You are on page 1of 16

Available online at www.sciencedirect.

com
ScienceDirect

Mathematics and Computers in Simulation 160 (2019) 23–38


www.elsevier.com/locate/matcom

Original articles

Capturing the wall turbulence in CFD simulation of human


respiratory tract
Vivek Kumar Srivastava , Akshoy R. Paulb ,∗, Anuj Jainb
aMotihari College of Engineering, Motihari 845401, Bihar, India
b Motilal Nehru National Institute of Technology Allahabad, Prayagraj 211004, India
Received 21 September 2017; received in revised form 9 June 2018; accepted 19 November 2018
Available online 7 December 2018

Abstract
The present study is concentrated on simulating the turbulent airflow in human respiratory tract using computational fluid
dynamics (CFD). Attention is emphasized on local Reynolds number in terms of y + (y-plus) value for the near wall mesh
refinement using a finite volume method (FVM) based CFD solver to capture wall turbulence parameters. Turbulence in human
respiratory tract is formed at high breathing conditions, typically during the running or exercise. The turbulence formation in
human respiratory tract also occurs due to its irregular cross-sections and curved surfaces of the respiratory walls. CT scan based
realistic respiratory model is considered for the study. Two turbulence models, namely realizable k–ε and low Reynolds number
k –ω models are used in the study to capture turbulent flow phenomena for inspiratory flow condition at 60 liter/minute. During
turbulent flow, large velocity gradient occurs near the respiratory wall and hence correct y+ value is essential for accurate prediction
of wall bounded flow using these turbulence models. The outcome of the paper is based on the judicious selection of y+ value and
the turbulence models, which are found vital to correctly simulate airflow in respiratory tract at heavy-breathing conditions.
⃝c 2018 International Association for Mathematics and Computers in Simulation (IMACS). Published by Elsevier B.V. All rights
reserved.
Keywords: Human respiratory tract; Local Reynolds number; Turbulence kinetic energy; Turbulence energy dissipation; CFD

1. Introduction
Turbulence is still the unpredictable phenomena in many real life problems. Turbulence study of human respiratory
system is even more challenging due to irregular bronchus, flexible wall and its asymmetrical nature, especially at
heavy breathing conditions. Therefore, it is necessary to find out suitable turbulence model and suitable computational
grids that capture the correct internal physiological flows of human respiratory tract during heavy breathing conditions.
Weibel [29] had given dimensions of 23-generation human respiratory model based on uniform cross sectional pipes.
But, turbulence studies based on this model do not reflect the realistic physiological conditions as actual respiratory
tracts are irregular and asymmetrical in shapes. Most of the previous studies on human respiratory tract considered
simplistic Weibel respiratory model for turbulence study [11,16,33,34]; whereas few other studies reported realistic
computed tomography (CT) scan based respiratory model for turbulent flow [5,9,18,20,25].
∗ Corresponding author.
E-mail address: arpaul@mnnit.ac.in (A.R. Paul).

https://doi.org/10.1016/j.matcom.2018.11.019
0378-4754/⃝ c 2018 International Association for Mathematics and Computers in Simulation (IMACS). Published by Elsevier B.V. All rights
reserved.
24 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

