Detached Eddy Simulation of the flow around a simplified vehicle sheltered by wind barrier in transient yaw crosswind.
Telenta, Marijo ; Subelj, Matjaz ; Tavcar, Joze 等
1. Introduction
Today's ground vehicles are sensitive to crosswind
disturbance. Nonetheless, the focus of vehicle development is on
providing lighter and streamlined vehicles. The vehicle's crosswind
sensitivity deteriorates due to vehicle's aerodynamic design
improvements. A lot of research is dedicated in analyzing the influence
of the vehicle shape on the vehicle's crosswind sensitivity. There
are two objectives for vehicle's aerodynamic design. The first
objective is reduction of drag coefficient in order to improve fuel
efficiency. The second objective is improvement of driving stability
where lift and pitch are important for the straight line driving, and
yawing moment and side force are important for the crosswind sensitivity
[1]. However, these objectives represent conflicting goals. Therefore,
in this work other means for improving crosswind sensitivity are
considered. Instead of analyzing the influence of different vehicle
designs on the vehicle's crosswind stability, wind barrier is
considered for lowering vehicle's crosswind exposure.
Previous studies modeled fluid flow through porous geometries while
not considering the details of the barrier's geometry. The main
focus was to define a suitable resistance model for a given geometry of
a barrier. Previous studies, [2], [3], and [4] used the Reynolds
averaging method with turbulence closure for a two-dimensional fluid
flow simulation in which the porous barrier was represented as a
momentum sink. As stated in [5], numerical methods utilizing the
momentum sink approach for wind barrier modeling treat complex
unresolved flow near and through the gaps at a superficial level. A
deeper understanding of the turbulent structure dynamics is required to
evaluate the barrier sheltering effect. Author's previous work [6],
[7] addressed this issue. URANS numerical simulations, verified with
experimental data, were done modeling the fluid flow through
geometrically accurate three-dimensional barrier model in order to
resolve the flow near and through the porous barrier. The objective was
to investigate the interaction between the bleed flow and the reverse
flow for different barrier configurations. Present paper extents the
research scope with more advanced turbulence models, namely Detached
Eddy Simulation (DES).
DES is computationally more expensive than RANS. However, it
provides more accurate results and gives information about the flow
structures which is out of reach for RANS methods. RANS provides only
the mean information about the flow and the unsteady information is
lost. Also, flow calculation accuracy is dependent on the turbulence
model used. It is difficult to define a RANS model that accurately
represents the Reynolds stresses in the region of separated flow such as
a wake behind the barrier. In addition, complicated flow structures are
developed in the wake region behind the barrier. These wake structures
are dominated by large turbulent structures which can be resolved by DES
method. Increase in the computer capability made DES simulation possible
nowadays. DES is used instead of Large Eddy Simulations (LES) since LES
is not feasible for higher Reynolds number flow which is the case in
this work. Therefore, DES was utilized to simulate the time-development
of the flow around a generic vehicle behind a wind barrier subjected to
a sudden strong crosswind.
The investigation scope in the current work includes vehicle
shelter by geometrically accurate wind barrier. Present numerical
simulation mirrors the experimental work of [1] which is further
expanded with barrier model introduction in computational domain. The
goal was to analyze the barrier influence on aerodynamic loads
development that vehicle is subjected to. Wind tunnel testing of a
vehicle in crosswind was done in [1] where transient yaw crosswind
scenario on a simplified vehicle shape corresponding to sport utility
vehicle was performed. The flexibility of the CFD makes transient
crosswind studies easier to realize than experimental studies. However,
no DES studies with time-dependent boundary conditions investigating the
wind barrier applications were found prior to this work. To date,
transient crosswind with DES method has been investigated on a bus
geometry [8], simple vehicle shapes [9], and high speed trains [10],
[11]. In this work, unsteady crosswind simulation for sheltered moving
vehicle behind the wind barrier subjected to a deterministic gust wind
represented by a continuous and smooth step function is reported.
