Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/262696697

Automatic Mesh Generation for Full-Cycle CFD Modeling of IC Engines:


Application to the TCC Test Case

Conference Paper  in  SAE Technical Papers · April 2014


DOI: 10.4271/2014-01-1131

CITATIONS READS

10 1,545

4 authors, including:

Tommaso Lucchini Gianluca D’Errico


Politecnico di Milano Politecnico di Milano
129 PUBLICATIONS   1,565 CITATIONS    133 PUBLICATIONS   1,857 CITATIONS   

SEE PROFILE SEE PROFILE

Roberto Torelli
Argonne National Laboratory
38 PUBLICATIONS   155 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

CFD Modeling of Internal Nozzle Flow View project

All content following this page was uploaded by Tommaso Lucchini on 30 May 2014.

The user has requested enhancement of the downloaded file.


2014-01-1131

Automatic mesh generation for full-cycle CFD modeling of


IC engines: application to the TCC test case

Author, co-author list (Do NOT enter this information. It will be pulled from participant tab
in MyTechZone)
Affiliation (Do NOT enter this information. It will be pulled from participant tab in MyTechZone)

c 2014 SAE International


Copyright

ABSTRACT Introduction
The definition of a robust methodology to perform a Nowadays, multi-dimensional models are widely em-
full-cycle CFD simulation of IC engines requires as ployed as design tools for internal combustion en-
first step the availability of a reliable grid generation gines. Depending on application and objectives, dif-
tool, which does not only have to guarantee a high ferent simulation types are generally performed.
quality mesh but also has to prove to be efficient in Steady-state conditions are modeled when new in-
terms of required time. In this work the authors dis- let port configurations have to be evaluated [3, 4]
cuss a novel approach entirely based on the Open- and the main interest is for flow coefficients and in-
FOAM technology, in which the available 3D grid gen- dexes of charge motions, like tumble or swirl num-
erator was employed to automatically create meshes bers. When fuel-air mixing or combustion processes
containing hexahedra and split-hexahedra from tri- have to be studied, piston geometry is included in
angulated surface geometries in Stereolithography the computational domain and mesh motion is also
(STL) format. The possibility to introduce local re- needed [5, 6]. Depending on the engine type, a por-
finements and boundary layers makes this tool suit- tion of the cycle or the entire one are simulated. For
able for IC engine simulations. Grids are sequentially example, in Diesel engines with axi-symmetric pis-
generated at target crank angles which are automati- ton bowl only compression, combustion and expan-
cally determined depending on user specified settings sion phases are considered with flow field imposed
such as maximum mesh validity interval and quality at intake valve closure (IVC) time from measured or
parameters like non-orthogonality, skewness and as- computed swirl ratio at steady-state conditions [7]. To
pect ratio. This ensures high quality grids for the en- simulate combustion or fuel-air mixture formation pro-
tire cycle and requires a very reduced amount of user cesses in SI engines, the full-cycle or, at least, the
time. The proposed approach has been introduced intake stroke needs to be included due to the higher
into the Lib-ICE code, which is a set of libraries and complexity of the in-cylinder charge motions that are
solvers based on the OpenFOAM technology and de- generated by the interaction between the incoming
veloped by the authors for IC engine modeling. Ex- air jet and cylinder walls [8, 9, 10]. Currently, most
perimental validation was carried out by simulating of the simulations are carried out by using the RANS
the full cycle in the so-called TCC (Transparent Com- method with the standard k − ε model and reliable ap-
bustion Chamber) engine [1, 2], whose experimen- proaches are available to describe both fuel-air mix-
tal data are available through the Engine Combus- ing and combustion [11, 12, 13, 14, 15, 16, 17]. In-
tion Network database (ECN). In particular, a detailed crease of CPU performance and massive paralleliza-
comparison between computed and experimental in- tion makes possible to achieve results in a reason-
cylinder pressure, turbulence intensity distribution and able amount of time. In particular, full-cycle simula-
velocity field was performed so that it was possible to tions are more and more necessary because most of
assess the requirements in terms of minimum mesh the engine design process is focused on spray tar-
size and numerical method accuracy to be employed geting and combustion optimization, due to the need
with the proposed methodology. to reduce both pollutant emissions and fuel consump-
tion. While case setup and run are handled almost au-
tomatically with limited user operation, the main bot-
tleneck is still represented by the mesh generation OpenFOAM technology
R and implemented into the
stage which requires a significant amount of time and LibICE code. Such technique was then incorporated
experience for geometry processing/cleaning and to in the methodology developed by the authors over