Turbulence in human airways is formed due to the curvature, constriction, irregular and bifurcation junctions, and
affect throughout the respiratory tract. Turbulent is formed in the upper respiratory tract having low Reynolds number
(Re < 9000). Srivastav et al. [25] earlier carried out CFD simulation in human respiratory tract at general Reynolds
number (Re = ρV d/µ) of 3269, which is calculated based on an airflow rate of 45 liter/minute and hydraulic diameter
(d) measured at the inlet of trachea.
Luo and Liu [17] simulated inspiratory flow in a CT-scan respiratory model using Low Reynolds Number (LRN)
k–ω turbulence model. It is found that with increase in Reynolds number, airflow shifts to the left main bronchus
and moves further downstream. Moreover, secondary flow present in the turbulence flow seems weaker than that in
the laminar flow condition. Forman et al. [5] studied aerosol deposition in a CT-scan based respiratory tract. LRN
k–ω turbulence model was used to obtain deposition efficiency of aerosol inhaled with air. The authors predicted that
an increase in Reynolds number influences the aerosol deposition due to creation of turbulence. Also, the turbulent
flow changed into transition and subsequently into laminar regime as the flow moves towards lower bronchus due to
reduction of respiratory tract diameter. Nithiarasu et al. [20] used unstructured ‘Characteristic meshing Based Split’
(CBS) scheme for computational study of airflow in the human upper airway. In the study, the authors used one-
equation Spalrat-Allmaras (SA) turbulence model for low and moderate Reynolds number flows in the respiratory
tract and found that SA model is capable of capturing the flow physics of respiratory tract. Luo and Liu [18] used CT
scan based model to study the turbulence behavior in the human respiratory tract. LRN k–ω turbulence model was
used in the study. The authors predicted that laryngeal jet in the upper respiratory tract produces turbulence which has
significant effect on airflow physics and particle deposition in it. Laminar to turbulent fluid–nano particle deposition
and the dynamic simulation was performed by Zhang and Kleinstreuer [31]. They compared different turbulence
models and found that Shear Stress Transport (SST) k–ω turbulence model offers a better prediction of turbulence
kinetic energy (TKE). In the same context, Liu et al. [14] studied airflow and particle deposition in children’s upper
respiratory system using LRN k–ω turbulence model.
Jeong et al. [8] used low Reynolds number k–ω model to study the obstructive sleep apnea disease. They obtained
minimum intra-luminal pressure and maximum aerodynamic force at the pharynx region. Patient-specific lower
respiratory tract model under different boundary conditions was studied by Backer et al. [4]. They predicted that
in order to obtain accuracy limits of gamma scintigraphy scan, both the specific model and boundary conditions play
a crucial role. Inthavong et al. [6] used LRN k–ω turbulence model to study the deposition of micron-sized particle
under different breathing conditions. They found that micron-sized drug deposition depends on the breath holds of the
patient.
Tawhai and Lin [27] discussed the appropriate turbulence models for image-based modeling of human respiratory
tract. Luo et al. [16] compared the different turbulence models (k–ε, k–ω and LES) to study the flow physics in
first generation respiratory tract model based on Weibel model [29]. Among the different turbulence models used,
Large Eddy Simulation (LES) model was found suitable which captures details of internal flow physics of the human
respiratory model. Lambert et al. [12] also used LES model to study the regional deposition of particles in CT scan
model. Greater proportion of particles is found in the left lung as compared to the right lung. In LES model, turbulence
flow is recognized by eddies with wide range of length and time scales. However, in LES, the dynamics of the large
scale turbulence are computed explicitly. Hence, LES needs to be three-dimensional and transient leading to higher
computation time and cost (Ansys user’s guide, [3]).
Table 1 summarizes different turbulence models used by the researchers for flow through human respiratory tract.
Based on the application of various turbulence models, the following comments can be made: (i) SA model is cost
effective but does not cover whole range of eddy scale. (ii) It is found that for constant-pressure boundary layer flow,
both k−ε and k–ω models give good performance. However, during boundary layers with adverse pressure gradient,
LRN k–ω model gives better prediction of flow features [2]. (iii) LES model involved filtering function which passes
large eddies and rejects the small eddies. The main advantages of LES model is that large eddies, which are hard
to model, are simulated directly. The un-resolved turbulence scales (small eddies) are modeled using sub-grid scale
(SGS) model. But this model is computationally expensive.
LRN k–ω turbulence model is the appropriate turbulence model because it not only captures the turbulence behavior
in upper respiratory tract, but also captures the laminar flow behavior at low Reynolds number in lower bronchus.
Thus, LRN k–ω turbulence model is comparatively suitable model for the modeling of laminar–transitional–turbulent
behavior of human respiratory tract [11,18,20].
In human respiratory system, turbulent flow is significantly influenced by the presence of irregular wall
(cartilaginous rings in trachea and first bronchus), curvature and dividers of the respiratory tract. The accurate
V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38 25

Table 1
Turbulence models used for computation of airflow in human respiratory tract.
Authors Geometry Turbulence models Remarks
Nithiarasu et al. [20] CT scan Spalrat Allmaras (SA) Spalrat-Allmaras (SA) model is one equation turbulence
model, which is cost effective and capable to capture the
internal flow physics.
Luo et al. [16] Simplified model k–ε, k–ω and LES LES was found suitable model with experimental data.
Kleinstreuer and Zhang [10] Oral cavity and trachea LRN k–ω Best fitted for laminar–transitional–turbulent flow
Luo and Liu [17] CT Scan LRN k–ω Suitable for laminar–transitional–turbulent flow.
Luo and Liu [18] CT Scan LRN k–ω Suitable for laminar–transitional–turbulent flow
Liu et al. [14] Simplified model LRN k–ω Good for re-circulating flow, accurately predicting velocity
profile, pressure drop and shear stress for low Reynolds
number.
Zhang and Kleinstreuer [31] CT Scan SST k–ω SST k–ω model gives a better prediction of turbulence kinetic
energy.
Tawhai and Lin [27] Review paper LES LES relies on a sub grid scale (SGS) turbulence model which
is fairly accurate and is becoming affordable with modern
computing technologies.
Lambert et al. [12] CT Scan LES Explicitly resolves large-scale Energy-containing turbulent
eddies and parameterizes small-scale eddies with a SGS
model.
Liu et al. [14] Simplified model LRN k–ω Studied human upper respiratory tract.
Jamasp [7] CT Scan SST k–ω & LES LES is used in the context that higher flow rates caused the
airflow to be more complicated.
Srivastav et al. [26] CT scan LRN k–ω Drug delivery on tumorous respiratory tract model.