Advanced boundary conditions are implemented to simulate a gust wind
propagating through the computational domain. Two types of vehicle
aerodynamic research are found: time-averaged aerodynamics testing and
transient aerodynamic testing [12]. Time averaged technique subjects the
vehicle to a constant yaw angle, whereas with the transient techniques
the vehicle is subjected to a rapid change in yaw angle.
Although there are many wind induced traffic accidents every year,
the wind barrier shelter effect on vehicle aerodynamic is not yet
properly investigated. Available studies are still inadequate to give
the complete picture of the flow structures around the moving vehicle
behind the wind barrier in gust wind conditions.
2. Numerical methods
Delayed Detached Eddy Simulation (DDES) is performed to study the
flow behind the wind barrier. DDES is a hybrid technique for prediction
of separated turbulent flows at high Reynolds numbers. Development of
this technique was motivated by the prohibitive computational costs of
applying Large-Eddy Simulation (LES). Thus, high-Reynolds number
separated flows have been predicted using steady or unsteady
Reynolds-averaged Navier-Stokes equations (RANS and URANS). However, the
disadvantage of the RANS methods applied to massive separations is that
the statistical models are designed and calibrated on the basis of the
mean parameters of thin turbulent shear flows containing numerous
relatively standard eddies. Such eddies are not representative of the
comparatively fewer and geometry-dependent structures that typically
characterize massively separated flows.
Reynolds-averaged Navier-Stokes (RANS) equations, Eq. (1) and Eq.
(2):
[partial derivative][[bar.u].sub.i]/[partial
derivative][[bar.x].sub.i] = 0, (1)
[MATHEMATICAL EXPRESSION NOT REPRODUCIBLE IN ASCII] (2)
where [bar.p] is averaged pressure, [[bar.u].sub.i] is averaged
velocity, [rho] is the air density, [mu] dynamic viscosity, and Reynolds
stresses are [rho][bar.[u'.sub.i][u'.sub.i]].
The filtered Navier-Stokes equations for LES, Eq. (3) and Eq. (4),
are as follows [13]:
[partial derivative][[??].sub.i]/[partial derivative][x.sub.i] = 0.
(3)
[MATHEMATICAL EXPRESSION NOT REPRODUCIBLE IN ASCII] (4)
where the velocity [u.sub.i] is separated into the filtered,
resolved part [[??].sub.i] and sub-filtered, unresolved part
[u'.sub.i], and v is kinematic viscosity.
The switch between RANS and LES in DDES, Eq. (5) and Eq. (6), is
expressed as follows [14]:
[??] = d - [f.sub.d]max(0,d - [C.sub.DES][DELTA]), (5)
where:
[f.sub.d] = 1 - tanh ([[8[r.sub.d]].sup.3]), (6)
[r.sub.d] = [v.sub.t] + v/[([U.sub.i,j][U.sub.i,j]).sup.0.5]
[k.sup.2][d.sup.2]. (7)
[DELTA] is the maximum edge length of the local computational cell,
d distance from the wall, [??] length scale, [r.sub.d] parameter,
[f.sub.d] function, and [C.sub.DES] =0.65 is model constant.
3. Barrier and vehicle model
One barrier configuration model was used in the numerical
simulation, as seen in Fig. 1, a. It consists of five horizontal bars,
all with 90[degrees] inclination angle. The barrier length is L =
9.5[L.sub.V], where [L.sub.V] = 0.48 m is the vehicle length. Porosity
of the barrier is 25%. The computational domain is also shown in Fig. 1,
b, where the barrier model length L and height H are 4.56 m and 0.4 m,
respectively. Vehicle model has a box-like geometry where the width
[W.sub.V] and the height [H.sub.V] are the same with value of 0.2 m.
Ground clearance of the vehicle is 0.03 m. In order to avoid moving
mesh, vehicle position is fixed and the velocity is imposed at the
domain inlet.