build a high-quality grid. This aspect is further wors- the years for full-cycle engine simulations [27, 28, 8],
ened by modern engine combustion chambers due where the entire cycle is simulated by using a multi-
to the complex shapes of valves, piston bowl, ports ple number of deforming grids, each one valid within
and cylinder head. To generate a grid, conventional a certain crank angle interval. In the proposed ap-
programs offer a pre-processing interface where it is proach, the user has only to provide the combustion
possible to import the CAD geometry in surface or vol- chamber geometry in surface file format with piston
ume format. Afterwards, the user manually cleans the at top dead center (TDC) and valves at minimum ar-
geometry, creates several volume blocks inside it and bitrary lift. Surface points are then moved to posi-
generates a proper mesh for each block. Once the tions corresponding to the start time of the simula-
first template grid is done, specific tools are available tion and a Cartesian, body-fitted grid-generator cre-
to create the entire necessary set to complete the full- ates the first mesh which is automatically moved until
cycle simulation. Due to the large involvement of man- quality parameters are satisfied or maximum validity
ual user work in mesh generation stage, several criti- interval is covered. At this stage, a new grid is gen-
cal issues might arise that negatively affect the quality erated and both mesh motion and automatic mesh
of the results. Among them, we can mention the capa- generation steps are sequentially performed until the
bility to maintain similar mesh structure, size and qual- entire set of grids is created. The proposed approach
ity for the different configurations to be tested. Possi- employs the utility available in the OpenFOAM code,
ble, fully automatic alternative solutions exist. Initially, called snappyHexMesh, generating automatically 3D
semi-automatic methods were proposed in [18] and meshes containing hexahedra and split-hexahedra
currently applied up to these days. Automatic tetrahe- from triangulated surface geometries in Stereolithog-
dral mesh generation was employed in [19, 20] using raphy (STL) format. Advantages of such utility are rep-
the Delaunay algorithm. Such approach is very fast resented by the possibility to insert boundary layers
and allows to generate high-quality grids. However, it on wall surfaces and local refinements in regions of in-
has several drawbacks related to the numerical diffu- terest. snappyHexMesh allows grid generation in paral-
sion induced by a non flow-oriented mesh structure, lel with a consequent reduction of the pre-processing
very high number of cells, computational efficiency time required.
and difficulty in creating boundary layers on valve wall
boundaries. Recent examples of the use of fully tetra- To test the proposed technique for mesh generation
hedral grids for IC engine simulations can be found and handling, an entire engine cycle was simulated
in [21, 22]. An alternative approach is represented by for the TCC engine. The choice of such geometry
Cartesian, cut-cell grids as it was described in detail was justified by several aspects. First, geometry and
in [23]. Such methodology has been successfully ap- experimental data are publicly available through the
plied over the years to simulate gas exchange, fuel-air Engine Combustion Network database, making such
mixing and combustion both in Diesel and GDI en- experiment a common basis for comparing different
gines [24, 25, 26]. There, the engine mesh is gen- CFD approaches for mesh management and turbu-
erated from the surface geometry and a Cartesian lence modeling. Furthermore, the layout of the valves
mesh. Cells outside the geometry are removed and is quite critical when Cartesian-based grids are em-
intersecting cells are cut to fit to the boundary geom- ployed, since the flow enters into the cylinder with ap-
etry by introducing additional faces. To capture most proximately a 45◦ orientation with respect to the mesh
of the geometry details of ports, valves and combus- structure. Hence, an accurate numerical setup needs
tion chamber, local refinement is also included close to be defined to properly predict the gas flow. Finally,
to the boundaries of interest. Cartesian cut-cell grids a large amount of experimental data of in-cylinder ve-
are computationally very efficient and also very accu- locity field and turbulence is available, making a com-
rate because of their very low non-orthogonality. How- prehensive validation of the proposed methodology
ever, prediction of charge motions can be negatively possible.
affected by the absence of boundary layers on the
valve wall boundaries and also numerical interpola-
tion errors might arise due to the need to regenerate Mesh management
the computational mesh at each time-step. This last
aspect also introduces a non-negligible computational In the proposed approach for CFD simulation of IC
overhead. engines a multiple number of meshes is employed, so
that each mesh is valid in a certain crank angle inter-
In this paper, a novel approach for automatic gen- val and during it the grid points are moved and the grid
eration of engine grids was developed using the topology is eventually changed. Consistent interpola-