presentation of near-wall treatment obtains correct prediction of bounded turbulent flow. The y+ calculates the local
Reynolds number at the wall and is used in CFD √ to define coarse or fine mesh near the wall [24]. Local Reynolds
ρu y
number is defined as:y + = µτ p , where u τ = τw /ρw is the friction velocity, y p is the distance from point p (first
grid line) to the wall, ρ is the fluid density, and µ is the fluid viscosity at any point p [2]. Airflow in respiratory tract
is the case of low Reynolds number, where viscous sub-layer is formed and y+ value throughout the wall is less than
5. Most of the previous studies ignored the effect of selection of y+ value on the results of flow simulation.
Salim and Cheah [24] presented that viscosity affected region is divided into three zones: Viscous sub-layer (y+
< 5), Buffer layer (5 < y+ < 30) and Fully turbulent or log-law region (y+ > 30 to 60). Researchers used near-
wall-treatment option in commercial packages but in spite of using these options it sometimes show buffer layer near
the wall. To reduce this buffer layer, inner wall region needs very fine mesh to capture the flow physics of near wall
bounded flow. Kleinstreuer and Zhang [10] had taken value of y + ⩽ 1 for case of LRN turbulence modeling. Longest
and Vinchurkar [15] were also maintained y + value approximately 1 at all boundary walls for LRN k–ω turbulence
model.
The first property of turbulence is recognized by the fluctuating component of velocity and is defined in terms
of turbulence kinetic energy (TKE). Turbulence kinetic energy of the fluid flow is defined as the square root of the
fluctuating components
( of velocity.
)
TKE = f u ′2 + v ′2 + w ′2 , where u ′ , v ′ and w′ are the fluctuating velocity components. Since most of the multi-
phase flow phenomena in human respiratory tract is either drug delivery and toxic particle deposition or tissue injury
occurring on the respiratory wall, very fine mesh is required near the wall to capture the wall bounded turbulent flow.
This may be done either using wall treatment option or by maintaining y+ value. In spite of using best wall treatment
option, it does not compute correct y+ value (< 5 for viscous sublayer) at the human respiratory wall.
Different turbulence models have been used by the researchers for solving airflow and particle deposition in
respiratory tract. The high breathing condition typically occurs during running or exercise condition that not only
involves turbulence flow but also transition and laminar flows. This is because summation of cross sectional area of
a respiratory tract increases as move towards lower bronchus and as a result velocity is decreased. Thus appropriate
turbulence model is required that enables to capture the internal flow physics in all the three breathing conditions (low,
moderately and high) simultaneously.
It is well known that particles (toxic or drug) are driven by inhaled air. In respiratory system, patient sucks air from
mouth based on his capacity and tries to hold the breath to enable drug to reach deep into the lungs. In such a situation,
26 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

the flow pattern is not same as in normal respiratory cycle through nose. The airflow pattern is quite steady for short
duration. Therefore, similar pattern is considered in the present study as is also followed by other researchers [28,32].
It is seen that researchers have applied different turbulence models for simulating the airflow and two-phase study
of human respiratory tract. However, the importance of local Reynolds number has not been reported in the previous
studies, which plays a crucial role for computing wall turbulence and related parameters — like wall shear stress
(WSS), which is an indicative of wall injury to the respiratory tract. The objective of the present study is therefore,
to investigate the effects of wall turbulence on bifurcation junctions of respiratory tract in terms of turbulent kinetic
energy (TKE), turbulent energy dissipation (TED), wall shear stress (WSS) and local Reynolds number (y+ ) value
while using popular turbulence models in respiratory flow dynamics.

2. Human respiratory tract model


The respiratory tract model is constructed from CT scan data of human using medical imaging software
MIMICSTM [19]. The scan is done from the inlet of trachea to lower part of the respiratory tract.
The range of Hounsfield unit used for visualization of the respiratory tract boundary is between −1024 and −600.
Total 519 CT slices with slice-distance of 0.625 mm and resolution of 512 × 512 pixel are used for the complete
reconstruction of sixth-generation (G0 to G6) respiratory tract model and are shown in Fig. 2. The left and right
lungs are divided into several bifurcation junctions, such as left lung (bifurcation from BiF-2L to 5L) and right lung
(bifurcation from BiF-2R to 6R) and are shown in Fig. 2.

3. Grid generation
Accuracy of the computational results depends on the computational cell/element size and number of elements
present in the flow domain. As realistic human respiratory tract is irregular in shape, hence, tetrahedral element has
been adopted to improve the quality of meshing, especially near the boundary. Surface and volume meshing is done
using the triangular and tetrahedral elements respectively. Grid independency test (GIT) is performed to confirm
the computational accuracy of the airway model. Grid refinement is done using adaptation techniques provided in
Ansys-ICEM and is shown in Fig. 3. Initially 1.75 million elements is generated in the respiratory tract model, which
is further refined two times for 2.85 and 3.7 million elements. It is found that the results become grid independent
for 2.85 million elements. Maximum local Reynolds number (y+ ) corresponding to 2.85 million elements is 7. It is
further refined several times near the respiratory wall boundary to obtain y+ less than 3. Finally 4.09 million tetrahedral
elements are created corresponding to y+ ≈ 3.

