Embate J, Steinmeier E, Field J, Wormald PJ. Standardized video-endoscopy and surgical field grading scale for endoscopic sinus surgery: a multi-centre study. Laryngoscope. 2008; 118: 314-319. DOI: 10.1097/MLG.0b013e318157f764 External link
117.
Fokkens W, Lund V, Mullol J. EP3OS 2007: European position paper on rhinosinusitis and nasal polyps 2007. A summary for otorhinolaryngologists. Rhinology. 2007; 45: 97-101.
118.
Bhattacharyya N, Fried MP. The accuracy of computed tomography in the diagnosis of chronic rhinosinusitis. Laryngoscope. 2003; 113: 125-129. DOI: 10.1097/00005537-200301000-00023 External link
119.
Howells RC, Ramadan HH. Usefulness of computed tomography and magnetic resonance in fulminant invasive fungal rhinosinusitis. Am J Rhinol. 2001; 15: 255-261.
120.
Lim WK, Kumar M. Computed tomography of rhinosinusitis in immunodeficient patients: not only fungal sinusitis. Arch Otolaryngol Head Neck Surg. 2003; 129: 1355-1356. DOI: 10.1001/archotol.129.12.1355-b External link
121.
Bhattacharyya N. Test-retest reliability of computed tomography in the assessment of chronic rhinosinusitis. Laryngoscope. 1999; 109: 1055-1058. DOI: 10.1097/00005537-199907000-00008 External link
122.
Ganz H. Hals-Nasen-Ohren-Heilkunde. Berlin, New York: Walter de Gruyter; 1996.
123.
Mahmood U, Ridgway J, Jackson R, Guo S, Su J, Armstrong W, Shibuya T, Crumley R, Chen Z,Wong B. In vivo optical coherence tomography of the nasal mucosa. Am J Rhinol. 2006; 20: 155-159.
124.
Canciani M, Barlocco EG, Mastella G, de Santi MM, Gardi C, Lungarella G. The saccharin method for testing mucociliary function in patients suspected of having primary ciliary dyskinesia. Pediatr Pulmonol. 1988; 5: 210-214. DOI: 10.1002/ppul.1950050406 External link
125.
Nathan RA, Eccles R, Howarth PH, Steinsvag SK, Togias A. Objective monitoring of nasal patency and nasal physiology in rhinitis. J Allergy Clin Immunol. 2005; 115: S442-459. DOI: 10.1016/j.jaci.2004.12.015 External link
126.
Kleinschmidt EG, Witt G. Zur Beurteilung der nasalen Mukoziliartätigkeit mit einem modifizierten Saccharintest. Laryngorhinootologie. 1995; 74: 286-288. DOI: 10.1055/s-2007-997741 External link
127.
Di Giuda D, Galli J, Calcagni ML, Corina L, Paludetti G, Ottaviani F, De Rossi G. Rhinoscintigraphy: a simple radioisotope technique to study the mucociliary system. Clin Nucl Med. 2000; 25: 127-130. DOI: 10.1097/00003072-200002000-00010 External link
128.
Deitmer T. A modification of the saccharine test for nasal mucociliary clearance. Rhinology. 1986; 24: 237-240.
129.
Sun SS, Hsieh JF, Tsai SC, Ho YJ, Kao CH. The role of rhinoscintigraphy in the evaluation of nasal mucociliary clearance function in patients with sinusitis. Nucl Med Commun. 2000; 21: 1029-1032. DOI: 10.1097/00006231-200011000-00007 External link
130.
Peksoy I, Ugur MB, Altin R, Cinar F, Uzun L, Cabuk M, Kart L. Evaluation of nasal mucociliary functions with rhinoscintigraphy in coal workers' pneumoconiosis. ORL J Otorhinolaryngol Relat Spec. 2005; 67: 163-167. DOI: 10.1159/000086470 External link
131.
Taylor RE. Abnormal mucociliary transport study in a patient with Kartagener syndrome. Clin Nucl Med. 2006; 31: 240-242. DOI: 10.1097/01.rlu.0000204550.59867.e2 External link
132.
Hüttenbrink KB, Wrede H, Lagemann S, Schleicher E, Hummel T. Lokalisierte, endonasale Messung der mukoziliaren Transportgeschwindigkeit als neues Verfahren für die nasale Diagnostik. Laryngorhinootologie. 2006; 85: 24-31. DOI: 10.1055/s-2005-870354 External link
133.
Takeuchi K, Suzumura E, Majima Y, Sakakura Y. Effect of atropine on nasal mucociliary clearance. Acta Otolaryngol. 1990; 110: 120-123. DOI: 10.3109/00016489009122525 External link
134.
Caruso G, Gelardi M, Passali GC, de Santi MM. Nasal scraping in diagnosing ciliary dyskinesia. Am J Rhinol. 2007; 21: 702-705. DOI: 10.2500/ajr.2007.21.3107 External link
135.
Athanasiadis T, Beule AG, Robinson BH, Robinson SR, Shi Z, Wormald PJ. Effects of a novel chitosan gel on mucosal wound healing following endoscopic sinus surgery in a sheep model of chronic rhinosinusitis. Laryngoscope. 2008; 118: 1088-1094. DOI: 10.1097/MLG.0b013e31816ba576 External link
136.
Dimova S, Maes F, Brewster ME, Jorissen M, Noppe M, Augustijns P. High-speed digital imaging method for ciliary beat frequency measurement. J Pharm Pharmacol. 2005;57(4):521-6.
137.
Jorissen M, Bessems. Normal ciliary beat frequency after ciliogenesis in nasal epithelial cells cultured sequentially as monolayer and in suspension. Acta Otolaryngol. 1995; 115: 66-70. DOI: 10.3109/00016489509133349 External link
138.
Jorissen M, Willems T. Ciliary function analysis for the diagnosis of primary ciliary dyskinesia. Acta Otolaryngol. 2000; 120: 291-295. DOI: 10.1080/000164800750001116 External link
139.
Abanses JC, Arima S, Rubin BK. Vicks VapoRub induces mucin secretion, decreases ciliary beat frequency, and increases tracheal mucus transport in the ferret trachea. Chest. 2009; 135: 143-148. DOI: 10.1378/chest.08-0095 External link
140.