2
tion of the computed flow field from one mesh to the From combustion chamber surface to first mesh
next one is performed by means of a second-order, Template Surface at
First mesh
inverse distance weighting method (with escapes for surface at simulation start
exact hit) [29]. TDC time

Mesh Motion

An automatic mesh motion technique was developed


to accommodate the displacement of internal grid
points according to the prescribed boundary motion
[30, 27]. The Laplace equation is solved with the Fi-
nite Volume method for the cell centers velocity field Figure 1: Generation of the first mesh from surface geome-
uc with constant or variable diffusivity γ: try of the combustion chamber.
∇ · (γ∇uc ) = 0 (1)
Automatic mesh generation for the full-cycle
The grid point velocity field up is computed by extrap-
Initial mesh at curr = 0
olation from uc and used to modify the point positions: Move mesh for
Crank angle 0
xnew = xold + up ∆t, (2)
curr = curr
0= curr
Eq. 1 is solved with boundary conditions represented
by the prescribed boundary motion. To further pre- Mesh YES
serve the mesh quality during motion it is also pos- quality and
sible to specify mesh-motion for an arbitrary number Generate a new duration
mesh with satisfied?
of points in pre-defined regions before solving the snappyHexMesh
motion equation. The developed mesh motion tech-
nique is flexible with respect to the mesh structure, NO
supporting unstructured polyhedral cells of arbitrary
shapes. This is quite important when direct-injection Move surface NO
curr =
engines have to be simulated, since their complex ge- geometry to current
end ?
crank angle curr
ometry usually require combination of hexahedra and
tetrahedra in regions where high deformations takes YES
place.
End of mesh
generation process

Automatic mesh generation


Figure 2: Automatic mesh generation process.

Figs. 1 - 2 summarize the methodology developed


in this work by the authors for automatic mesh be used for gas exchange, fuel-air mixing and com-
generation. The process is intended to ensure high bustion simulations. Initially, the user has to provide
quality grids according to the following user specified a cleaned surface of the combustion chamber where
parameters: all the boundaries of interest (piston, valve poppet,
valve bottom, liner, cylinder head, valve stem, inlet,
- maximum validity interval for each mesh, to outlet, . . . ) are identified with a proper name. Only
avoid excessive stretching of the cells and loss stereolithographic formats are supported. Piston must
of resolution; be located at TDC and valves at minimum lift, usu-
- maximum allowed mesh non-orthogonality and ally ranging in the 0.1 - 0.25 mm interval. A proper
skewness values, to increase stability and simu- utility moves automatically both piston and valves at
lation accuracy [31]; positions corresponding to the crank angle where the
simulation is started. At this stage, the first mesh is
- Topological and geometrical validity of the mesh
automatically generated and then it is moved using
[32].
the automatic mesh motion technique and solving the
Laplace equation with a fixed, user-specified time-
At the end of the process, an entire set of meshes step ranging from 0.1 to 0.25 CAD. The mesh is con-
satisfying the listed criteria will be created and can tinuously deformed until duration, quality and validity

3
criteria are satisfied. As soon as one of these fails, a) Combustion chamber surface
the combustion chamber surface geometry is moved and background mesh
to the current crank angle and a new mesh is gener-
ated. This process is sequentially performed until the
end of the simulation is reached.

Mesh generation tool

The snappyHexMesh method available in Open-


FOAM differs from the traditional way of doing pre-
processing for CFD. This method uses automatic pro-
cedure to create orthogonal hexahedral mesh either
around or inside a given geometry surface, which has
to be provided in stereolithographic format. In prin-
ciple this method enables fast and robust meshing
of complex geometries. In this work, snappyHexMesh b) Castellated mesh
was applied to generate engine meshes and here
its operation is shortly described and summarized in
Figs. 3(a) - (c).