4. Governing equations and turbulence models


Conservation of mass
∂ρ ∂(ρu i )
+ =0
∂t ∂x j
Conservation of linear momentum
∂u i ∂(u i u j ) 1 ∂p µ ∂ 2ui
+ =− +
∂t ∂x j ρ ∂ xi ρ ∂x j∂x j
where µ, ρ are viscosity and density of flowing fluid, p is pressure, u is the velocity components u i, j (i, j = 1, 2, 3) in
x, y and z direction. The airflow is assumed to be steady and incompressible in the present study.

4.1. Turbulence models

Turbulence is defined by the two parameters, generation of turbulence kinetic energy (k) and dissipation of
this energy (ε). These two parameters of turbulent flow are incorporated in the two-equation turbulence models.
Wilcox [30] described that a turbulence model should be one which can be applied to a given turbulent flow by
defining the most appropriate boundary/initial conditions. The complexity of the turbulence models may differ in
each problem. Considering the human respiratory tract and all the three flow phenomena (laminar, transition and
turbulent) are happened simultaneously therein, suitable turbulence model must be used to obtain real physiological
flow characteristics in the respiratory tract system.
V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38 27

4.1.1. Realizable k−ε turbulence model


The realizable k–ε turbulence model is suitable for the flow involving flow separation, boundary layer under strong
pressure gradient and recirculation and therefore is used in the present study. Enhanced wall treatment is adopted to
determine the flow parameters near the wall.
The transport equations for k (turbulent kinetic energy) and ε (turbulent dissipation rate) in the realizable k−ε
turbulent model are given below.
k-equation:
∂ ∂ µt ∂k
[( ) ]
(ρku j ) = µ+ + G k − ρε
∂x j ∂x j σk ∂ x j
ε-equation:
∂ ∂ µt ∂ε ε2
[( ) ]
(ρεu j ) = µ+ + ρC1 Sε − ρC2 √
∂x j ∂x j σε ∂ x j k + νε
[ ]
η
where, C1 = max 0.43, η+5 and C2 = 1.9, Cµ = 0.09 are constants.
η = S kε where, S is the modulus of the mean rate of strain tensor, defined as S =

2Si j Si j
µ is dynamic viscosity.
2
Turbulent/eddy viscosity is given as µt = ρCµ kε
Turbulent Prandtl number for k, σk = 1,
Turbulent Prandtl number for ε, σε = 1.2

4.1.2. Low Reynolds Number (LRN) k−ω turbulence model


It is found that LRN k−ω turbulence model is relatively better for the modeling of laminar–transitional–turbulent
airflows in human respiratory tract [11,17,18]. The transport equations for k (turbulent kinetic energy) and ω (specific
turbulent dissipation rate) in the k−ω turbulent model are furnished below.
k-equation:
∂k ∂u i ∂ ∂k
[ ]
uj = τi j − β kω +

(v + σk νT )
∂x j ∂x j ∂x j ∂x j
ω-equation:
∂ω ω ∂u i ∂ ∂ω
[ ]
uj = α τi j − βω2 + (v + σω νT )
∂x j k ∂x j ∂x j ∂x j
ν, νT and τi j are kinetic molecular viscosity, turbulent viscosity and Reynolds stress tensor respectively. νT =
Cµ f µ k/ω and function
f µ = exp −3.4/(I + RT /50)2
[ ]

with RT = k/µω, µ being dynamic molecular viscosity. Model constants are


Cµ = 0.09, α = 0.555, β = 0.8333, β ∗ = 1 and σk = σw = 0.5

5. Boundary conditions
Boundary conditions in the CFD analysis play an important role for accurate prediction of the flow physics. Inlet
mass flow condition is applied at the inlet of trachea and pressure outlet condition at the outlets. Inlet velocity is 4.56
m/s for an airflow rate of 60 liter/minute corresponding to Re ≈ 5186, which necessarily denote turbulent condition
for pipe/conduit flow. Outlet pressure is equal to zero gauge pressure and no-slip condition is set at the respiratory
wall.

6. Near wall turbulence modeling


The near-wall treatment determines the accuracy of the wall stresses and of the near-wall turbulence prediction.
Hence, appropriate near-wall turbulence modeling is crucial in order to capture important flow features. In this respect,
wall functions are used, which is a set of empirical rules that are based on the logarithmic law of the wall. The wall
28 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

Fig. 1. Application of wall functions to CFD simulation.