Even-Tzur N, Kloog Y, Wolf M, Elad D. Mucus secretion and cytoskeletal modifications in cultured nasal epithelial cells exposed to wall shear stresses. Biophys J. 2008; 95: 2998-3008. DOI: 10.1529/biophysj.107.127142 External link
141.
Watelet JB, Gevaert P, Holtappels G, Van Cauwenberge P, Bachert C. Collection of nasal secretions for immunological analysis. Eur Arch Otorhinolaryngol. 2004; 261: 242-246. DOI: 10.1007/s00405-003-0691-y External link
142.
Ha KR, Psaltis AJ, Tan L, Wormald PJ. A sheep model for the study of biofilms in rhinosinusitis. Am J Rhinol. 2007; 21: 339-345. DOI: 10.2500/ajr.2007.21.3032 External link
143.
Beule AG, Hosemann W. Bakterielle Biofilme. Laryngorhinootologie. 2007; 86: 886-895. DOI: 10.1055/s-2007-967031 External link
144.
Clement PA, Gordts F. Consensus report on acoustic rhinometry and rhinomanometry. Rhinology. 2005; 43: 169-179.
145.
Clement PA. Committee report on standardization of rhinomanometry. Rhinology. 1984; 22: 151-155.
146.
Snow JB, Ashley WP. Ballenger's Otorhinolaryngology Head and Neck Surgery. McGrath-Hill Professional; 2009.
147.
Thulesius HL, Thulesius HO, Jessen M. What happens to patients with nasal stuffiness and pathological rhinomanometry left without surgery? Rhinology. 2009; 47: 24-27.
148.
Hilberg O. Objective measurement of nasal airway dimensions using acoustic rhinometry: methodological and clinical aspects. Allergy. 2002; 57 Suppl 70: 5-39. DOI: 10.1046/j.0908-665x.2001.all.doc.x External link
149.
Bermuller C, Kirsche H, Rettinger G, Riechelmann H. Diagnostic accuracy of peak nasal inspiratory flow and rhinomanometry in functional rhinosurgery. Laryngoscope. 2008; 118: 605-610. DOI: 10.1097/MLG.0b013e318161e56b External link
150.
Mlynski G, Low J. Die Rhinoresistometrie-eine Weiterentwicklung der Rhinomanometrie. Laryngorhinootologie. 1993; 72: 608-610. DOI: 10.1055/s-2007-997964 External link
151.
Grützenmacher S, Mlynski G, Mlynski B, Lang C. Die Objektivierung des Schwellungszustandes der Nasenschleimhaut-ein Vergleich von vier Messmethoden. Laryngorhinootologie. 2003; 82: 645-649. DOI: 10.1055/s-2003-42684 External link
152.
Temmel AF, Toth J, Marks B, Jager S, Berger U, Reiser K, Horak F. Rhinoresistometry versus rhinomanometry-an evaluation. Wien Klin Wochenschr. 1998; 110: 612-615.
153.
Mayhew TM, O'Flynn P. Validation of acoustic rhinometry by using the Cavalieri principle to estimate nasal cavity volume in cadavers. Clin Otolaryngol Allied Sci. 1993; 18: 220-225. DOI: 10.1111/j.1365-2273.1993.tb00835.x External link
154.
Hilberg O, Jackson AC, Swift DL, Pedersen OF. Acoustic rhinometry: evaluation of nasal cavity geometry by acoustic reflection. J Appl Physiol. 1989; 66: 295-303.
155.
Burres SA. Acoustic rhinometry of the oriental nose. Am J Rhinol. 1999; 13: 407-410. DOI: 10.2500/105065899781367519 External link
156.
Straszek SP, Schlunssen V, Sigsgaard T, Pedersen OF. Reference values for acoustic rhinometry in decongested school children and adults: the most sensitive measurement for change in nasal patency. Rhinology. 2007; 45: 36-39.
157.
Cankurtaran M, Celik H, Coskun M, Hizal E, Cakmak O. Acoustic rhinometry in healthy humans: accuracy of area estimates and ability to quantify certain anatomic structures in the nasal cavity. Ann Otol Rhinol Laryngol. 2007; 116: 906-916.
158.
Riechelmann H, Rheinheimer MC, Wolfensberger M. Acoustic rhinometry in pre-school children. Clin Otolaryngol Allied Sci. 1993; 18: 272-277. DOI: 10.1111/j.1365-2273.1993.tb00846.x External link
159.
Straszek SP, Moeller A, Hall GL, Zhang G, Stick SM, Franklin PJ. Reference values for acoustic rhinometry in children from 4 to 13 years old. Am J Rhinol. 2008; 22: 285-291. DOI: 10.2500/ajr.2008.22.3147 External link
160.
Tantilipikorn P, Jareoncharsri P, Voraprayoon S, Bunnag C, Clement PA. Acoustic rhinometry of Asian noses. Am J Rhinol. 2008; 22: 617-620. DOI: 10.2500/ajr.2008.22.3229 External link
161.
Miyamoto Y, Takeuchi K, Majima Y. Measurement of nasal patency by acoustic rhinometry in Japanese school children. Auris Nasus Larynx. 2009; 36: 406-410. DOI: 10.1016/j.anl.2008.09.001 External link
162.
Haight JS, Djupesland PG, Qjan W, Chatkin JM, Furlott H, Irish J, Witterick I, McClean P, Fenton RS, Hoffstein V,Zamel N. Does nasal nitric oxide come from the sinuses? J Otolaryngol. 1999; 28: 197-204.
163.
Kim JW, Min YG, Rhee CS, Lee CH, Koh YY, Rhyoo C, Kwon TY, Park SW. Regulation of mucociliary motility by nitric oxide and expression of nitric oxide synthase in the human sinus epithelial cells. Laryngoscope. 2001; 111: 246-250. DOI: 10.1097/00005537-200102000-00011 External link
164.
Durland WF Jr, Lane AP, Durland KW, Smith TL, Johnson KL, Prazma J, Pillsbury HC. Nitric oxide is a mediator of the late-phase response in an animal model of nasal allergy. Otolaryngol Head Neck Surg. 2000; 122: 706-711. DOI: 10.1016/S0194-5998(00)70201-4 External link
165.
Silkoff PE, McClean PA, Caramori M, Slutsky AS, Zamel N. A significant proportion of exhaled nitric oxide arises in large airways in normal subjects. Respir Physiol. 1998; 113: 33-38. DOI: 10.1016/S0034-5687(98)00033-4 External link
166.