[FIGURE 1 OMITTED]
4. Solver
A commercial CFD code, Ansys Fluent 14.5, was used to solve
incompressible Navier-Stokes equations of fluid motion. It uses
cell-centered numerics, via a segregated approach, on a collocated,
unstructured grid. The Delayed DES (DDES) method with the standard
Spalart Allmaras (SA) model was used. The diffusive fluxes of the
momentum and turbulent equations are discretized using the central
difference (CD) scheme. The bounded central difference scheme was used
for convective fluxes in LES and RANS regions. Second order upwind was
used for the spatial discretization of the convection terms of the
turbulence model. Second order upwind scheme is less diffusive and
offers more accurate solution over the first order upwind scheme. The
least square method was used for the gradient method. Standard pressure
interpolation was used. Time integration was performed with the
second-order backward Euler scheme and the bounded second-order implicit
Euler scheme for turbulence variables. The DDES simulation is
initialized with steady state RANS simulations. The SST k-[omega]
turbulence model is used in RANS and simulations are run until
convergence is reached. In the DDES method, pressure implicit with
splitting of operator (PISO) algorithm was set for the pressure-velocity
coupling. The time step is chosen to comply with the solver requirements
for stability. Additionally, recommendations for time step size are
followed from [15]. The time step was set to 0.0001 s, and the CFL
number for this time step was approximately 0.5.
5. Grid
The ICEM-CFD commercial grid generator software was utilized to
create the numerical grid, as seen in Fig. 2. A high quality
unstructured hexahedral grid was created following the recommendation
for grid generation from [15]. Flow regions with different gridding
requirement exist in the numerical simulation. The grid should follow
the recommendations as far as possible to be efficient for the DES
method. The mesh consists of O-grid and H-grid topologies. This allowed
for a finer grid close to the wind tunnel walls and model surface
because a no-slip boundary condition is used on those surfaces. Values
of [y.sup.+] are set low (below 1) near the ground, the barrier and the
vehicle surfaces. Refinement zones are created in critical areas, such
as separations and wakes. Only one grid was created for the numerical
simulation. However, the grid sensitivity was conducted in a prior
numerical simulation where only the vehicle was analyzed under steady
crosswind. Hence, a grid with 33 million elements was created, where the
smallest element size was 4 mm.
[FIGURE 2 OMITTED]
5.1. Assessment of the grid resolution
The resolution characteristics of the grid near the walls were
discussed above by reference to the grid spacing in wall units. In the
interior of the flow, the grid resolution can be assessed by comparing
the grid spacing [DELTA] to an estimate of the Kolmogorov length [eta].
The Kolmogorov length scale, Eq. (8), characterizes the length
scale of the dissipative motion.
[eta] = [([v.sup.3]/[epsilon]).sup.1/4], (8)
where [epsilon] is the dissipation rate and v = 1.511x[10.sup.-5]
[m.sup.2]/s is the kinematic viscosity.
Fig. 3 shows the vertical profile of the ratio [DELTA]/[eta]
between the vehicle and the barrier, along cuts through the shear layer
and in the region beyond the wall. A substantial part of the dissipation
is resolved where the grid spacing is 12 [eta]. One can see from Fig. 3
that this level of discretization was achieved with levels of
[DELTA]/[eta] < 12 over the entire domain.
Further support for the grid resolution is provided by the ratio
v/[v.sub.t], as seen in Fig. 4, which gives an indication of the ratio
of resolved and modeled contributions to the dissipation. When eddy
viscosity is larger, the RANS modeled contribution is larger. Higher
turbulence viscosity is located near the vehicle and represents the RANS
area, whereas lower turbulence viscosity represents the LES area. Fig. 4
shows the instantaneous turbulence viscosity between the vehicle and the
barrier. As one can see from Fig. 4, the lower region represents the
RANS region near the barrier, and the upper part represents the RANS
region of the vehicle. In-between is located the LES region.