Source: Excerpted from Anderson et al. (2012).

functions are essential to avoid having dense meshes in CFD simulations or they may be needed since particular
turbulence models are not valid in the viscosity-affected near-wall region. The wall functions estimate the velocities
(u, v and w), turbulent kinetic energy (k), turbulent dissipation (ε) or turbulent stresses in the RANS models in the first
cell close to the wall.

6.1. Standard wall function

Standard wall function applies the boundary conditions some distance away from the wall so that the turbulence
model is not solved close to the wall. These wall functions allow calculations to be carried out with the first grid
point, P, in the region where the wall function is valid, rather than on the wall itself, as shown in Fig. 1. The boundary
conditions are applied at P, which represents the first grid point while W represents the corresponding point on the
wall. Thus, the wall functions allow the rapid change of flow variables occurring within the near-wall region to be
accounted for without resolving the viscous near-wall region.

6.2. Enhanced wall function

Near wall treatment with ‘enhanced wall function’ is used in the present study. Enhanced wall treatment is a
near-wall modeling method that merges a two-layer model with so called enhanced wall functions. If the near-wall
mesh is fine enough to be able to resolve the viscous sub-layer, then the enhanced wall treatment will be identical
to the traditional two-layer zonal model (Ansys-Fluent User Guide). It provides better prediction near the wall in
comparison of others.
The general recommendation for a standard wall function is 30 < y+ < 100, preferably in the lower Reynolds
number region. At high Reynolds number, the logarithmic law is valid up to higher y+ and the upper limit may
increase to 300–500. For low-Re models and enhanced wall functions the first grid point should be close to y+ =
1, and there should be at least ten grid points in the viscosity-affected near-wall region, i.e. y+ < 20. It is however,
worthy to mention that as with the choice of selecting an appropriate turbulence model, the choice among the near-
wall modeling approaches is strongly coupled to the physics of the particular flow and the computational resources
available.

7. CFD simulation
A finite volume method (FVM) based CFD solver — Ansys-Fluent 15 is used for the computational simulation
of aerosol deposition in image based respiratory tract model. Convergence criteria for the governing equations
were set 10−5 for all the cases. Pressure, momentum and turbulent kinetic energy are discretized using second-
order upwind scheme while specific dissipation rate was done by QUICK (Quadratic Upstream Interpolation for
V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38 29

Fig. 2. Human respiratory tract model.

Fig. 3. Computational grid of the respiratory tract model.

Convective Kinetics) scheme [13]. The SIMPLE algorithm was applied in the CFD solver for pressure velocity
coupling [22].

8. Validation of CFD results


The validation of computational results is done with the results reported by Nowak et al. [21]. For the same
respiratory model and boundary conditions, velocity profile was obtained at the mid plane of fourth generation. The
validation is performed for the inspiratory flow rate of 40 L/min. Both the results are shown in Fig. 4, which is within
the acceptable limit.
30 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

Fig. 4. Relative velocity at mid section of bronchus-4.

9. Results and discussion


Turbulence studies in a CT scan model at steady flow conditions are presented. In this section, local Reynolds
number, velocity contours, wall shear stress distribution and effect of turbulence are computed.

9.1. Local Reynolds number (y+ ) distribution

It defines the structure of boundary layer (near wall treatment). The near wall boundary layers are divided into
three regions: viscous sub-layer, buffer layer and fully turbulent region. Grid adaptation technique is used for different
refinement levels to obtain within the viscous sub-layer range.
Local y+ value corresponding to different computational elements is plotted in Fig. 5. It is suggested that y+ must
be refined up to 1 to obtain the viscous sub-layer while it is acceptable for y+ < 5 as recommended by Wilcox [30].
Total number of elements reached 7.06 million when y+ = 2 and hence, it requires more computational time.
The grid refinement near the wall of inlet cross-section is shown in Fig. 6. The same grid refinement is retained
throughout the wall. It is seen that wall boundary thickness increases with the increase of y+ value. Wall refinement
for y+ ≈ 7 and y+ ≈ 3 or 2 formed buffer and viscous sub-layer respectively. This refined mesh (y+ ≈ 3, 2) provides
better computational prediction near the respiratory wall.

9.2. Velocity distribution

In order to compare the flow characteristics for two different values of y+ , velocity contours are plotted at two
different locations (plane-1 and 2) of the respiratory tract and are shown in Figs. 7 and 8. It is clear from the velocity
contours (plane-1, Fig. 7) that asymmetrical velocity zones are formed throughout the region. The local maximum
velocity corresponding to the y+ < 7 and y+ < 3 are 9.05 m/s and 9.33 m/s respectively, showing 3.01% deviation.
Different flow features are observed for two different values of y+ (Fig. 7). Maximum velocity gradient occurs near
the inner wall, which is captured very well at y+ < 3. Also, plane-1 is located at the cartilaginous ring where local
vortices create disturbance and these local vortices decide the path of incoming particles. Thus, y+ values must be
taken between 1 and 5 (in viscous sub-layer region) to obtain good computational results and also to predict the
realistic physiological flow features near the wall of the respiratory tract.
As the airflow crosses first bifurcation it skews towards the inner wall, and therefore reflect the occurrence of large
velocity gradient. Plane-2 is located at middle of left main bronchus, where velocity is skewed towards the inner wall
due to centrifugal forces. It is seen in Fig. 8 (zoomed view) that lower value of y+ (<3) (viscous sub-layer) captured
flow physics better than larger y+ (< 7) value (buffer layer).
V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38 31

Fig. 5. Maximum wall y+ versus number of elements.