Di Nardo W, Pitocco D, Di Leo MA, Picciotti PM, Di Stasio E, Collina C, Santini S, Scarano E, Ghirlanda G. Modifications in nasal function and nitric oxide serum level in type 1 diabetes. J Otolaryngol Head Neck Surg. 2008; 37: 611-615.
167.
Hogman M, Drca N, Ehrstedt C, Merilainen P. Exhaled nitric oxide partitioned into alveolar, lower airways and nasal contributions. Respir Med. 2000; 94: 985-991. DOI: 10.1053/rmed.2000.0872 External link
168.
Qian W, Sabo R, Ohm M, Haight JS, Fenton RS. Nasal nitric oxide and the nasal cycle. Laryngoscope. 2001; 111: 1603-1607. DOI: 10.1097/00005537-200109000-00021 External link
169.
Kirihene RK, Rees G, Wormald PJ. The influence of the size of the maxillary sinus ostium on the nasal and sinus nitric oxide levels. Am J Rhinol. 2002; 16: 261-264.
170.
Silkoff P, Chatkin J, Qian W, Chakravorty S, Gutierrez C, Furlott H, McClean P, Rai S, Zamel N, Haight J. Nasal nitric oxide: a comparison of measurement techniques. Am J Rhinol. 1999; 13: 169-178. DOI: 10.2500/105065899781389803 External link
171.
American Thoracic Society; European Respiratory Society. ATS/ERS recommendations for standardized procedures for the online and offline measurement of exhaled lower respiratory nitric oxide and nasal nitric oxide, 2005. Am J Respir Crit Care Med. 2005; 171: 912-930. DOI: 10.1164/rccm.200406-710ST External link
172.
Gupta R, Gupta N, Turner SW. A methodology for measurements of nasal nitric oxide in children under 5 yr. Pediatr Allergy Immunol. 2008; 19: 233-238. DOI: 10.1111/j.1399-3038.2007.00616.x External link
173.
Qian W, Chatkin JM, Djupesland PG, McClean P, Zamel N, Irish JC, Haight JS. Unilateral nasal nitric oxide measurement after nasal surgery. Ann Otol Rhinol Laryngol. 2000; 109: 952-957.
174.
Bommarito L, Guida G, Heffler E, Badiu I, Nebiolo F, Usai A, De Stefani A, Rolla G. Nasal nitric oxide concentration in suspected chronic rhinosinusitis. Ann Allergy Asthma Immunol. 2008; 101: 358-362. DOI: 10.1016/S1081-1206(10)60310-9 External link
175.
Delclaux C, Malinvaud D, Chevalier-Bidaud B, Callens E, Mahut B, Bonfils P. Nitric oxide evaluation in upper and lower respiratory tracts in nasal polyposis. Clin Exp Allergy. 2008; 38: 1140-1147. DOI: 10.1111/j.1365-2222.2008.03006.x External link
176.
Santamaria F, De Stefano S, Montella S, Barbarano F, Iacotucci P, Ciccarelli R, Sofia M, Maniscalco M. Nasal nitric oxide assessment in primary ciliary dyskinesia using aspiration, exhalation, and humming. Med Sci Monit. 2008; 14: CR80-85.
177.
Wodehouse T, Kharitonov SA, Mackay IS, Barnes PJ, Wilson R, Cole PJ. Nasal nitric oxide measurements for the screening of primary ciliary dyskinesia. Eur Respir J. 2003; 21: 43-47. DOI: 10.1183/09031936.03.00305503 External link
178.
Monto AS. Epidemiology of viral respiratory infections. Am J Med. 2002; 112 Suppl 6A: 4S-12S. DOI: 10.1016/S0002-9343(01)01058-0 External link
179.
American Academy of Pediatrics. Subcommittee on Management of Sinusitis and Committee on Quality Improvement. Clinical practice guideline: management of sinusitis. Pediatrics. 2001; 108: 798-808.
180.
Aitken M, Taylor JA. Prevalence of clinical sinusitis in young children followed up by primary care pediatricians. Arch Pediatr Adolesc Med. 1998; 152: 244-248.
181.
LaFranchi S. Disorders of the Thyroid gland. In: Behrman R, Klingman R, Jenson H, eds. Nelson's Textbook of pediatrics. Philadelphia: WB Saunders; 2004.
182.
Belenky W, Madgy D. Nasal obstruction. In: Bluestone C, Stoll S, Kenna M, eds. Pediatric Otolaryngology. Philadelphia: WB Saunders; 1996.
183.
Mur Sierra A, Vinolas Tolosa M, Sanchez Garcia-Vao C, Garcia Lopez AC, Busquets Monge RM, Garcia Algar O, Lloreta Trull J, Bargues Cardelus R. Asociacion entre el consumo de heroina durante la gestacion y anomalias estructurales de los cilios respiratorios en el periodo neonatal. An Esp Pediatr. 2001; 55: 335-338.
184.
Wallace DV, Dykewicz MS, Bernstein DI, Blessing-Moore J, Cox L, Khan DA, Lang DM, Nicklas RA, Oppenheimer J, Portnoy JM, Randolph CC, Schuller D, Spector SL,Tilles SA. The diagnosis and management of rhinitis: an updated practice parameter. J Allergy Clin Immunol. 2008; 122: S1-84. DOI: 10.1016/j.jaci.2008.06.003 External link
185.
Ballow M, O'neil K. Approach to the patient with recurrent infections. In: Middleton E, Reed C, Ellis E, Adkinson N, Younginer I, Busse W, eds. Allergy principles and practice. St. Louis: Mosby; 1993.
186.
Tasker A, Dettmar PW, Panetti M, Koufman JA, J PB, Pearson JP. Is gastric reflux a cause of otitis media with effusion in children? Laryngoscope. 2002; 112: 1930-1934. DOI: 10.1097/00005537-200211000-00004 External link
187.
Alobid I, Guilemany JM, Mullol J. Nasal manifestations of systemic illnesses. Curr Allergy Asthma Rep. 2004; 4: 208-216. DOI: 10.1007/s11882-004-0028-y External link
188.
Boot JM, Oranje AP, de Groot R, Tan G, Stolz E. Congenital syphilis. Int J STD AIDS. 1992; 3: 161-167.
189.
