Ideally, CFD simulation results can be validated against well-controlled experiments where uncertainties regarding e.g. the flow geometry and mass and heat transfer have been minimized, and measurement errors are under control. Under such conditions, the CFD model is subject to little uncertainty, and errors associated with poor choice of modelling strategies and submodels can readily be assessed, so that the model can be tuned to predict measured data with good accuracy. This might be the reason for the good agreement reported between in vitro and in silico RMM.
Due to difficulties and challenges associated with acquiring and assessing the quality and reproducibility of objective in vivo clinical data, including RMM and CT/MRI imaging data, there is substantial uncertainty related to the quantitative comparison of nasal resistance and pressure-flow curves obtained from in vivo and in silico RMM. E.g., an essential part of in silico RMM, not inherent in standard in vivo RMM, is the requirement of a detailed description of the nasal cavity geometry. In silico RMM typically utilize medical imaging data to acquire the nasal cavity geometry, but unless specific actions and precautions are taken, it is unknown to what extent the medical images adequately describe the state of the nasal cavity during in vivo RMM.
Computed flow variables strongly depend on the flow geometry. Thus, the lack of knowledge about the instantaneous state of the nasal cavity during in vivo RMM causes major uncertainty regarding the quantitative comparison of nasal resistance and other flow parameters obtained from in vivo and in silico RMM. A discussion of uncertainties and errors in the comparison of in vivo and in silico RMM is therefore incomplete without a separate discussion of the relevance of medical imaging data with respect to describing the nasal cavity at the time of in vivo RMM.
I.6.1 Physiological factors affecting the temporal variability of the nasal cavity geometry
To attain accurate predictions of in vivo RMM results through in silico RMM simulations, it is imperative to employ a digital nasal cavity model that faithfully represents the dynamic state of the nasal cavity during the in vivo RMM examination.
The alignment reported between
in silico RMM and
in vitro RMM in physical replicas of nasal cavities [
127,
129,
130] starkly contrasts the observed disagreement with
in vivo RMM [
117,
125]. This suggests that CFD models are correctly configured, but somehow fail to adequately represent the nasal cavity's actual state and function during
in vivo RMM examination.
Two physiological mechanisms that can cause digital geometry models to misrepresent the nasal cavity geometry during in vivo RMM, are: 1) the nasal cycle, known to cause a periodic, temporal variation in nasal cavity volume, and 2) nasal cavity compliance, which may cause spontaneous expansion/contraction of the nasal cavity volume due to over/under pressure during respiration. In the following subsections, brief discussions are given about these two physiological phenomena. Other causes for errors and uncertainty associated with the acquisition of a digital nasal cavity geometry are discussed below.
Nasal cycle
The nasal cycle causes spontaneous engorgement, hence cross-sectional variability, in the nasal cavities. Consequently, the nasal cycle affects the reproducibility of objective rhinometric measurements adversely and complicates the objective assessment of nasal patency when comparing pre- and post-operative measurements. Additionally, it poses challenges when comparing in vivo and in silico RMM results, as the nasal cavity's shape and volume may differ between in vivo RMM examination and the acquisition of medical imaging data used for in silico RMM.
In a study by Hasegawa and Kern [
74], which involved 50 subjects, bilateral and unilateral nasal resistance measurements were conducted over a 6-7 hour period. They observed that the bilateral nasal resistance remained relatively constant despite cyclic variations in unilateral resistances. The average ratio of highest to lowest unilateral resistances was 4.6 on the right side and 4.4 on the left side, with peak values reaching 16.3 and 13.7, respectively. Hence, accounting for the nasal cycle is crucial when assessing nasal resistance. It is noteworthy that the nasal cycle was found to be non-reproducible in all of the five subjects who underwent re-testing, as the durations and amplitudes of their nasal cycles varied. To quantify the nature of the nasal cycle, in numerical terms, Flanagan and Eccles [
75] conducted hourly unilateral airflow measurements over 8-hour periods in 52 subjects.