Initially, a block-structured grid has to be provided and


its size represents the initial mesh density. On the ba-
sis of the geometry to be meshed and user-specified
settings, the so-called castellated mesh is created by
removing all the cells outside the combustion cham-
ber, as it can be seen in Fig. 3(b). Close to bound-
aries or in regions of particular interest, the castel-
lated mesh can be further refined either for a better
matching of the real geometry or to properly predict
c) Body-fitted mesh
phenomena of interest such as incoming gas flow,
flame kernel growth or air-fuel mixture formation. An
iterative procedure, controlled by maximum specified
non-orthogonality and skewness values, will be then
used to morph the castellated mesh to the combus-
tion chamber surface. The final result of the proce-
dure is shown in Fig. 3(c), displaying a body fitted
mesh conforming to surface boundaries and properly
accounting for the main geometry details including
sharp edges.

The optional possibility to insert boundary layers on Figure 3: Application of the snappyHexMesh tool to engine
surfaces of interest was also exploited for the valve mesh generation.
boundaries in order to better predict the flow entering
into the cylinder. Details of the mesh generation pro-
cess in the valve region, from castellated to body fitted snappyHexMesh runs in parallel and operates a load
mesh with boundary layers are illustrated in Fig. 4. balancing step every iteration. Both these aspects
make mesh generation process very fast and guar-
Since it is not possible to model the contact between antee an optimal domain decomposition for the sub-
valve poppet and cylinder head, the following proce- sequent gas flow simulations.
dure was used: the valve is considered to be closed
when its lift is lower than a minimum value (range is
0.1 - 0.25 mm). At that points, a new mesh is created
and baffles are automatically introduced to physically
separate the cylinder and port domains.

4
a) Castellated mesh the same face (PQ). Non-orthogonality affects the dis-
cretization accuracy of the diffusion term in transport
equations [31, 34]. In industrial CFD simulations non-
orthogonal meshes are commonly used to account for
complex geometry features. However, if α > 80 the
non-orthogonality is severe requiring to limit or dis-
card the non-orthogonal component of the diffusion
term and to increase the number of non-orthogonal
correctors [31]. α ≥ 90 is an index of mesh invalidity
since this happens with degenerate cells.

b) Body-fitted mesh The grid skewness error reduces the accuracy of face
integrals to the first order [31] when the interpolated
face value does not lie in the its center as shown in
Figure 5(b). The skewness parameter for a face is de-
fined as:
kmk
skewness = (3)
kPQk
Where the vector m is the distance between the face
center and the intersection between the face area and
the vector PQ. As for nonorthogonality, skew meshes
are rather common in IC engine simulations mainly
c) Body-fitted mesh when tetrahedra are used. Skewness can be consid-
with boundary layers ered severe when it is higher than 4. However, values
lower than 10 can be still considered acceptable.

f S
P m
S fi d
α Q
P d Q
f

(a) (b)
Figure 4: Application of the snappyHexMesh tool to generate
body-fitted meshes with boundary layers in the valve region. Figure 5: Examples of a non orthogonal (a) and a skew
mesh (b).

Mesh quality indexes


Results and discussion
Using a finite volume method with deforming grids,
the computed solution is strongly affected by mesh Sandia TCC Engine
size and quality, which are continuously changing due
to domain volume variation induced by moving bound- The proposed approach was assessed and validated
aries. A significant work about mesh quality metrics by experiments carried out in a Transparent Combus-
was performed in the past [33] where different shape tion Chamber (TCC) engine with a two-valve head
and size quality indexes were defined, but they can- and a pancake combustion chamber which were
not be easily extended to polyhedral grids and are mainly adopted to perform experiments about the na-
only valid for tetrahedral and hexahedral cells. In [31], ture of stochastic flows in internal combustion engines
Jasak showed that discretization accuracy of the finite [1, 2, 35]. This aspect justified the simplified geometry
volume method depends on the mesh skewness and layout which was intended to generate a turbulent flow
non-orthogonality. in the cylinder by means of the interaction between
the incoming gas jet, piston and liner walls. Most of
An example of non-orthogonal mesh is displayed in the studies on this engine were used to assess and
Figure 5(a). The mesh non-orthogonality is defined develop sub-grid models for Large Eddy Simulations
for a face as the angle α between the face area vec- [36, 37, 38], while to the authors’ knowledge no RANS
tor (S) and the vector joining the cell centers sharing simulations were carried out on such geometry so far.

5
The engine layout is displayed in Figs. 6 (a) - (b), Ta-
ble 1 illustrates the main geometry data and valve lift
profiles and phasing are displayed in Fig. 7.