Scadding GK. Non-allergic rhinitis: diagnosis and management. Curr Opin Allergy Clin Immunol. 2001; 1: 15-20.
190.
Toppozada H. The human nasal mucosa in the menopause (a histochemical and electron microscopic study). J Laryngol Otol. 1988; 102: 314-318.
191.
Getchell ML, Chen Y, Ding X, Sparks DL, Getchell TV. Immunohistochemical localization of a cytochrome P-450 isozyme in human nasal mucosa: age-related trends. Ann Otol Rhinol Laryngol. 1993; 102: 368-374.
192.
Nappi C, Di Spiezio Sardo A, Guerra G, Di Carlo C, Bifulco G, Acunzo G, Sammartino A, Galli V. Comparison of intranasal and transdermal estradiol on nasal mucosa in postmenopausal women. Menopause. 2004; 11: 447-455. DOI: 10.1097/01.GME.0000113849.74835.53 External link
193.
Caruso S, Grillo C, Agnello C, Di Mari L, Farina M, Serra A. Olfactometric and rhinomanometric outcomes in post-menopausal women treated with hormone therapy: a prospective study. Hum Reprod. 2004; 19: 2959-2964. DOI: 10.1093/humrep/deh465 External link
194.
Edelstein DR. Aging of the normal nose in adults. Laryngoscope. 1996; 106: 1-25. DOI: 10.1097/00005537-199609001-00001 External link
195.
Agius AM, Smallman LA, Pahor AL. Age, smoking and nasal ciliary beat frequency. Clin Otolaryngol Allied Sci. 1998; 23: 227-230. DOI: 10.1046/j.1365-2273.1998.00141.x External link
196.
Ho JC, Chan KN, Hu WH, Lam WK, Zheng L, Tipoe GL, Sun J, Leung R, Tsang KW. The effect of aging on nasal mucociliary clearance, beat frequency, and ultrastructure of respiratory cilia. Am J Respir Crit Care Med. 2001; 163: 983-988.
197.
Kalmovich LM, Elad D, Zaretsky U, Adunsky A, Chetrit A, Sadetzki S, Segal S, Wolf M. Endonasal geometry changes in elderly people: acoustic rhinometry measurements. J Gerontol A Biol Sci Med Sci. 2005; 60: 396-398.
198.
Settipane GA, Chafee FH. Nasal polyps in asthma and rhinitis. A review of 6,037 patients. J Allergy Clin Immunol. 1977; 59: 17-21. DOI: 10.1016/0091-6749(77)90171-3 External link
199.
Joe S, Benson A. Nonallergic rhinitis. In: Cummings C, eds. Otolaryngology Head and Neck Surgery. Phildalphia: Elsevier Mosby; 2005. pp. 990-1000.
200.
Bateman ND, Woolford TJ. The rhinological side-effects of systemic drugs. Clin Otolaryngol Allied Sci. 2003; 28: 381-385. DOI: 10.1046/j.1365-2273.2003.00712.x External link
201.
Muller LM, Gorter KJ, Hak E, Goudzwaard WL, Schellevis FG, Hoepelman AI, Rutten GE. Increased risk of common infections in patients with type 1 and type 2 diabetes mellitus. Clin Infect Dis. 2005; 41: 281-288. DOI: 10.1086/431587 External link
202.
Ahluwalia A, Sood A, Lakshmy R, Kapil A, Pandey RM. Nasal colonization with Staphylococcus aureus in patients with diabetes mellitus. Diabet Med. 2000; 17: 487-488.
203.
Saxena AK, Panhotra BR, Venkateshappa CK, Sundaram DS, Naguib M, Uzzaman W, Al Mulhim K. The impact of nasal carriage of methicillin-resistant and methicillin-susceptible Staphylococcus a ureus (MRSA & MSSA) on vascular access-related septicemia among patients with type-II diabetes on dialysis. Ren Fail. 2002; 24: 763-777. DOI: 10.1081/JDI-120015679 External link
204.
Panierakis C, Goulielmos G, Mamoulakis D, Maraki S, Papavasiliou E, Galanakis E. Staphylococcus aureus nasal carriage might be associated with vitamin D receptor polymorphisms in type 1 diabetes. Int J Infect Dis. Int J Infect Dis. 2009 Nov;13(6):e437-43. DOI: 10.1016/j.ijid.2009.02.012 External link
205.
Selimoglu MA, Selimoglu E, Kurt A. Nasal mucociliary clearance and nasal and oral pH in patients with insulin-dependent diabetes. Ear Nose Throat J. 1999; 78: 585-588, 590.
206.
Deniz M, Uslu C, Ogredik EA, Akduman D, Gursan SO. Nasal mucociliary clearance in total laryngectomized patients. Eur Arch Otorhinolaryngol. 2006; 263: 1099-1104. DOI: 10.1007/s00405-006-0111-1 External link
207.
Sachdeva A, Sachdeva OP, Gulati SP, Kakkar V. Nasal mucociliary clearance & mucus pH in patients with diabetes mellitus. Indian J Med Res. 1993; 98: 265-268.
208.
Yue WL. Nasal mucociliary clearance in patients with diabetes mellitus. J Laryngol Otol. 1989; 103: 853-855. DOI: 10.1017/S0022215100110291 External link
209.
Han DH, An SY, Kim SW, Kim DY, Rhee CS, Lee CH, Min YG. Primary and secondary fungal infections of the paranasal sinuses – clinical features and treatment outcomes. Acta Otolaryngol Suppl. 2007; 78-82. DOI: 10.1080/03655230701624913 External link
210.
Rashid M, Bari A, Majeed S, Tariq KM, Haq I, Niwaz A. Mucormycosis: a devastating fungal infection in diabetics. J Coll Physicians Surg Pak. 2005; 15: 43-45.
211.
Deshazo RD. Syndromes of invasive fungal sinusitis. Med Mycol. 2009; 47 Suppl 1: S309-314. DOI: 10.1080/13693780802213399 External link
212.
Cannady SB, Batra PS, Citardi MJ, Lanza DC. Comparison of delivery of topical medications to the paranasal sinuses via "vertex-to-floor" position and atomizer spray after FESS. Otolaryngol Head Neck Surg. 2005; 133: 735-740. DOI: 10.1016/j.otohns.2005.07.039 External link
213.