Patel
et al. [
131] employed CFD to investigate the impact of the nasal cycle on objective measures. They suggested that paradoxical post-operative worsening of NAO observed in simulations could be attributed to the nasal cycle. Gaberino
et al. [
65] created virtual mid-cycle models to correct for the nasal cycle, resulting in improved correlation between objective and subjective measures of nasal patency. Moghaddam
et al. [
90] pointed out that mucosal engorgement due to the nasal cycle can significantly affect CT images, thereby influencing the correlation between CFD and subjective and objective nasal patency scores. Susaman
et al. [
132] emphasized that rhinologists need to take the existence of the nasal cycle, which affects a large percentage of the population, into account when examining and measuring the nose.
Several authors [
33,
133,
134] have pointed out that postural effects on the nasal resistance and the nasal cycle should be expected. The nasal resistance tends to be higher in the supine position. Consequently, differences in posture between medical imaging procedures and RMM examinations may lead to geometrical misrepresentation in
in silico RMM simulations.
To mitigate the effects of the nasal cycle, it is common practice to perform RMM both before and after applying a decongestive nasal spray which shrinks the large veins in the nasal epithelium. This approach enables the evaluation of the
anatomical nasal patency [
45].
The presence of the nasal cycle suggests that if CT/MRI images and RMM measurements are not acquired within a short timeframe, in the same state of decongestion, and in the same posture, they may not reflect the same nasal geometries.
Nasal cavity compliance
In a study conducted by Fodil
et al. [
135], a simplified model of the nasal cavity was used to demonstrate that, depending on the pathological condition, the assumption of rigid nasal cavity walls is only valid for low flowrates. The rigid wall assumption generally failed for pressure drops above
20 Pa. In contradiction, Bailie
et al. [
136] claimed that during resting breathing, one can regard the nasal cavity as a rigid structure. More recent research by Akmenkalne
et al. [
137] corroborated the earlier work of Fodil
et al. [
135]. They investigated the mobility of the lateral nasal wall under the influence of breathing and emphasized that even during quiet breathing, we must take into account the deflection of the nasal walls. O'Neill and Tolley [
72] used a simplified mathematical model based on Bernoulli's principle to compute the total pressure loss through the nasal cavity as a sum of minor losses. Their model allowed the nasal gateway (valve) to dynamically adjust its cross-sectional area based on local pressure and a stiffness coefficient, providing a quantitative rationale for observed discrepancies between AR and RMM. Cherobin
et al. [
117] observed that while
in silico RMM was in good agreement with
in vitro RMM, it diverged significantly from
in vivo RMM. This disparity was partly attributed to the rigid wall assumption in CFD, which matched the properties of the rigid nasal cavity replica used in
in vitro experiments but might have failed to adequately represent physiological nasal cavity compliance. Schmidt
et al. [
121] reported systematic underprediction of nasal resistance in CFD simulations but found no significant difference between patients with or without nasal valve collapse or between inhalation and exhalation phases.
Considering that nasal cavity expansion/contraction in response to over-/under-pressure during exhalation/inhalation, has an effect on the nasal resistance, it is anticipated that the pressure-flow curves will exhibit asymmetry between these respiratory phases. To assess this, one can compare mirrored exhalation curves with inhalation curves obtained from in vivo RMM. If the two sets of curves align well, it suggests that nasal cavity compliance is minimal and cannot account for the substantial differences between in silico and in vivo RMM results. However, it is important to note that this argument does not consider the Venturi effect. When the Venturi effect dominates over the hydrostatic effect, it can lead to contraction during both inhalation and exhalation, resulting in an overall increase in nasal resistance.
This line of reasoning is consistent with the observations of Akmenkalne
et al. [
137], who demonstrated contraction during both inhalation and exhalation in quiet breathing. For elevated breathing and forced sniffing, the hydrostatic pressure component appeared to dominate, causing contraction during inhalation and expansion during exhalation. An interesting observation in their
Figure 4 was that after a period of forceful sniffing, the deflection of the lateral nasal wall reversed its direction, transitioning from negative to positive but trending downwards. However, this reversal and slow nasal wall relaxation time did not appear to correlate with flow or pressure curves, implying minimal impact on nasal resistance.