The overall optical access is maximized to allow ac-


quisition of three-dimensional in-cylinder flow field for
a proper investigation of near-wall, boundary layer
flows. Experimental data are acquired with optical
multi-dimensional high-speed diagnostics techniques.
In particular, macro- and micro-PIV techniques were
used for flow measurements, which can be set up to
cover the full 86 mm stroke of the engine. Further-
more, zoomed-in measurements allow spatial resolu-
tions below 50 mm. Data were recorded every 5 CAD
for 70 cycles, allowing a very detailed characteriza-
tion of the average in-cylinder flow-field and related Figure 7: Intake and exhaust valve lift profiles of the TCC
fluctuations. engine.

The corresponding mesh sizes are approximately 600


thousand and 1.5 million cells at TDC and BDC, re-
spectively. More details about the mesh structure are
displayed in Figs. 8 (a) - (b) where it is also possible
to see that the spark-plug geometry was included in
the computational domain.
Figure 6: Layout of the SANDIA TCC engine: (a) engine at
test-bench; (b) combustion chamber geometry. (a) (b)
Bore 92 mm
Stroke 86 mm
Connecting Rod Length 244 mm
Compression Ratio 10
IVO 359 ◦
IVC 592 ◦
EVC 364 ◦ Figure 8: Mesh structure used for the TCC engine simula-
EVO 131 ◦ tion: (a) inlet port and valve region; (b) cylinder symmetry
Speed 800 rpm plane.
Valve diameters 30 mm
Maximum valve lift 8.9 mm Grid generation process starts from EVO, where
the full-cycle simulation begins. Effects of non-
Table 1: Summary of engine geometry data and simulated orthogonality on the mesh generation process were
operating conditions. evaluated, by setting that during motion grids maintain
both their topological and geometrical validity, allow-
ing a maximum non-orthogonality of 90◦ and avoiding
Methodology assessment negative face areas. With such constrains, average
mesh duration is approximately 10 CAD and the cu-
The performances of the proposed approach for au- mulative mesh count as function of crank-angle is re-
tomatic mesh generation were first tested in terms ported in Fig. 9 where it is possible to see that mesh
of number of required grids and required CPU time. validity mainly depends on how the adopted finite-
Such investigation was carried out using the following volume algorithm is affected by cylinder mesh de-
parameters for the grid generation process: formation, valve motion and interaction between pis-
ton and valves. In particular, grid compression com-
- 2 mm mesh size inside cylinder and ports; bined with valve motion during exhaust stroke re-
quires a higher number of grids (33 vs 25) com-
- local refinement up to 0.125 mm close to cylinder
pared to the intake phase, where cylinder mesh is ex-
head, piston, liner and valve boundaries;
panded. Compression stroke requires less grids, due
- additional refinement box below the cylinder to the absence of valve motion. Mesh count drastically
head with a 1 mm size. increases during valve overlap around IVC since in

6
such conditions mesh validity is strongly worsened by IVC periods, while the number of grids required for in-
piston-valve interaction and high deformation of small take, compression and exhaust phases remain almost
cells on the top of the valves. Compression and ex- unchanged. This aspect can be drastically seen when
pansion phases requires much lower grids, due to the maximum mesh non-orthogonality is reduced from 75
absence of valve motion. to 70 degrees. Since charge motions and turbulence
are mainly generated during the intake stroke, intro-
Exhaust Intake Compression
stroke stroke stroke ducing more grids in the other parts of the cycle is
expected to increase the computational time but not
80
the quality of the results. For this reason, grid gener-
70 IVC ation process based on maintaining topological and
BDC geometrical validity of the mesh was adopted in this
60 IVO EVC
15 work.
Cumulative mesh count

50 Finally, Fig. 11 illustrates performance of the grid gen-


EVO
eration process when run on multiple CPU. The sin-
40 25
gle processor case was not considered due to the
30 large size of the meshes involved. Mesh generation
and simulations were run on an AMD 64 cores ma-
20 chine, with 4 processors and 4 GB RAM for each
33 core. For what concerns the scalability of the pro-
10
cess, it is possible to see that a good speed-up fac-
0
tor is achieved until 12 cores are used. With a higher
120 245 370 495 620 745 870 number of them, overheads introduced by the contin-
Crank angle [ ] uous need to redistribute the meshed domain across
Figure 9: Cumulative mesh count for the mesh generation the different processors compensates the increase of
process, limited only by topological and geometrical validity available computational resources. On 16 cores the
of each mesh. entire mesh generation process takes approximately
8 hours (including the required pre-processing) which
Exhaust Intake Compression
is significantly less than what is generally needed to
stroke stroke stroke
create manually just the first template mesh.
100
IVC 2.5
90
BDC
80
IVO EVC
Cumulative mesh count