Karagama YG, Lancaster JL, Karkanevatos A, O'Sullivan G. Delivery of nasal drops to the middle meatus: which is the best head position? Rhinology. 2001; 39: 226-229.
214.
DelGaudio JM, Wise SK. Topical steroid drops for the treatment of sinus ostia stenosis in the postoperative period. Am J Rhinol. 2006; 20: 563-567. DOI: 10.2500/ajr.2006.20.2904 External link
215.
Valentine R, Athanasiadis T, Thwin M, Singhal D, Weitzel EK, Wormald PJ. A prospective controlled trial of pulsed nasal nebulizer in maximally dissected cadavers. Am J Rhinol. 2008; 22: 390-394. DOI: 10.2500/ajr.2008.22.3191 External link
216.
Grobler A, Weitzel EK, Beule A, Jardeleza C, Cheong YC, Field J, Wormald PJ. Pre- and postoperative sinus penetration of nasal irrigation. Laryngoscope. 2008; 118: 2078-2081. DOI: 10.1097/MLG.0b013e31818208c1 External link
217.
Beule A, Athanasiadis T, Athanasiadis E, Field J, Wormald PJ. Efficacy of different techniques of sinonasal irrigation after modified Lothrop procedure. Am J Rhinol Allergy. 2009; 23: 85-90. DOI: 10.2500/ajra.2009.23.3265 External link
218.
Jorissen M. Postoperative care following endoscopic sinus surgery. Rhinology. 2004; 42: 114-120.
219.
Hosemann W, Schindler E, Wiegrebe E, Göpferich A. Innovative frontal sinus stent acting as a local drug-releasing system. Eur Arch Otorhinolaryngol. 2003; 260: 131-134.
220.
Merkus F, Verhoef J, Schipper N, Marttin E. Nasal mucociliary clearance as a factor in nasal drug delivery. Adv Drug Deliv Rev. 1998; 9: 13-38.
221.
Merkle HP, Ditzinger G, Lang SR, Peter H, Schmidt MC. In vitro cell models to study nasal mucosal permeability and metabolism. Adv Drug Deliv Rev. 1998; 29: 51-79. DOI: 10.1016/S0169-409X(97)00061-6 External link
222.
Shimpi S, Chauhan B, Shimpi P. Cyclodextrins: application in different routes of drug administration. Acta Pharm. 2005; 55: 139-156.
223.
Khanvilkar K, Donovan M, Flanagan D. Drug transfer through mucus. Adv Drug Deliv Rev. 2001; 48: 173-193. DOI: 10.1016/S0169-409X(01)00115-6 External link
224.
Osth K, Grasjo J, Bjork E. A new method for drug transport studies on pig nasal mucosa using a horizontal Ussing chamber. J Pharm Sci. 2002; 91: 1259-1273. DOI: 10.1002/jps.10123 External link
225.
Jadhav KR, Shaikh IM, Ambade KW, Kadam VJ. Applications of microemulsion based drug delivery system. Curr Drug Deliv. 2006; 3: 267-273. DOI: 10.2174/156720106777731118 External link
226.
Pavan B, Dalpiaz A, Ciliberti N, Biondi C, Manfredini S, Vertuani S. Progress in drug delivery to the central nervous system by the prodrug approach. Molecules. 2008; 13: 1035-1065. DOI: 10.3390/molecules13051035 External link
227.
MacGregor FB, Robson AG, Pride NB. Topical corticosteroids potentiate mucin secretion in the normal nose. Clin Otolaryngol Allied Sci. 1996; 21: 76-79. DOI: 10.1111/j.1365-2273.1996.tb01029.x External link
228.
Dayal P, Pillay V, Babu RJ, Singh M. Box-Behnken experimental design in the development of a nasal drug delivery system of model drug hydroxyurea: characterization of viscosity, in vitro drug release, droplet size, and dynamic surface tension. AAPS PharmSciTech. 2005; 6: E573-585. DOI: 10.1208/pt060472 External link
229.
Nielsen LS, Schubert L, Hansen J. Bioadhesive drug delivery systems. I. Characterisation of mucoadhesive properties of systems based on glyceryl mono-oleate and glyceryl monolinoleate. Eur J Pharm Sci. 1998; 6: 231-239. DOI: 10.1016/S0928-0987(97)10004-5 External link
230.
Bhat P, Flanagan M. The limitating role of mucus in drug absorption: Drug permeation through mucus solution. Int J Pharm. 1995; 126: 179-187. DOI: 10.1016/0378-5173(95)04120-6 External link
231.
Larhed AW, Artursson P, Grasjo J, Bjork E. Diffusion of drugs in native and purified gastrointestinal mucus. J Pharm Sci. 1997; 86: 660-665. DOI: 10.1021/js960503w External link
232.
Corbo D, Liu J, Chien Y. Characterisation of the barrier properties of mucosal membranes. J Pharm Sci. 1990; 79: 202-206. DOI: 10.1002/jps.2600790304 External link
233.
McMartin C, Hutchinson L, Hyde R, Peters G. Analysis of structural requirements for the absorption of drugs and macromolecules from the nasal cavity. J Pharm Sci. 1987; 76: 535-540. DOI: 10.1002/jps.2600760709 External link
234.
Cornaz A, Buri P. Nasal mucosa as an absorption barrier. Eur J Pharm Biopharm. 1994; 40: 261-270.
235.
Ropke M, Carstens S, Holm M, Frederiksen O. Ion transport mechanisms in native rabbit nasal airway epithelium. Am J Physiol. 1996; 271: L637-645.
236.
Flynn AN, Itani OA, Moninger TO, Welsh MJ. Acute regulation of tight junction ion selectivity in human airway epithelia. Proc Natl Acad Sci U S A. 2009; 106: 3591-3596. DOI: 10.1073/pnas.0813393106 External link
237.
Luppi B, Bigucci F, Mercolini L, Musenga A, Sorrenti M, Catenacci L, Zecchi V. Novel mucoadhesive nasal inserts based on chitosan/hyaluronate polyelectrolyte complexes for peptide and protein delivery. J Pharm Pharmacol. 2009; 61: 151-157. DOI: 10.1211/jpp.61.02.0003 External link
238.
Settipane RA, Lieberman P. Update on nonallergic rhinitis. Ann Allergy Asthma Immunol. 2001; 86: 494-507; quiz 507-498.