The influence of nasal compliance on the hysteresis in RMM pressure-flow curves illustrated in Section I.3.3, has been discussed by Vogt and co-authors [
77,
138]. Wernecke
et al. [
77], Vogt and Zhang [
138], Vogt
et al. [
139], Bozdemir
et al. [
76], and Frank-Ito and Garcia [
115] presented pressure-flow curves featuring hysteresis where the portions of the curve corresponding to the accelerating and decelerating inspiratory phases were switched when compared to the simplified model showcased in
Figure 2b of the present paper. Measurement results by Groß and Peters [
140] support the time arrows in
Figure 2b, but they attributed the observed pressure-flow curve hysteresis to the measurement technique, rather than nasal airflow dynamics. An adequate explanation for this disagreement is lacking.
It is anticipated that the influence of nasal compliance on RMM pressure-flow curves will manifest as asymmetry between inspiratory and expiratory pressure-flow curves as well as, referring to
Figure 2b, a widening of the hysteresis loop. It may, however, be imagined a situation where the Venturi effect dominates over the static pressure such that local under pressure is effectively independent of flow direction and causes (partial) collapse both during inhalation and exhalation. Owing to the phase disparity between volumetric flow rate and flow resistance, brief periods of counterintuitive over and under pressure may occur within the nasal cavity at the culmination of the inspiratory and expiratory phases, respectively (see
Figure 2). Consequential local expansion/contraction may introduce complexity to the response of the pressure-flow curves.
I.6.4 Sources of uncertainties and errors in in silico rhinomanometry (CFD)
When presented with experimental data from flow measurements, such as
in vivo or
in vitro RMM, it rests upon the CFD engineer to set up a CFD model that is able to reproduce the measured data, or to explain observed discrepancies between computed and measured data. A significant preparatory task in CFD modelling is to describe the flow system both qualitatively and quantitatively with respect to geometry (including flow restrictions/walls, inlets, and exits), material properties, and other factors that may affect the flow. Even if the geometry of the flow system is well known, there is a multitude of parameters and settings that must be chosen carefully when setting up the CFD model. Potential sources of error associated with setting up and running CFD simulations have been thoroughly covered by the European Research Community on Flow, Turbulence and Combustion [
146] and many others, e.g. Andersson
et al. [
147], Rodriguez [
94], and Roychowdhury [
95]. Inthavong
et al. [
105] reviewed
in silico approaches to simulation of nasal airflow.
Besides fundamental errors and limitations in the program code of the CFD software, such as program bugs or truncation- and round-off errors, errors in simulation results can be caused by e.g.:
Poor computational mesh quality [
94].
Inadequate spatial or temporal refinement.
Poorly selected solver settings and numerical schemes [
148].
Incorrect definition of flow physics, including e.g. boundary conditions, material properties, and approximations.
Inaccurate or incorrect solution due to poor convergence and/or failure to conserve mass, momentum, or energy.
The main uncertainties are related to the lack of knowledge about the flow problem to be modelled. These can be divided into four main categories related to 1) flow physics; 2) geometry; 3) required spatial and temporal numerical resolution; and 4) boundary conditions. Even if these are implemented correctly without errors, there may be uncertainty associated with their correct description. For simulation parameters associated with high uncertainty, sensitivity analysis may be required to assess the influence of variations in these parameters.
Flow physics
Uncertainties surrounding the flow physics within the nasal cavity encompass aspects that, theoretically, could be elucidated through measurements or experiments. However, practical challenges arise in conducting in vivo measurements on patients, and a lack of in vitro experimental data complicates the matter. Consequently, an ongoing debate persists regarding fundamental aspects of nasal flow physics. This includes the deciding between quasi-steady and transient modeling, determining the optimal turbulence modeling strategy, and addressing other considerations such as the dependence of air's material and transport properties on pressure, temperature, and humidity.
Modelling of temporal phenomena in respiratory flow
The physiological, respiratory flow in the nasal cavity is normally of pulsative nature. The literature review summarized in
Table 1 suggests that steady flow modelling, by far, is the most popular approach in computational rhinology, however. It is appropriate to question the validity of the assumption of quasi-steady flow in respiratory flow modelling. E.g., how does the transient nature of the flow affect temporal effects such as hysteresis, developing flow boundary layers, and meandering of wakes or jets?