[FIGURE 3 OMITTED]
[FIGURE 4 OMITTED]
Grid dependence study was not performed. Instead, the largest
feasible grid size was used, in this case hexahedral grid with 33
million elements. Since the size of the grid influences the numerical
results, only one numerical simulation with largest possible grid size
was performed. Instantaneous and turbulent flow is analyzed in the
present work. The boundary conditions are time-dependent and therefore
the data is instantaneous.
6. Boundary conditions
The dimensions of the computational domain were set to simulate
open domain, as seen in Fig. 1, b. The upstream inlet and outlet were
positioned at a distance of 0.5L upstream and L downstream from the
barrier, respectively. The domain height and width were 10H and 20H,
respectively. No-slip boundary conditions were applied on the vehicle
and the ground surfaces. Appropriate boundary layer and blockage ratio
were simulated. Velocity inlet boundary condition with a uniform
velocity profile was specified at the domain inlet; in this case the
upstream and the lateral sides of the domain. A small turbulence
intensity of 0.3% was imposed at the inlet, which corresponds to the
experimental case. A pressure outlet was applied at the domain outflow.
The inflow velocity was [u.sub.v] = 13 m/s and maximum crosswind
velocity was [u.sub.crosswind] = 4.73 m/s. The corresponding Reynolds
number based on the barrier height is [Re.sub.H] = 1.7 x [10.sup.5]. To
simulate transient gust wind, transient boundary conditions on the
upstream and the lateral sides of the domain are used. In addition to
the front velocity inlet, identical velocity inlet boundary conditions
are specified on the lateral sides of the domain. A slip wall boundary
condition was set at the top surface of the domain. A moving wall
boundary condition with 13 m/s streamwise velocity was set for the
ground and the barrier surfaces, whereas a stationary wall was set for
the vehicle surfaces. A moving grid was avoided with this boundary
configuration. No wall functions were used for boundary layer modeling.
6.1. Unsteady crosswind
A single strong gust wind was simulated. The crosswind scenario
simulates the experimental study done by [1]. A vehicle model is
propelled at a constant velocity through the wind tunnel exhaust. It
depicts the scenario represented in Fig. 5, a. The gust wind consists of
a jet flow and two mixing layers. The jet flow was modeled as a step
function where the smooth transitions represent the mixing layers, Fig.
5, b. The smooth transitions were modeled as cosine functions [9]. The
maximum crosswind velocity length was set to 5[L.sub.V], and the cosine
period to 1.5[L.sub.V]. The maximum crosswind velocity was set to
correspond to the 20[degrees] yaw angle of the incoming wind in respect
to the vehicle. The yaw angle value of 20[degrees] is considered the
most critical for vehicle safety. At the front inlet, the crosswind is
only a function of time, whereas at the lateral side inlets the
crosswind is a function of the time and space. The transient boundary
condition was introduced in the solver via user defined functions
(UDFs).
7. Results
The first 9000 time-steps are run with only headwind and were not
considered because these time-steps correspond to a transient period
when the flow is unsteady. This corresponds to one flow-through time,
i.e., the time needed for one particle to go through the entire
computational domain. Afterwards, gust wind was introduced in the
computational domain.
The components of the aerodynamic forces projected on the vehicle
axis are the drag in the streamwise direction, the side force in the
lateral direction, and the lift in the upward vertical direction.
Coherent structures of the flow are investigated by using the second
invariant of the velocity gradient, the Q-criterion, Eq. (10). The
visual inspection of the turbulence structures was done using the
iso-surfaces of the Q-criterion.
[FIGURE 5 OMITTED]
The definition of the Q-criterion [16]:
Q = [C.sub.Q] ([[OMEGA].sup.2]- [[??].sup.2]), (9)
where [C.sub.Q] = 0.5, [??] is the absolute value of the strain
rate, and [OMEGA] is the absolute value of the vorticity.