239.
Baraniuk JN, Shusterman DJ. Clinical Allergy and Immunology. Nonallergic rhinitis. New York, London: informa Healthcare; 2008.
240.
Graf P. Rhinitis medicamentosa: aspects of pathophysiology and treatment. Allergy. 1997; 52: 28-34. DOI: 10.1111/j.1398-9995.1997.tb04881.x External link
241.
Graf P. Rhinitis medicamentosa: a review of causes and treatment. Treat Respir Med. 2005; 4: 21-29. DOI: 10.2165/00151829-200504010-00003 External link
242.
Fleece L, Mizes JS, Jolly PA, Baldwin RL. Rhinitis medicamentosa. Conceptualization, incidence, and treatment. Ala J Med Sci. 1984; 21: 205-208.
243.
Fox N. Chronic effect of epinephrine and ephedrine on the nasal mucosa. Arch Otolaryngol. 1931; 30: 73-76.
244.
Graf P, Hallen H, Juto JE. The pathophysiology and treatment of rhinitis medicamentosa. Clin Otolaryngol Allied Sci. 1995; 20: 224-229. DOI: 10.1111/j.1365-2273.1995.tb01853.x External link
245.
Graf P, Juto JE. Sustained use of xylometazoline nasal spray shortens the decongestive response and induces rebound swelling. Rhinology. 1995; 33: 14-17.
246.
Baldwin RL. Rhinitis medicamentosa (an approach to treatment). J Med Assoc State Ala. 1975; 47: 33-35.
247.
Lacroix JS. Adrenergic and non-adrenergic mechanisms in sympathetic vascular control of the nasal mucosa. Acta Physiol Scand Suppl. 1989; 581: 1-63.
248.
Elwany SS, Stephanos WM. Rhinitis medicamentosa. An experimental histopathological and histochemical study. ORL J Otorhinolaryngol Relat Spec. 1983; 45: 187-194. DOI: 10.1159/000275642 External link
249.
Talaat M, Belal A, Aziz T, Mandour M, Maher A. Rhinitis medicamentosa: electron microscopic study. J Laryngol Otol. 1981; 95: 125-131. DOI: 10.1017/S0022215100090526 External link
250.
Knipping S, Holzhausen HJ, Goetze G, Riederer A, Bloching MB. Rhinitis medicamentosa: electron microscopic changes of human nasal mucosa. Otolaryngol Head Neck Surg. 2007; 136: 57-61. DOI: 10.1016/j.otohns.2006.08.025 External link
251.
Lin CY, Cheng PH, Fang SY. Mucosal changes in rhinitis medicamentosa. Ann Otol Rhinol Laryngol. 2004; 113: 147-151.
252.
Akerlund A, Bende M. Sustained use of xylometazoline nose drops aggravetes vasomotor rhinitis. Am J Rhinol. 1991; 5: 157-160. DOI: 10.2500/105065891781874983 External link
253.
Graf P, Enerdal J, Hallen H. Ten days' use of oxymetazoline nasal spray with or without benzalkonium chloride in patients with vasomotor rhinitis. Arch Otolaryngol Head Neck Surg. 1999; 125: 1128-1132.
254.
Morris S, Eccles R, Martez SJ, Riker DK, Witek TJ. An evaluation of nasal response following different treatment regimes of oxymetazoline with reference to rebound congestion. Am J Rhinol. 1997; 11: 109-115. DOI: 10.2500/105065897782537197 External link
255.
Yoo JK, Seikaly H, Calhoun KH. Extended use of topical nasal decongestants. Laryngoscope. 1997; 107: 40-43. DOI: 10.1097/00005537-199701000-00010 External link
256.
Hakansson B, Forsgren A, Tegner H, Toremalm NG. Inhibitory effects of nasal drops components on granulocyte chemotaxis. Pharmacol Toxicol. 1989; 64: 321-323. DOI: 10.1111/j.1600-0773.1989.tb00656.x External link
257.
Bjerknes R, Steinsvag SK. Inhibition of human neutrophil actin polymerization, phagocytosis and oxidative burst by components of decongestive nosedrops. Pharmacol Toxicol. 1993; 73: 41-45. DOI: 10.1111/j.1600-0773.1993.tb01955.x External link
258.
Hallen H, Graf P. Benzalkonium chloride in nasal decongestive sprays has a long-lasting adverse effect on the nasal mucosa of healthy volunteers. Clin Exp Allergy. 1995; 25: 401-405. DOI: 10.1111/j.1365-2222.1995.tb01070.x External link
259.
Bernstein IL. Is the use of benzalkonium chloride as a preservative for nasal formulations a safety concern? A cautionary note based on compromised mucociliary transport. J Allergy Clin Immunol. 2000; 105: 39-44. DOI: 10.1016/S0091-6749(00)90175-1 External link
260.
Graf P, Hallen H. Effect on the nasal mucosa of long-term treatment with oxymetazoline, benzalkonium chloride, and placebo nasal sprays. Laryngoscope. 1996; 106: 605-609. DOI: 10.1097/00005537-199605000-00016 External link
261.
Knipping S, Holzhausen HJ, Riederer A, Bloching M. Ultrastrukturelle Veranderungen der Nasenschleimhaut des Menschen nach Abusus von topischen alpha-Sympathomimetika. HNO. 2006; 54: 742-748. DOI: 10.1007/s00106-005-1370-7 External link
262.
Tas A, Yagiz R, Yalcin O, Uzun C, Huseyinova G, Adali MK, Karasalihoglu AR. Use of mometasone furoate aqueous nasal spray in the treatment of rhinitis medicamentosa: an experimental study. Otolaryngol Head Neck Surg. 2005; 132: 608-612. DOI: 10.1016/j.otohns.2005.01.010 External link
263.
Ramey JT, Bailen E, Lockey RF. Rhinitis medicamentosa. J Investig Allergol Clin Immunol. 2006; 16: 148-155.
264.
Keyserling HF, Grimme JD, Camacho DL, Castillo M. Nasal septal perforation secondary to rhinitis medicamentosa. Ear Nose Throat J. 2006; 85: 376, 378-379.
265.
Toohill RJ, Lehman RH, Grossman TW, Belson TP. Rhinitis medicamentosa. Laryngoscope. 1981; 91: 1614-1621. DOI: 10.1288/00005537-198110000-00005 External link
266.