The simulation of transient flow adds complexity to CFD simulations compared to modelling steady state flow. Many authors have argued that nasal airflow can be approximated by quasi-steady flow [
149,
150,
151,
152,
153,
154].
In the current context, the concept of quasi-steady state implies that the time-response of the overall flow phenomena within a system is much quicker than the variation of transient phenomena occurring in the system. The system's behavior can thus be assumed to be in instantaneous equilibrium with the transient phenomena, enabling its approximation by steady state simulations. In the case of nasal airflow, quasi-steady state suggests that the nasal flow parameters can be determined from the instantaneous respiratory pressure and velocity boundary conditions, at any given moment. This implies that pressure-flow curves in in silico RMM can be generated through a series of steady state simulations conducted at different volumetric flowrates, instead of relying on a transient simulation of the entire breathing cycle.
The Womersley number, named after J. R. Womersley [
155] who studied pulsatile flow in arteries, is defined as the ratio between the transient inertial and viscous forces and is commonly expressed as
where
is the channel diameter,
is the pulsation frequency, and
is the kinematic viscosity. The Womersley number can be used to characterize an unsteady flow as quasi-steady or not [
156]. The flow may be considered quasi-steady if
. Inserting for
,
, and
, the expected Womersley number in unilateral nasal airflow is approximately
. This is in the intermediate range, where the oscillatory nature of the flow is not dominating but may have some influence.
Doorly
et al. [
157] discussed whether a series of quasi-steady simulations is sufficient to characterize tidal breathing. They referred to Shi
et al. [
153] and suggested that the quasi-steady assumption is valid for quiet breathing. Bosykh
et al. [
158] showed that a transient model produced almost identical results as steady state simulations produced by themselves as well as others. Furthermore, they observed that asymmetry of the respiratory cycle had little effect on the flow pattern in the nasal cavity, compared to a sinusoidal inhalation/exhalation profile, which follows naturally from quasi-steady behavior. Bradshaw
et al. [
159] highlighted several phenomena observed in their transient simulations that cannot be seen in steady flow. In particular, their results indicate that transient simulations of the entire breathing cycle are essential in order to correctly capture air conditioning via heating/cooling and humidification.
A noteworthy characteristic of
in vivo RMM pressure-flow curves is the presence of a hysteresis pattern [
77]. This hysteresis has been attributed, among other factors, to unsteady/inertial pressure drop contribution stemming from varying flowrates during respiration, and it can naturally not be predicted by steady state flow simulations. See Section I.3.3 for more details.
Modelling of turbulent, transitional, and laminar flow
The complex, dramatically varying flow channel cross-sections in the nasal cavity, can have significant impact on the development of turbulent structures within the flow due to e.g., flow separation, recirculation, varying pressure gradients, secondary flows, developing wall boundary layers, merging of separate flow streams, flow instabilities, etc. The understanding and prediction of turbulence in such scenarios typically requires very detailed CFD models and experiments. It can be expected that the behavior of such flow systems are highly non-linear and three-dimensional. E.g., Tretiakow
et al. [
160] found that the flow in the ostiomeatal complex (e.g. degree of turbulence) depended on the overall geometric features of the nasal cavity (e.g. nasal septum deviation).
Only DNS is an exact representation of the Navier-Stokes equations. All other turbulence models contain approximations with individual limitations and ranges of validity. E.g., while RANS models are ensemble averaged and unable to model individual turbulent eddies, LES models are able to track eddies larger than a given filter size (typically a function of the computational mesh size) and employ subgrid models to describe the effect of smaller eddies. In principle, LES should approach DNS in the limit of small filter sizes.