[MATHEMATICAL EXPRESSION NOT REPRODUCIBLE IN ASCII] (10)
[OMEGA] = [square root of (2[[OMEGA].sub.ij][[OMEGA].sub.ij])],
(11)
[MATHEMATICAL EXPRESSION NOT REPRODUCIBLE IN ASCII]. (12)
[MATHEMATICAL EXPRESSION NOT REPRODUCIBLE IN ASCII]. (13)
7.1. Transient crosswind
The present work focuses on understanding of the flow mechanisms,
which occur when the gust wind acts on the barrier and the sheltered
vehicle. A time-dependent gust wind was introduced as boundary data
after the headwind was first run through the computational domain. Fig.
6 shows the propagation of the gust wind through the computational
domain in four different time moments: 1. no crosswind, 2. crosswind
approaching the barrier, 3. crosswind approaching the vehicle, 4.
crosswind behind the barrier and the vehicle.
[FIGURE 6 OMITTED]
7.2. Numerical accuracy
Fig. 7 shows the agreement among the characteristic points of the
experimental measurements and the numerical data for no-barrier
scenario. The gust length in the numerical simulation is longer for two
vehicle lengths than the one from the experimental measurements, and
adjustments on the time axis were performed for the experimental data to
evaluate the agreement between the experimental data and the numerical
results. As one can see, present results are in good agreement with the
experimental measurements. In addition, Fig. 7 shows the side force
coefficient trace for the fixed yaw tests. The side force coefficient
value for the fixed yaw test was time-averaged. As one can see from Fig.
7, the side force coefficient for transient yaw test displays an
overshoot compared to the fix yaw test, which indicates the importance
of performing the transient yaw tests for vehicle crosswind sensitivity.
[FIGURE 7 OMITTED]
7.3. Aerodynamic forces and moments
The axis system for the force and moment coefficients is the center
axis of the vehicle model, where a positive side force occurs with
positive wind loading and a positive yawing moment occurs when the nose
of the model turns leeward. The non-dimensional coefficients, [c.sub.F]
and [c.sub.M], for the forces and the moments, are defined in
[FIGURE 8 OMITTED]
[FIGURE 9 OMITTED]
Eq. (14) and Eq. (15):
[c.sub.F] = F/qA, (14)
[c.sub.M] = M/qA, (15)
where q is the dynamic pressure of the incoming wind, Eq. (16):
q = 1/2 [rho][u.sup.2.sub.R], (16)
where F is the force (drag, lift, and side force), M is the moment
(roll, pitch, and yaw), [rho] is the density of air, A is the frontal
area of the vehicle model, and [u.sub.R] is the resultant velocity.
The non-dimensional pressure coefficient:
[c.sub.p] = p - [p.sub.[infinity]]/q, (17)
where [p.sub.[infinity]] is the pressure in the free-stream.
Figs. 8-9 show the aerodynamic force and moment coefficients on the
vehicle with and without the barrier, respectively. As one can see, all
the aerodynamic loads are lower when the barrier is introduced except
for the pitch moment coefficient. Vehicle crosswind stability is
influenced mainly by the yaw moment coefficient and the side force
coefficient. Hence, vehicle crosswind stability is improved
significantly with the barrier shelter.
7.4. Visualization of the turbulent structures and pressure mapping
Fig. 10 shows the vehicle position relative to the gust wind.
Position 1 represents the vehicle at the beginning of the gust wind.
Position 2 represents the start of the vehicle experiencing the maximum
velocity of the gust wind. Position 3 represents the middle of the gust
wind, position 4 represents the end of the gust wind maximum velocity,
at the position 5 the vehicle starts to exit the gust wind, and at the
position 6 the vehicle completely exits the gust wind. Figs. 11-14 show
the visual representation of the large-scale flow structures using the
iso-surfaces of the Q-criterion colored by CFL value. The Q-criterion
value in this work is 100 000 [s.sup.-2]. One can see from Figs. 11-14
that the flow is complex and unstable. Also, the flow is remarkably
changed with the wind barrier introduction. There are dominant and well
defined coherent vortices originating from the front surface of the
vehicle at positions 3, 4 and 5 when there is no wind barrier shelter.