Passali D, Salerni L, Passali GC, Passali FM, Bellussi L. Nasal decongestants in the treatment of chronic nasal obstruction: efficacy and safety of use. Expert Opin Drug Saf. 2006; 5: 783-790. DOI: 10.1517/14740338.5.6.783 External link
267.
Graf P. Long-term use of oxy- and xylometazoline nasal sprays induces rebound swelling, tolerance, and nasal hyperreactivity. Rhinology. 1996; 34: 9-13.
268.
Ferguson BJ, Paramaesvaran S, Rubinstein E. A study of the effect of nasal steroid sprays in perennial allergic rhinitis patients with rhinitis medicamentosa. Otolaryngol Head Neck Surg. 2001; 125: 253-260.
269.
Elwany S, Abdel-Salaam S. Treatment of rhinitis medicamentosa with fluticasone propionate--an experimental study. Eur Arch Otorhinolaryngol. 2001; 258: 116-119. DOI: 10.1007/s004050000309 External link
270.
Caffier PP, Frieler K, Scherer H, Sedlmaier B, Goktas O. Rhinitis medicamentosa: therapeutic effect of diode laser inferior turbinate reduction on nasal obstruction and decongestant abuse. Am J Rhinol. 2008; 22: 433-439. DOI: 10.2500/ajr.2008.22.3199 External link
271.
Milosevic D, Janosevic L, Dergenc R, Vasic M. Pathologic conditions associated with drug-induced rhinitis. Srp Arh Celok Lek. 2004; 132: 14-17. DOI: 10.2298/SARH0402014M External link
272.
Graf PM, Hallen H. One year follow-up of patients with rhinitis medicamentosa after vasoconstrictor withdrawal. Am J Rhinol. 1997; 11: 67-72. DOI: 10.2500/105065897781446865 External link
273.
Moore EJ, Kern EB. Atrophic rhinitis: a review of 242 cases. Am J Rhinol. 2001; 15: 355-361.
274.
Bunnag C, Jareoncharsri P, Tansuriyawong P, Bhothisuwan W, Chantarakul N. Characteristics of atrophic rhinitis in Thai patients at the Siriraj Hospital. Rhinology. 1999; 37: 125-130.
275.
Sibert JR, Barton RP. Dominant inheritance in a family with primary atrophic rhinitis. J Med Genet. 1980; 17: 39-40. DOI: 10.1136/jmg.17.1.39 External link
276.
Mehrotra R, Singhal J, Kawatra M, Gupta SC, Singh M. Pre and post-treatment histopathological changes in atrophic rhinitis. Indian J Pathol Microbiol. 2005; 48: 310-313.
277.
Jiang HL, Park IK, Shin NR, Kang SG, Yoo HS, Kim SI, Suh SB, Akaike T, Cho CS. In vitro study of the immune stimulating activity of an atrophic rhinitis vaccine associated to chitosan microspheres. Eur J Pharm Biopharm. 2004; 58: 471-476. DOI: 10.1016/j.ejpb.2004.05.006 External link
278.
Effat KG, Madany NM. Microbiological study of role of fungi in primary atrophic rhinitis. J Laryngol Otol. 2009; 123: 631-634. DOI: 10.1017/S002221510800399X External link
279.
Garcia GJ, Bailie N, Martins DA, Kimbell JS. Atrophic rhinitis: a CFD study of air conditioning in the nasal cavity. J Appl Physiol. 2007; 103: 1082-1092. DOI: 10.1152/japplphysiol.01118.2006 External link
280.
Singh M, Jain V, Singh SP, Gupta SC. Endoscopic dacryocystorhinostomy in cases of dacryocystitis due to atrophic rhinitis. J Laryngol Otol. 2004; 118: 426-428. DOI: 10.1258/002221504323219536 External link
281.
Yanagisawa E, Scher DA. Endoscopic views of choanal stenosis in secondary atrophic rhinitis. Ear Nose Throat J. 2003; 82: 666-668.
282.
Dutt SN, Kameswaran M. The aetiology and management of atrophic rhinitis. J Laryngol Otol. 2005; 119: 843-852. DOI: 10.1258/002221505774783377 External link
283.
Ishikawa S, Nakayama T, Watanabe M, Matsuzawa T. Visualization of flow resistance in physiological nasal respiration: analysis of velocity and vorticities using numerical simulation. Arch Otolaryngol Head Neck Surg. 2006; 132: 1203-1209. DOI: 10.1001/archotol.132.11.1203 External link
284.
Hirt R, Paulsen F, Neumann K, Knipping S. Immunhistochemischer Nachweis von Caspase 3 bei verschiedenen Nasenschleimhauterkrankungen des Menschen. HNO. 2009; 57: 466-472. DOI: 10.1007/s00106-009-1905-4 External link
285.
Sayed RH, Abou-Elhamd KE, Abdel-Kader M, Saleem TH. Study of surfactant level in cases of primary atrophic rhinitis. J Laryngol Otol. 2000; 114: 254-259. DOI: 10.1258/0022215001905508 External link
286.
Pruliere-Escabasse V, Escudier E, Balheda R, Soria JC, Coste A, Massard C. Rhinitis and epistaxis in patients treated by anti-angiogenic therapy. Invest New Drugs. 2009; 27: 285-286. DOI: 10.1007/s10637-008-9168-6 External link
287.
Ly TH, de Shazo RD, Olivier J, Stringer SP, Daley W, Stodard CM. Diagnostic criteria for atrophic rhinosinusitis. Am J Med. 2009; 122: 747-753. DOI: 10.1016/j.amjmed.2008.12.025 External link
288.
Wachsberger A. A new technic in the surgical treatment of ozena. Arch Otolaryngol. 1934; 19: 370-382.
289.
Boyce J, Eccles R. Do chronic changes in nasal airflow have any physiological or pathological effect on the nose and paranasal sinuses? A systematic review. Clin Otolaryngol. 2006; 31: 15-19. DOI: 10.1111/j.1749-4486.2006.01125.x External link
290.
Eccles R. Nasal airflow in health and disease. Acta Otolaryngol. 2000; 120: 580-595. DOI: 10.1080/000164800750000388 External link
291.