70 2

60 EVO
Speed-up factor

50 1.5
40
Max n.o. = 90°
30
Max n.o. = 80° 1
20
Max n.o. = 75°
10
Max n.o. = 70° 0.5
0
120 245 370 495 620 745 870
Crank angle [ ]
0
Figure 10: Effect of maximum allowed non-orthogonality on 0 4 8 12 16 20
the mesh generation process.
Number of cores
Fig. 10 illustrates how the cumulative mesh count is Figure 11: Parallel performance of the mesh generation pro-
cess.
affected by maximum allowed non-orthogonality, with
tested values in the 70-90◦ range. When generated,
each mesh has an initial maximum non-orthogonality
of 60 degrees. As expected, when reducing the maxi-
mum allowed non-orthogonality during motion, the to-
tal number of required meshes grows. However, such
increase takes place mainly during valve overlap and

7
Experimental Validation the Cartesian grid orientation could be used to prop-
erly reproduce the in-cylinder flow field. This aspect is
Full-cycle simulations of the TCC engine were car- rather critical for the simulated engine, since the flow
ried out to validate the proposed approach. The sim- enters the cylinder with approximately a 45◦ orienta-
ulated domain is shown in Fig. 12 where it is possible tion with respect to the mesh structure. Under such
to see that the entire combustion chamber (including conditions, numerical diffusivity plays a big role and
spark-plug) and part of intake and exhaust manifolds might lead to a wrong prediction of flow field, charge
were included in the computational mesh. On the inlet motions and turbulence. To this end, three different
and outlet boundaries, experimental time-varying to- simulation setups were considered as displayed in
tal pressure and temperature profiles were imposed. Tab. 3, representing a combination of mesh sizes
In Fig. 13 it is possible to see the typical behavior of and numerical methods. Two different grids were used
pressure waves in a single cylinder engine operating which mainly differ in the mesh structure close to the
at non-fired conditions. The standard turbulence k − ε valves, as illustrated in Fig. 14 (a) - (b): the second
model was used with 5% turbulence intensity at the in- mesh has an additional refinement region close to
let and the integral length equal to 10% of the manifold the valves with a 0.25 mm size. This increases the
diameter. Coefficients used for the turbulence model number of cells, up to 1.3 at TDC and 1.7 millions
are listed in Tab. 2. at BDC. Numerical diffusivity is expected to have a
much higher influence on the results with the coarser
meshes than the fine meshes. To understand how
numerical methods influence the computed flow-field
and if the meshes are all robust enough to sup-
port high-order methods, both first and second-order
schemes for space discretization were tested on the
fine mesh.

(a) Coarse mesh (b) Fine mesh


Figure 12: Simulated computational domain.

Figure 14: Comparison between the structures adopted for


the two tested mesh configurations: (a) coarse mesh; (b)
fine mesh.

Case Mesh Numerical Method


1 Coarse First order
2 Fine First order
3 Fine Second order
Table 3: Simulation setup in terms of mesh size and numer-
ical schemes adopted for the TCC engine.

In order to achieve the convergence of results in terms


Figure 13: Evolution of pressure (solid) and temperature
(dashed) on the inlet (blue) and outlet (red) boundaries for
of computed velocity field in both manifolds and cylin-
an engine cycle. der for each crank angle, two engine cycles were sim-
ulated. The validity of the proposed setup in terms
Cµ C1 C2 C3 αh αk αε P rt of mesh to mesh interpolation, geometry discretiza-
0.09 1.44 1.92 -0.33 1 1 0.769 0.85 tion and imposed boundary conditions was verified by
comparing in Fig. 15 computed and experimental data
Table 2: k − ε model coefficients used in this work. of in-cylinder pressure profiles.

Validation was mainly carried out to understand the A proper matching was achieved, showing that
validity of the proposed setup in terms of boundary trapped cylinder mass is correctly predicted at the
conditions and number of used meshes and to see if end of the gas-exchange process. Furthermore, a

8
piston at maximum acceleration;

• 450 CAD: inlet valve at almost lift and piston at


maximum velocity;

• 490 CAD: inlet valve at mean lift (∼ 4.9 mm) and


piston at maximum deceleration;

• 540 CAD: inlet valve at mean lift (∼ 4.9 mm) and


piston at bottom dead center (zero velocity);

• 590 CAD: inlet valve closure time.