Fig. 6. Computational grid arrangement near the wall for various y+ (zoomed view).

The whole respiratory tract is irregular and having curved surfaces, where flow separation affects the particle
direction and deposition. Therefore, correct selection of y+ value will help to improve the computational results of
human airways modeling.

9.3. Velocity contours

The velocity contours at different cross section of the human respiratory model are shown in Fig. 9. Airflow velocity
inside the human airways helps to find the internal flow physics of the human lungs. It is found that skewed velocity
towards the anterior wall of trachea due to presence of cartilaginous rings. Once the airflow reaches at bifurcation
junction it bifurcates into two parts namely left and right bronchus flow. The airflow patterns in human respiratory
tract will help the medical practitioner to understand the internal flow physics of respiratory tract.
32 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

Fig. 7. Velocity contours at plane-1 (at y+ < 7 and y+ < 3).

Fig. 8. Velocity contours at plane-2 (at y+ < 7 and y+ < 3).


V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38 33

Fig. 9. Velocity contours at different locations.

9.4. Wall shear stress distribution

Average wall shear stress (WSS) value at left and right bifurcation junctions is computed which is shown in Figs. 10
and 11. In order to see the effects of y+ on different bifurcation walls, three different values of y+ (4, 3 and 2) are
considered in the study. Average WSS is found closer at all the bifurcation junctions (BiF-2L, 3L and 5L) except at
BiF-4L. The average WSS on BiF-4L at y+ ≈ 2 is 0.54 N/m2 while it shows 0.58 N/m2 for y+ ≈ 3. Hence there
has 6.8% deviation for two WSS values at different y+ . Wall injury occurs in human respiratory tract due to peak
value of WSS, which depends upon the acting time. Thus, wall refinement must be done to capture the accurate flow
characteristics near the wall.
There is an asymmetrical distribution of mass flow at left and right lung; therefore, WSS is computed separately.
Higher WSS is obtained for y+ = 2 as compared to y+ = 3 or 4. It is seen from Fig. 11 that average WSS increases
with decrease of y+ value from 4 to 2. The WSS deviation at BiF-5R for the y+ ≈ 3 and 2 is 4.13%. Thus wall grid
refinement in computational domain increases the accuracy of CFD simulation.
34 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

Fig. 10. Wall shear stress at left bifurcation.

Fig. 11. Wall shear stress at right bifurcation.

9.5. Effects of turbulence

Turbulence plays important role in the upper respiratory tract because large eddies are broken into smaller eddies
through the cascading process near the wall of respiratory tract. In cascading process, kinetic energy interacts the
turbulence at largest scales at which energy is extracted from the mean flow. Also various parameters are varied with
space and time [1].
Two additional transport equations — one for turbulence kinetic energy, another for dissipation of turbulence
kinetic energy are required to close the Reynolds averaged Navier–Stokes (RANS) equations. Turbulence in the
respiratory tract is produced because of constriction, sudden expansion and contraction of respiratory wall. In order
to find out the suitable turbulence model for human respiratory tract, two turbulence models are chosen for CFD
simulation. Since human respiratory tract has irregular and curved wall, therefore dissipation of kinetic energy occurs
at these locations.
V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38 35

Fig. 12. Turbulence kinetic energy (m2 /s2 ) distribution at bifurcation junctions (LRN k–ω model).

Realizable k−ε model gives improved results for swirl flows and flow separation. However LRN k–ω model works
well with adverse pressure gradient and separating flow. One of the advantages of LRN k–ω model is that it does not
require wall function.

9.5.1. Turbulence kinetic energy


Turbulence kinetic energy at different bifurcation junctions for LRN k–ω turbulence model is presented in Fig. 12.
Maximum TKE is found at bifurcation junction indicating large velocity gradient at these locations. In similar context,
Radhakrishnan and Kassinos [23] also studied that locations with maximum TKE are those where velocity fluctuates
rapidly. Thus, Turbulence has prominent effect at the bifurcation junctions of the respiratory tract. Initially, local TKE
is maximum at first bifurcation (BiF-1) but decreases as flow moves towards lower bifurcation junction.
TKE for realizable k–ε model is shown in Fig. 13. The maximum local kinetic energy at BiF-1 is found at 0.43
m2 /s2 against the value 0.72 m2 /s2 obtained from LRN k–ω model. Thus a significant variation of TKE is obtained at
different bifurcation junction for the two turbulence models.
It is seen from the plotted contours that maximum local TKE impacts on the center of flow dividers, which plays a
crucial role during deposition of toxic particles.
36 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

Fig. 13. Turbulence kinetic energy (m2 /s2 ) distribution at bifurcation junctions (realizable k–ε model).