These structures are inclined in the direction of the resultant
direction of the gust wind. This is not the case for the wind barrier
scenario where prevalent vortical structures are small and parallel to
the vehicle traveling direction for the entire duration of the gust
wind. In particular, the flow structures of the sheltered vehicle during
the gust wind are similar to that before the sheltered vehicle entrance
into the gust. Also, one can notice the additional vortices formed
between the vehicle and the barrier.
[FIGURE 10 OMITTED]
[FIGURE 11 OMITTED]
[FIGURE 12 OMITTED]
[FIGURE 13 OMITTED]
[FIGURE 14 OMITTED]
Figs. 13-14 show the pressure distribution on the vehicle surfaces.
The pressure distribution indicates the influence of the vortices on the
vehicle's aerodynamic loads. As one can see, pressure coefficients
are much lower for the case without the barrier shelter and consequently
vehicle experiences larger aerodynamic loads. Also, one can notice in
Fig. 14 that at the positions 5 and 6 higher values of the negative
pressure coefficient are displayed; however, these values occur for a
very short time and on a small area of the vehicle surface. In addition,
one can see from Fig.14 that at the position 4, the large negative
pressure coefficient values on the vehicle's leeward surface
correspond to the side force coefficient and yaw moment coefficient peak
values. However, these peak values are still lower than those for the
case without the wind barrier.
8. Conclusion
In this work, a geometrically accurate three-dimensional wind
barrier model was used in the numerical study. An advanced aerodynamic
CFD simulation, DES, was utilized in analyzing the transient crosswind
scenario. Time-dependent boundary conditions were used to simulate the
gust wind propagation. One barrier configuration was analyzed. Hexahedra
grid with 33 million elements was used in the present numerical study.
Aerodynamic coefficients, pressure mapping, and flow visualization are
used to analyze the effects of the transient crosswind on the vehicle
aerodynamics. The goal of this research was to quantify and visualize
the barrier's shelter influence on the vehicle aerodynamics in
transient crosswind scenario.
LES is not feasible for higher Reynolds number flow which is the
case in this work and DES provides an accurate prediction of the dynamic
change in the aerodynamic coefficient. Therefore, DES was used to
investigate the influence of the barrier shelter on the vehicle aerody
namics in gust wind. The aim was to study the flow around the vehicle in
unsteady wind scenario and sheltered vehicle's aerodynamic load
development. Vortical structures are significantly changed with the wind
barrier application compared to the no-barrier case. Large and well
defined vortical structures develop when the vehicle is not sheltered by
the wind barrier. These vortical structures increase the negative
pressure on the vehicle's surface which promotes increase of the
vehicle aerodynamic loads. The present work shows that the vehicle
aerodynamic force and moment coefficients are much lower with the
barrier shelter except for the pitch moment coefficient. Moreover,
vehicle's crosswind stability is improved with introduction of the
wind barrier since vehicle's yaw moment coefficient and side force
coefficient are significantly lower.
The present work provides insight into the vortical structure
development which is result of the interaction among the vehicle, the
barrier and the gust wind. It offers deeper understanding of the flow
mechanism around the sheltered vehicle in the transient crosswind. In
particular, the dynamics of the coherent structures, as well as the flow
structure evolution and interaction with the wind barrier introduction
for the vehicle protection is presented. No prior work to date has
analyzed the time development of the vehicle aerodynamic loads behind
the geometrically accurate wind barrier in gust wind conditions.
http://dx.doi.Org/10.5755/j01.mech.21.3.8942
References
[1.] Chadwick, A. 1999. Crosswind Aerodynamics of Sport Utility
Vehicle, PhD Thesis, Cranfield University.
[2.] Packwood, A. 2000. Flow through porous fences in thick
boundary layers: comparisons between laboratory and numerical
experiments, J. Wind Eng. Ind. Aerodyn. 88: 75-90.
http://dx.doi.org/10.1016/S0167-6105(00)00025-8.