To validate the proposed methodology, a detailed
Figure 15: Comparison between computed and experimen- comparison between computed and experimental
tal cylinder pressure profiles for the full cycle under motored data of in-cylinder velocity and turbulence intensity
conditions. distributions was carried out for the selected crank
angles. Experimental instantaneous velocities were
recorded on two different windows, parallel to both
rather good agreement between experimental and
cylinder and valve axes. The first window provides de-
computed pressure values was achieved during the
tailed information about the incoming flow in the cylin-
intake stroke, meaning that losses in both inlet mani-
der. It has a 40.6 x 62.8 mm size, with a correspond-
fold and valves are properly reproduced.
ing grid of 52 x 80 points equally spaced. The second
measurement window is larger (74 x 115.7 mm) than
Once the setup was verified, comparisons between
the first one, but it has a lower resolution (same num-
computed and experimental data of in-cylinder flow
ber of points: 52 x 80) and provides an overview of the
fields were carried out. In particular, measured data
overall cylinder flow field. Interpolating computed data
were available in a 26 mm width x variable height win-
at exact measurement locations and accounting only
dow, located in a plane passing through the inlet valve
for projected values on the optical windows ensure
axis and parallel to the cylinder symmetry plane. Both
a consistent validation with experimental data. Here,
measured velocity vectors and magnitude were avail-
some details about the post-processing technique are
able as well as the distribution of the turbulence inten-
briefly described. Experimental velocity magnitude in
sity. Fig. 16 shows the exact location of the measure-
each point of the measurement plane is computed
ment window inside the cylinder.
from the average of its horizontal and vertical com-
ponents as follows:
v
u N !2 Nc
!2
1 u X c X
U (x, y, θ) = t Ux,i (θ) + Uy,i (θ)
Nc i=1 i=1
(4)
where Nc is the total number of cycles where velocity
data were recorded. For each experimental velocity
component, corresponding turbulence intensities are
estimated as:
Figure 16: Location of the optical window used for flow field
and turbulence measurements.
Nc
1 X 2
u′x (x, y, θ) = U x (x, y, θ) − Ux,i (x, y, θ) (5)
Effects of mesh size and numerical accuracy were Nc i=1
evaluated at different times during intake and com- Nc
pression stroke and flow field results were then com- 1 X 2
u′y (x, y, θ) = U y (x, y, θ) − Uy,i (x, y, θ) (6)
pared with experimental data. For what concerns the Nc i=1
intake phase, the following instants were considered:

• 370 CAD: beginning of intake phase where high where U x and U y are the average horizontal and ver-
velocities are expected because of small valve lift tical components of velocity, respectively. Experimen-
(∼ 1.5 mm); tal turbulence intensity is computed for u′x , u′y , esti-
mating the corresponding turbulent kinetic energy first
• 410 CAD: inlet valve at mean lift (∼ 4.9 mm) and and then computing the corresponding average turbu-

9
lence intensity u′exp as follows: of the inlet valve and his presence is properly re-
produced by all the tested configurations as it can
1 ′2 1 ′2 be seen in Figs. 17 (b) - (d) where also no signifi-
kexp = u + u (7)
2 x 2√y cant differences between used meshes and numerical
u′exp = k (8) setup were found. However, the predicted vortex size
is smaller than the experimental one and this might be
For a consistent comparison between experimental due to the choice of using a minimum valve lift of 0.25
and computed data, calculated velocity field is inter- mm, which might limit the air inflow to the cylinder at
polated at measurement locations with an inverse- the beginning of the inlet stroke.
distance weighted interpolation technique [29], then
its magnitude is computed taking only the projection During the first part of the intake stroke, the typi-
of the velocity vector on the measurement plane into cal gas jet structure is established at the valve en-
account. Computed turbulence intensity u′calc is esti- trance, in particular Fig. 18 illustrates that experimen-
mated from turbulent kinetic energy kcalc interpolated tally there is still not interaction between gas flow and
at measurement locations as follows: piston. The shape of jet is correctly predicted when
r the fine mesh is used, either with first or second order
′ 2 schemes as it can be seen in Figs. 18(c) - (d). In-
ucalc = kcalc (9)
3 stead, Fig. 18(b) shows that numerical diffusion due
to either first order schemes or coarse mesh induces
a non-physical deviation of the jet towards the cylin-
370 CAD - Velocity magnitude der liner in the horizontal direction. However, accurate
(a) Exp. (b) Coarse Mesh numerical schemes are necessary to properly predict
jet penetration, while first order schemes makes the
jet larger and shorter.