9.5.2. Turbulence energy dissipation (TED)


The dissipation of turbulence kinetic energy (ε) of different bifurcation junction is plotted in Fig. 14. The maximum
probability of dissipation of kinetic energy is found at the bifurcation junction, which further leads the wall injury.
Dissipation of turbulence kinetic energy is typically occurred near the wall. All the smallest scale eddies near the
respiratory wall and curved boundaries are dissipated because of viscous stresses.

10. Conclusions

The Computational study of airflow in three-dimensional, CT scan model of human respiratory tract is done with
turbulent flow condition. In order to see the effects of turbulence TKE, TED and y+ values are computed at different
locations, especially at bifurcations of the respiratory wall. The results led to the following conclusions:
• Maximum turbulent kinetic energy (TKE) occurs at bifurcation junctions where high velocity gradient exists.
Maximum TKE is found at upper bifurcation; however it decreases as flow moves towards lower bifurcation
V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38 37

Fig. 14. Turbulence energy dissipation (m2 /s3 ) (realizable k–ε model).

junctions. Turbulent energy dissipation (TED) also occurs at bifurcation junctions because of curved surfaces
and sudden expansion.
• The near-wall mesh must be refined enough to obtain viscous sub-layer range. The y+ value must be judiciously
put in computation for the correct prediction of near wall flow features.
• This also makes the turbulence model suitable for simulating flow where flow separation occurs. Hence a very
fine mesh is required near the wall in order to capture high velocity gradient which exists in viscous sub-layer
region.
• The importance of choosing correct local Reynolds number (y+ ) while using a turbulence model in the flow
through respiratory tract has been established. The outcome of the present study will therefore, help the
researchers to choose optimum range of local Reynolds number (y+ ) and turbulence model for simulating correct
flow features in the human respiratory system.
38 V.K. Srivastav, A.R. Paul and A. Jain / Mathematics and Computers in Simulation 160 (2019) 23–38

Acknowledgment
The authors are grateful to Science and Engineering Research Board (SERB), Govt. of India for providing
necessary financial assistance vides sanction order no. SR/S3/MERC/0073/2012 to carry out the research work.