While RANS based models are much cheaper than LES or DNS based models, in terms of computational power requirements, they are known to have many limitations. These models were typically created to solve specialized industrial problems with good balance between accuracy and computational cost. Model parameters were thus tuned to predict standard, industrial flow scenarios. It is not given that these models are suitable for modelling of flow in complex geometries such as the nasal cavity. Moreover, these models are known to have severe limitations with respect to modelling transitional flow. Thus, if the nasal flow is transitioning between laminar and turbulent flow along the length of the nasal cavity and due to the respiratory variation of the flow velocity, these models might not be able to predict the flow accurately.
Most authors discussing turbulence in the upper airways seem to consider Reynolds number only, as a criterium for the onset of turbulence. They fail to consider that it takes time to develop turbulence. In pipe-flow at Reynolds numbers above the critical Reynolds number, it generally takes more than 10 pipe diameters flow-length to fully develop the turbulent velocity profile. During restful tidal breathing, approximately 25 nasal volumes are inhaled/exhaled during one cycle [
161], and considering that the length of the nasal passage is between five and fifteen times its hydraulic diameter, the flow field is thus unlikely to be fully developed. Even for steady flow, it seems unlikely that the flow can be fully developed due to the varying cross-sectional area along the nasal cavity, and the many anatomical features that affect the flow pattern. There may, however, be regions within the nasal cavity which experience periods of transitional/turbulent flow during a respiratory cycle. It can be expected that most ensemble averaged turbulence models are unsuited for such complex, spatially and temporally varying laminar/transitional/turbulent flow fields.
For flow channels with cross-sections that slightly deviate from a cylindrical shape, the hydraulic diameter has proven useful in predicting flow resistance using standard friction loss correlations, such as the Haaland correlation (Equation (7)). For cross-sectional shapes deviating from cylindrical shapes, inaccuracies have been reported [
73]. This indicates that Reynolds numbers based on hydraulic diameter of complex cross-sections such as those in the nasal cavity, may not be appropriate for predicting the transition from laminar to turbulent flow.
Transition between laminar and turbulent unsteady flow is still, despite its importance in many engineering applications, not fully understood [
162]. Recently, Guerrero
et al. reviewed the literature and performed direct numerical simulation (DNS) to investigate the transient behavior of accelerating [
163] and decelerating [
164] turbulent pipe-flows. An early discussion of the transition between laminar and turbulent flow in pulsatile flow in the cardiovascular system was published by Yellin [
165], who observed that large instantaneous Reynolds numbers did not result in transition to turbulence everywhere. Gündoğdu and Çarpinlioğlu [
166,
167] presented the theoretical background for pulsatile laminar, transitional, and turbulent flows, and reviewed theoretical and experimental investigations. Xu
et al. [
168] investigated the effect of pulsation on transition from laminar to turbulent flow for rigid, straight pipes. They observed that the delay of the transition to turbulence increases with decreasing Womersley number (
) and increasing pulsation amplitude. The implication is that the turbulent transition threshold for Womersley numbers close to 1 is above Reynolds number
3000, in straight pipes, but turbulent puffs may exist due to the high Reynolds number time intervals of a respiratory cycle.
The combined effect of delayed transition and relatively short flow channel suggests that fully developed turbulent flow within the nasal cavity is improbable during resting respiratory flow. When utilizing RANS turbulence models that assume fully developed turbulent flow in cases where the flow is laminar, transitional, or developing, there is a risk of overestimating turbulent viscosity and, consequently, overall flow resistance. If the validation of in silico models relies solely on nasal resistance measured through in vivo RMM, there is a potential bias towards models that overestimate turbulent viscosity. This bias may help align in silico and in vivo RMM pressure-flow curves but could lead to an incomplete representation of other pertinent phenomena. This example illustrates the perils of overly simplistic analyses in the study of complex problems like nasal airflow and emphasizes the necessity for additional objective, measurable metrics in the assessment of nasal airflow.
Schillaci and Quadrio [
148] compared laminar/RANS/LES simulations and concluded that the choice of numerical scheme is more important than the choice of turbulence model, although they emphasize that the chosen turbulence model should be able to handle three-dimensional, vortical, mostly laminar flow conditions. They suggested that LES or DNS is necessary to reliably simulate the full breathing cycle at intermediate intensity. Bradshaw
et al. [
159] performed hybrid RANS-LES simulations of the entire respiratory cycle and reported bilateral nasal airflow to be dominantly laminar. While LES is widely acknowledged as one of the most accurate turbulence modelling approaches, second only to DNS, it is worth noting that LES also encounters challenges in predicting transitional flow and the initiation of turbulence, as highlighted by Sayadi and Moin [
169].