Dykewicz MS, Fineman S, Skoner DP, Nicklas R, Lee R, Blessing-Moore J, Li JT, Bernstein IL, Berger W, Spector S, Schuller D. Diagnosis and management of rhinitis: complete guidelines of the Joint Task Force on Practice Parameters in Allergy, Asthma and Immunology. American Academy of Allergy, Asthma, and Immunology. Ann Allergy Asthma Immunol. 1998; 81: 478-518. DOI: 10.1016/S1081-1206(10)63155-9 External link
292.
Neher A, Gstottner M, Thaurer M, Augustijns P, Reinelt M, Schobersberger W. Influence of essential and fatty oils on ciliary beat frequency of human nasal epithelial cells. Am J Rhinol. 2008; 22: 130-134. DOI: 10.2500/ajr.2008.22.3137 External link
293.
Zenner H. Praktische Therapie von Hals-Nasen-Ohren-Krankheiten. Stuttgart, New York: Schattauer; 2008. pp. 494.
294.
Jaswal A, Jana AK, Sikder B, Nandi TK, Sadhukhan SK, Das A. Novel treatment of atrophic rhinitis: early results. Eur Arch Otorhinolaryngol. 2008; 265: 1211-1217. DOI: 10.1007/s00405-008-0629-5 External link
295.
Dykewicz MS, Fineman S. Executive Summary of Joint Task Force Practice Parameters on Diagnosis and Management of Rhinitis. Ann Allergy Asthma Immunol. 1998; 81: 463-468. DOI: 10.1016/S1081-1206(10)63152-3 External link
296.
Jiang RS, Hsu CY, Chen CC, Jan YJ, Jang JW. Endoscopic sinus surgery and postoperative intravenous aminoglycosides in the treatment of atrophic rhinitis. Am J Rhinol. 1998; 12: 325-333. DOI: 10.2500/105065898780182480 External link
297.
Lobo CJ, Hartley C, Farrington WT. Closure of the nasal vestibule in atrophic rhinitis--a new non-surgical technique. J Laryngol Otol. 1998; 112: 543-546. DOI: 10.1017/S0022215100141040 External link
298.
Lal D, Corey JP. Vasomotor rhinitis update. Curr Opin Otolaryngol Head Neck Surg. 2004; 12: 243-247. DOI: 10.1097/01.moo.0000122310.13359.79 External link
299.
Silvers WS. The skier's nose: a model of cold-induced rhinorrhea. Ann Allergy. 1991; 67: 32-36.
300.
Giannessi F, Ursino F, Fattori B, Giambelluca MA, Scavuzzo MC, Ceccarelli F, Ruffoli R. Immunohistochemical localization of 3-nitrotyrosine in the nasal respiratory mucosa of patients with vasomotor rhinitis. Acta Otolaryngol. 2005; 125: 65-71. DOI: 10.1080/00016480410016982 External link
301.
Papon JF, Brugel-Ribere L, Fodil R, Croce C, Larger C, Rugina M, Coste A, Isabey D, Zerah-Lancner F, Louis B. Nasal wall compliance in vasomotor rhinitis. J Appl Physiol. 2006; 100: 107-111. DOI: 10.1152/japplphysiol.00575.2005 External link
302.
Vayisoglu Y, Ozcan C, Pekdemir H, Gorur K, Pata YS, Camsari A. Autonomic nervous system evaluation using heart rate variability parameters in vasomotor rhinitis patients. J Otolaryngol. 2006; 35: 338-342. DOI: 10.2310/7070.2006.0080 External link
303.
Brandt D, Bernstein JA. Questionnaire evaluation and risk factor identification for nonallergic vasomotor rhinitis. Ann Allergy Asthma Immunol. 2006; 96: 526-532. DOI: 10.1016/S1081-1206(10)63546-6 External link
304.
Lieberman P, Kaliner MA, Wheeler WJ. Open-label evaluation of azelastine nasal spray in patients with seasonal allergic rhinitis and nonallergic vasomotor rhinitis. Curr Med Res Opin. 2005; 21: 611-618. DOI: 10.1185/030079905X41408 External link
305.
Arikan OK, Koc C, Kendi T, Muluk NB, Ekici A. CT assessment of the effect of fluticasone propionate aqueous nasal spray treatment on lower turbinate hypertrophy due to vasomotor rhinitis. Acta Otolaryngol. 2006; 126: 37-42. DOI: 10.1080/00016480510012219 External link
306.
Jacobs R, Lieberman P, Kent E, Silvey M, Locantore N, Philpot EE. Weather/temperature-sensitive vasomotor rhinitis may be refractory to intranasal corticosteroid treatment. Allergy Asthma Proc. 2009; 30: 120-127. DOI: 10.2500/aap.2009.30.3206 External link
307.
Ozcan C, Vayisoglu Y, Dogu O, Gorur K. The effect of intranasal injection of botulinum toxin A on the symptoms of vasomotor rhinitis. Am J Otolaryngol. 2006; 27: 314-318. DOI: 10.1016/j.amjoto.2006.01.008 External link
308.
Fleckenstein J, Raab C, Gleditsch J, Ostertag P, Rasp G, Stor W, Irnich D. Impact of acupuncture on vasomotor rhinitis: a randomized placebo-controlled pilot study. J Altern Complement Med. 2009; 15: 391-398. DOI: 10.1089/acm.2008.0471 External link
309.
Araujo E, Palombini BC, Cantarelli V, Pereira A, Mariante A. Microbiology of middle meatus in chronic rhinosinusitis. Am J Rhinol. 2003; 17: 9-15.
310.
Gordts F, Halewyck S, Pierard D, Kaufman L, Clement P. Microbiology of the middle meatus: a comparison between normal adults and children. J Laryngol Otol. 2000; 114: 184-188. DOI: 10.1258/0022215001905292 External link
311.
Gwaltney J. Microbiology of sinusitis. In: Druce H, eds. Sinusitis - Pathophysiology and treatment. New York: Dekker; 1994.
312.
Kirtsreesakul V, Tuntaraworasin J, Thamjarungwong B. Microbiology and antimicrobial susceptibility patterns of commensal flora in the middle nasal meatus. Ann Otol Rhinol Laryngol. 2008; 117: 914-918.
313.
Klossek J, Dubreuil L, Richet H, Richet B, Sedallian A, Beutter P. Bacteriology of the adult middle meatus. J Laryngol Otol. 1996; 110: 847-849. DOI: 10.1017/S0022215100135133 External link