References
[1] J. Anderson, Computational Fluid Dynamics, McGraw-Hill Education, 1995, ISBN 0070016852, 9780070016859.
[2] B. Anderson, R. Anderson, L. Hakansson, M. Mortensen, R. Sudiyo, B. van Wachem, Computational Fluid Dynamics for Engineers,
Cambridge University Press, ISBN: 9781107018952, 2012.
[3] Ansys-Fluent, Version 14.5, User Guide. Ansys Inc., USA. 2013.
[4] J.W.D. Backer, W.G. Vos, C.D. Gorle, P. Germonpre, B. Partoens, F.L. Wuyts, P.M. Parizel, W.D. Backer, Flow analyses in the lower airways:
Patient-specific model and boundary conditions, Med. Eng. & Phys. (30) (2008) 872–879.
[5] M. Forman, M. Jıcha, J. Katolicky, Aerosol deposition in human airways during breathing cycle, Appl. Comput. Mech. (1) (2007) 437–444.
[6] K. Inthavong, L.T. Choi, J. Tu, S. Dinga, F. Thien, Micron particle deposition in a tracheobronchial airway model under different breathing
conditions, Med. Eng. & Phys. (32) (2010) 1198–1212.
[7] Azarnoosh Jamasp, Computational Fluid Dynamics Simulation of the Airflow Through the Human Respiratory Tract (Master Thesis), The
University of Tnnessee at Chattanooga, Tennessee, 2016.
[8] S.J. Jeong, W.S. Kim, S.J. Sung, Numerical investigation on the flow characteristics and aerodynamic force of the upper airway of patient
with obstructive sleep apnea using computational fluid dynamics, Med. Eng. & Phys. (29) (2007) 637–651.
[9] N.H. Johari, K. Osman, N.H.N. Helmi, M.A.R.A. Kadir, Comparative analysis of realistic CT-scan and simplified human airway models in
airflow simulation, Comput. Methods Biomech. Biomed. Eng. (2013) http://dx.doi.org/10.1080/10255842.2013.776548.
[10] C. Kleinstreuer, Z. Zhang, Laminar-to-turbulent fluid-particle flows in a human airway model, Int. J. Multiph. Flow. (29) (2003) 271–289.
[11] C. Kleinstreuer, Z. Zhang, Airflow and particle transport in the human respiratory system, Annu. Rev. Fluid Mech. (42) (2010) 301–334.
[12] A.R. Lambert, P.T. Oshaughnessy, M.H. Tawhai, E.A. Hoffma, C.L. Lin, Regional deposition of Particles in an image-based airway model:
large-eddy simulation and left–right lung ventilation asymmetry, Aerosol Sci. Technol. (45) (2011) 11–25.
[13] B.P. Leonard, A stable and accurate convective modeling procedure based on Quadratic upstream interpolation, Comput. Methods Fluids
Appl. Mech. Eng. (19) (1979) 59–98.
[14] Z. Liu, L. Angui, X. Xiaoxia, G. Ran, Computational fluid dynamics simulation of airflow Patterns and particle deposition characteristics in
Children upper respiratory tracts, Eng. Appl. Comput. Fluid Mech. (6) (2012) 556–571.
[15] P.W. Longest, S. Vinchurkar, Validating CFD predictions of respiratory aerosol deposition: Effects of upstream transition and turbulence, J.
Biomech. (40) (2007) 305–316.
[16] X.Y. Luo, J.S. Hinton, T.T. Liew, K.K. Tan, LES modelling of flow in a simple airway model, Med. Eng. & Phys. (26) (2004) 403–413.
[17] H.Y. Luo, Y. Liu, Modeling the bifurcating flow in a CT-scanned human lung airway, J. Biomech. (41) (2008) 2681–2688.
[18] H.Y. Luo, Y. Liu, Particle deposition in a CT-scanned human lung airway, J. Biomech. (42) (2009) 1869–1876.
[19] Mimics, TM 9.1, Materialise, Belgium.
[20] P. Nithiarasu, C.B. Liu, N. Massarotti, Laminar and turbulent flow calculations through a model human upper airway using unstructured
meshes, Commun. Numer. Methods. Eng. (23) (2007) 1057–1069.
[21] N. Nowak, P.P. Kakade, A.V. Annapragada, Computational fluid dynamics simulation of airflow and aerosol deposition in human lungs, Ann.
Biomed. Eng. (31) (2003) 374–390.
[22] S.V. Patankar, Numerical Heat Transfer and Fluid Flow, Hemisphere Pub. Corp., NY, 1980.
[23] H. Radhakrishnan, S. Kassinos, CFD modeling of turbulent flow and particle deposition in human lungs, in: 31st Annual International
Conference of the IEEE EMBS, Minneapolis, Minnesota, USA, 2009.
[24] S.M. Salim, S.C. Cheah, Wall y+ strategy for dealing with wall-bounded turbulent flows, in: Proceedings of the International Multi-
Conference of Engineers and Computer Scientists 2009; II, IMECS 2009, March 18 –20, Hong Kong, 2009.
[25] V.K. Srivastav, A. Kumar, S.K. Shukla, A.R. Paul, A.D. Bhatt, A. Jain, Airflow and aerosol-drug delivery in a CT scan based human
respiratory tract with tumor using CFD, J. Appl. Fluid Mech. 7 (2) (2014) 245–256.
[26] V.K. Srivastav, A.R. Paul, A. Jain, Computational study of drug delivery in tumorous human airways, Int. J. Comput. Sci. Math. (2018) (in
press).
[27] M.H. Tawhai, C.L. Lin, Image-based modeling of lung structure and function, J. Magn. Reson. Imaging (32) (2010) 1421–1431.
[28] D.K. Walters, W.H. Luke, Computational fluid dynamics simulations of particle deposition in large-scale, multigenerational lung models, J.
Biomech. Eng. 133 (2011) 011003-1 to.
[29] E.R. Weibel, Morphometry of the Human of the Human Lung, Academic press, New York, 1963, pp. 110–144.
[30] D.C. Wilcox, Turbulence Modeling for CFD, DCW publishers, Canada, 1994.
[31] Z. Zhang, C. Kleinstreuer, Laminar-to-turbulent fluid–nanoparticle dynamics simulations: Model comparisons and nanoparticle-deposition
applications, Int. J. Numer. Methods Biomed. Eng. (22) (2011) 1930–1950.
[32] Z. Zhang, C. Kleinstreuer, C.S. Kim, Effects of curved inlet tubes on air flow and particle deposition in bifurcating lung models, J. Biomech.
34 (2001) 659–669.
[33] Z. Zhang, C. Kleinstreuer, C.S. Kim, Airflow and nanoparticle deposition in a 16-generation tracheobronchial airway model, Ann. Biomed.
Eng. 36 (12) (2008) 2095–2110.
[34] X.G. Zhao, X.X. Xu, SH.L. Tan, Y.J. Liu, Z.H.H. Gao, Characteristic of Gas-Solid Two-Phase Flow in the Human Upper Respiratory Tract
Model. 2009. DOI: 978-1-4244-2902-8.

You might also like