Other aspects
Other aspects of minor importance are just mentioned briefly, for completeness.
Temperature and humidity may affect the material properties of air. Some authors have suggested that these effects should be taken into account [
76]. Other authors have dismissed these effects [
151]. In the relevant temperature range, the mass density and viscosity of air varies with less than ten percent, and the effect of humidity is of the same order. For most situations, it is thus expected that this is of minor importance.
Due to the small effect of pressure on air material properties within the relevant pressure range, and low flow velocities, it is safe to assume atmospheric ambient pressure and constant air material properties.
Geometry
The nasal cavity is a highly complex flow-channel, with cross-sections that change dramatically in shape and area throughout the nose, and generally deviate significantly from cylindrical shape. Moreover, the nasal cavity is bounded by walls covered in mucosal lining, which may add a transient geometrical variation to the air-tissue interface, as well as constituting a non-rigid structure. The acquisition of a 3-dimensional, digital nasal cavity geometry model is prone to errors and uncertainties, as discussed in Section I.5.2. In addition, there are uncertainties regarding the geometrical level of detail required to set up accurate CFD models. For the CFD model to be able to accurately predict the behavior of physical phenomena, it is essential that the geometry model does not misrepresent the actual flow geometry too much. When manufacturing the nasal cavity model, essential questions that should be considered include:
How much of the surrounding volume outside the nose and in the oropharyngeal tract should be included? This consideration will affect to what extent the boundary conditions will affect the simulated flow fields. This question is closely intertwined with the discussion about boundary conditions, below.
How much of the paranasal sinuses should be included? CFD simulation of the flow in the maxillary sinus was performed by Zang
et al. [
170]. Their conclusion was that the airflow inside the maxillary sinus was much lower (<5%) of the airflow in the nasal cavity. Due to the narrow passage connecting the paranasal sinuses with the nasal cavity, it is expected that negligible gas exchange takes place between the two [
171]. This was supported by simulation results presented by Bradshaw
et al. [
159]. Kaneda
et al. [
126] reported that the inclusion of the paranasal sinuses did not improve the disagreement between computed and measured nasal resistances.
What is the role of the oral cavity? Paz et al. [
172] investigated the distribution between nasal and oral breathing under steady and unsteady flow. Chen et al. [
173] concluded that the inclusion of the oral cavity in CFD simulations of steady and unsteady nasal cavity flow had very little impact. Open mouth and oral breathing may, however, affect the RMM pressure-flow curve.
-
Should minor geometrical features such as nasal hair or the mucosal lining be considered?
- ○
Hahn
et al. [
151] found that the inclusion of nasal hair increased turbulent intensity in the external nares during inspiratory flow, but had little effect on downstream velocity profiles. Stoddard
et al. [
174] found that reduction of nasal hair density had positive impact on both subjective and objective measures of nasal obstruction, however.
- ○
Lee
et al. [
175] illustrated how the mucous layer may affect local flow velocities in the nasal cavity.
Uncertainty associated with the geometrical (mis-)representation of the nasal cavity in CFD models may be due to unknown factors affecting the process of manufacturing three-dimensional geometries from medical imaging data, via segmentation, or uncertainty regarding how well the CT/MRI imaging data represents the actual nasal cavity geometry during the RMM procedure.
Required spatial and temporal numerical resolution
Spatial and temporal discretization is required in order to enable the numerical solution of the governing equations of CFD (eqs. (9)-(11)). The rule of thumb is that the spatial and temporal resolution must be sufficient to resolve all spatial and temporal flow features of interest. In addition to determining the accuracy of CFD simulations, mesh and time step size may affect numerical stability and robustness of CFD solvers.