[3.] Fang, F.-M.; Wang, D.Y. 1997. On the flow around a vertical
porous fence, J. Wind Eng. Ind. Aerodyn. 67-68: 415-424.
http://dx.doi.org/10.1016/S0167-6105(97)00090-1.
[4.] Huang, L.-M.; Chan, H.C.; Lee, J.-T. 2012. A numerical study
on flow around nonuniform porous fences, J. Appl. Math. p.12.
http://dx.doi.org/10.1155/2012/268371
[5.] Bourdin, P.; Wilson, J.D. 2008. Windbreak aerodynamics: Is
Computational Fluid Dynamics Reliable?, Boundary Layer Meteorol. 126:
181-208 http://dx.doi.org/10.1007/s10546-007-9229-y.
[6.] Telenta, M.; Duhovnik, J.; Kosel, F.; Sajn, V. 2014. Numerical
and experimental study of the flow through a geometrically accurate
porous wind barrier model, J. Wind Eng. Ind. Aerodyn. 124: 99-108.
http://dx.doi.org/10.1016/jjweia.2013.11.010.
[7.] Telenta, M. 2014 Wind barrier motorway protection. LAP LAMBERT
Academic Publishing.
[8.] Hassan, H.; Krajnovic, S. 2007. DES of the flow around a
realistic bus model subjected to a side wind with 30 yaw angle, The
fifth IASME/WSEAS International Conference on Fluid Mechanics and
Aerodynamics, Athens.
[9.] Favre, T. 2011. Aerodynamics simulations of ground vehicles in
unsteady crosswind, Stockholm.
http://urn.kb.se/resolve?urn=um:nbn:se:kth:diva50242.
[10.] Krajnovic, S. 2008. Computer simulation of a train exiting a
tunnel through a varying crosswind, International journal of ailway
1(3): 99-105.
[11.] Krajnovic, S.; Ringqvist, P.; Nakade, K.; Basara, B. 2012.
Large eddy simulation of the flow around a simplified train moving
through a crosswind flow, J. Wind Eng. Ind. Aerodyn. 110: 86-99
http://dx.doi.org/10.1016/jjweia.2012.07.001.
[12.] Sustersic, G.; Prebil, I.; Ambroz, M. 2014. The snaking
stability of passenger cars with light cargo trailers, Strojniski
vestnik-Journal of Mechanical Engineering 60: 539-548.
http://dx.doi.org/10.5545/sv-jme.2014.1690.
[13.] Yogini P. 2010. Numerical investigation of flow past a
circular cylinder and in a staggered tube bundle using various
turbulence models, Lappeenranta
[14.] Spalart, P.; Deck, S.M.S.; Squires, K.D. 2006. New version of
detached-eddy simulation, Resistant to Ambiguous Grid Densities
Theoritical and Computational Fluid Dynamics 20: 181-195.
http://dx.doi.org/DOI 10.1007/s00162-006-0015-0.
[15.] Spalart, P. 2000. Strategies for turbulence modelling and
simulations, Int. J. Heat Fluid Flow. 21: 252-263.
http://dx.doi.org/10.1016/S0142-727X(00)00007-2.
[16.] Cucitore, R.; Quadrio, M.; Baron, A. 1999. On the
effectiveness and limitations of local criteria for the identification
of a vortex, Eur. J. Mech. B/Fluids. 18(2): 261-282.
http://dx.doi.org/10.1016/S0997-7546(99)80026-0.
Received December 17, 2014
Accepted March 20, 2015
Marijo Telenta *, Matjaz Subelj **, Joze Tavcar ***, Jozef Duhovnik
**** LECAD, Faculty of Mechanical Engineering, University of Ljubljana,
Slovenia, E-mail: *
[email protected] lj.si;
** matjaz.
[email protected]. uni-lj.si; *** joze.
[email protected].
uni-lj.si; ****
[email protected]. uni-lj.si