To assess the numerical solution's sensitivity to grid refinement, grid dependency tests should be performed. If the computed flow fields change negligibly by increasing the grid resolution, it is commonly assumed that grid independence is achieved. Frank-Ito
et al. [
176] reviewed the literature and investigated the requirements for grid-independence in their own steady, inspiratory, laminar sinonasal cavity airflow with particle deposition. They emphasized the importance of mesh refinement analysis to obtain trustworthy computational solutions. Similarly, to assess transient solutions' sensitivity to time step size, comparison between simulations employing relatively short and long time steps should be performed.
Brief discussions of how the spatial and temporal resolution can introduce uncertainties in CFD simulations of nasal airflow is given below.
Spatial Resolution
The spatial resolution of the computational mesh determines the level of detail in the computed flow fields. E.g., in the presence of steep velocity gradients, refined meshes are needed to avoid numerical diffusion. Moreover, turbulent eddies smaller than the mesh size must be modelled by closure laws and subgrid models, which introduce approximations.
Particularly, to describe flow profiles accurately in the vicinity of walls, the near-wall mesh must honor requirements by the turbulence model employed. The theory behind this is described in classical text books on turbulence by e.g. e.g. Tennekes and Lumley [
108], Pope [
109], or Wilcox [
110] and in CFD simulation software user and theory guides.
Distance to the wall is commonly expressed in
wall units, where the
dimensionless wall distance is expressed as
where
is the
shear velocity,
is the
wall shear stress, and
is the
distance to the wall. The
Law of the wall states that the
dimensionless velocity parallel to the wall is given by
where
denotes the flow velocity parallel to the wall. The intermediate range between the
viscous sublayer and the
log layer is known as the
buffer layer. The original idea by Launder and Spalding [
177] was to use Equation (14) as a
wall function for the velocity boundary conditions at the wall. This approach, which is used in some classic RANS turbulence models such, as the
type turbulence models, requires that the centroids of computational grid cells residing at the wall are in the log layer (
). Other RANS turbulence model types, such as the
and
models, and
with
enhanced wall treatment, require (or permit) that the near wall grid cells are within the viscous sublayer (
). While near wall grid cells in the buffer layer may be handled with blending functions, they are a major source of misrepresentation of wall shear stresses and should be avoided. LES and DNS approaches generally require
. Near wall grid cells outside the appropriate
range may result in incorrect turbulence production and turbulent viscosity, hence the flow resistance.
Due to the complex geometrical nature of the nasal cavities, it is expected that boundary layer thicknesses will experience significant spatial and temporal variations due to the breathing cycle. The best option might therefore be to ensure that the computational mesh is fine enough to maintain near wall cells in the viscous sublayer, everywhere, for all flowrates, and to utilize a suitable turbulence model. Inthavong
et al. [
106] discussed mesh resolution requirements for laminar nasal air flow and suggested that
in the near-wall grid cells.
Outside the near-wall region, the spatial resolution must be sufficient to resolve all relevant flow structures. It has been suggested that 4-6 million grid cells is generally sufficient to achieve mesh independence [
85,
106,
176], but this is a generalization that should be accepted with caution, since it might not be appropriate for all nasal geometries and volumetric flowrates.
Adaptive meshing, which is a technique that regenerates and/or adapts the mesh based on predefined flow field criteria, is an approach that may be well suited to respiratory breathing, where a wide range of flow characteristics and features can be expected, and a fixed mesh might not be the best choice for the entire range of volumetric flowrates. This technique allows the refinement of the mesh in regions of steep gradients or small flow features as well as coarsening of the mesh where spatial variations are modest. This further allows non-constant number of grid cells, reducing computational cost and giving shorter computation times when the flowrates are lower. It does come at the computational cost of remeshing/adjusting the computational mesh, however. Adaptive meshing is not exclusive to transient simulations but may also be used to improve accuracy in steady state simulations. The present author is not aware of any studies investigating this for nasal airflow.
Temporal Resolution
The temporal resolution determines simulations' ability to correctly describe transient variations in the flow fields. E.g., long time steps might not be able to capture quickly fluctuating phenomena. The time step size is commonly characterized by the dimensionless Courant-Friedrich-Lewy (CFL) number [
178,
179] which expresses the ratio of the advected distance during one time step,
, to the characteristic grid size,
,
where
denotes the flow velocity through the grid cell, and
is the time step size.
The number plays a critical role in ensuring the numerical stability and accuracy of CFD solvers. Explicit CFD solvers restrict information propagation to the maximum of one grid cell per time step (). Consequently, this limitation forces the use of exceedingly short time steps when dealing with small geometry features resolved by small grid cells, resulting in prolonged and costly simulations. Implicit solvers, on the other hand, permit longer time steps, but incorrect simulation results can be the result for too large numbers.
On a fixed computational mesh, with a fixed time step the CFL number tends to zero at the culmination of the inspiratory and expiratory phases of the respiratory cycle. Depending on the numerical scheme employed, short time steps can introduce numerical diffusion, blurring the details of the flow, but the main downside is an unnecessarily high number of time steps. To mitigate this issue, adaptive time stepping strategies can be employed, where the time step size increases as the volumetric flowrate decreases. This approach helps maintain favorable numbers and reduces computational cost. However, it's important to note that since the volumetric flowrate eventually dwindles to zero, a numerical scheme capable of handling low numbers remains essential.
Boundary conditions
Setting appropriate boundary conditions is a crucial step in configuring CFD models. Accurately describing flow parameters such as velocity, turbulence intensity, pressure, and temperature at the boundaries often presents a challenging task, leaving CFD engineers to rely on best available estimates or educated guesses. Consequently, it is important to position the boundaries at sufficient distance from the region of interest to prevent undue interference with the essential details of the flow. However, expanding the simulation domain to achieve this boundary distance unavoidably incurs higher computational costs. Thus, the determination of boundaries locations necessitates careful balance between accuracy and cost- efficiency.
In the specific context of nasal airflow analysis, the boundaries include the walls of the nasal cavity and the flow in- and outlets.
The walls are typically treated as non-slip, smooth boundaries. Nevertheless, the presence of the mucosal lining introduces the possibility that surface roughness and slip conditions might need consideration. While the nasal wall temperature is commonly assumed to fall within the range of normal body core temperature, this assumption may require more careful consideration if the inhaled air is significantly colder.
The nostrils serve as inlets to the nasal cavity during inhalation and outlets during exhalation. However, it is reasonable to suspect that truncating the computational domain at the nostrils may compromise the accurate description of airflow entering or exiting the nasal cavity. An alternative approach is to extend the computational domain to encompass the external airspace around the nose to achieve a more realistic airflow distribution at the nostrils. A study by Taylor
et al. [
180] suggested that the qualitative description of the inflow conditions at the nares may not be critical when computing general flow patterns and overall measures, but for detailed regional flow patterns, carefully chosen inflow conditions may be necessary.
Modeling the entire airway, including the lungs and alveoli, is impractical in nasal airflow studies. Therefore, the airway is typically truncated somewhere in the laryngopharyngeal tract. The location of this truncation has traditionally been based on available computational resources and the specific phenomena of interest. Although the location of truncation may be less critical during inhalation, more attention may be warranted during exhalation. Wu
et al. [
181] demonstrated in a physical experiment, that the flow in the pharynx is laminar during normal breathing, but Bradshaw
et al. [
159] highlighted the importance of including a realistic pharyngeal tract to achieve accurate flow conditions in the nasopharynx during exhalation. The pharyngeal tract is a complex, soft-tissue-enclosed flow channel susceptible to head and neck movements, swallowing, tongue movement, and compliance with over/under pressure due to breathing. The exhalatory flow pattern entering the nasopharynx is likely to be affected by this. The level of realism required in the pharyngeal tract to attain acceptable inflow to the nasopharynx is still unresolved.
Once the boundary locations are determined, careful selection of boundary types (e.g. Dirichlet or Neumann) is required to ensure uniqueness of solution, before defining boundary values. These boundary values can be constant or vary with time and/or position along the boundaries. There is generally significant uncertainty associated with the determination of the local boundary values, necessitating sensitivity analysis to evaluate the impact of boundary conditions.