410 CAD - Velocity magnitude


(a) Exp. (b) Coarse Mesh

(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)

(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)

Figure 17: Comparison between computed and experimen-


tal velocity magnitude distribution at 370 CAD. Units are in
[m/s], scale is 0 (blue) - 11 m/s (red).

Figs. 17 (a) - (d) report a comparison between com-


puted and experimental velocity distributions at 370 Figure 18: Comparison between computed and experimen-
CAD. At the beginning of the intake stroke, there is tal velocity magnitude distribution at 410 CAD. Units are in
still significant vortex located below the exhaust valve [m/s], scale is 0 (blue) - 70 m/s (red).
originating from the exhaust stroke as it can be seen
in Fig. 17(a). None of the simulated configurations A comparison between velocity vector distribution in
predicts correctly the intensity and shape of that vor- the measurement plane is shown in Fig. 19(a) - (d).
tex. because of the rather coarse mesh below the ex- In particular, when comparing Figs. 19(a) and (d) it
haust valve. A small vortex also appears on the left is possible to see that also the slight deviation of the

10
jet due to interaction between the incoming flow and 410 CAD - Turbulence intensity
the spark-plug is properly described and this demon- (a) Exp. (b) Coarse Mesh
strates the need to include such detail in the compu-
tational mesh.

410 CAD - Velocity vectors


(a) Exp. (b) Coarse Mesh

(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)

(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)

Figure 20: Comparison between computed and experimen-


tal turbulence intensity distribution at 410 CAD. Units are in
[m/s], scale is 0 (blue) - 30 m/s (red).

450 CAD - Velocity magnitude


(a) Exp. (b) Coarse Mesh
Figure 19: Comparison between computed and experimen-
tal velocity vectors at 410 CAD.

For what concerns the turbulence levels inside the


cylinder, in Fig. 20 experimental intensity looks at
this time much higher than computed data. Only the
fine mesh with second order schemes seems to be
able to predict non negligible turbulence levels inside
the cylinder with a distribution similar to experimen-
tal data. However, at 410 CAD predicted in-cylinder
turbulence might be also affected by the residual flow
originated during the exhaust phase where a coarse (c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)
mesh was used.

Figs. 21(a) - (d) report a comparison between com-


puted and experimental velocity magnitude at 450
CAD, where piston velocity is maximum and for this
reason a significant amount of air is sucked inside
the cylinder. Jet penetration is rather well predicted
by fine meshes, with second order schemes being
also able to better reproduce the velocity distribution
inside the jet. For what concerns the jet shape, ex-
perimental data show a slight vertical deviation at his
periphery which is not properly estimated during the Figure 21: Comparison between computed and experimen-
simulations. tal velocity magnitude distribution at 450 CAD. Units are in
[m/s], scale is 0 (blue) - 60 m/s (red).

11
To better understand such difference, Fig. 22 com- ture below the valve might be the reason for the differ-
pares experimental and computed velocity vectors ences between experimental and predicted shapes of
with second order schemes and fine mesh. the gas jets at 450 CAD.
450 CAD - Velocity vectors 490 CAD - Velocity magnitude
(a) Exp. (b) Fine mesh (2nd ord.) (a) Exp. (b) Coarse Mesh

(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)

Figure 22: Comparison between computed and experimen-


tal velocity vectors at 450 CAD: (a) Experimental data, (b)
Fine mesh with second order schemes.

450 CAD - Turbulence intensity


(a) Exp. (b) Fine mesh (2nd ord.)

Figure 24: Comparison between computed and experimen-


tal velocity magnitude distribution at 490 CAD. Units are in
[m/s], scale is 0 (blue) - 40 m/s (red).

490 CAD - Velocity vectors


(a) Exp. (b) Fine mesh (2nd ord.)

Figure 23: Comparison between computed and experimen-


tal turbulence intensity distribution at 450 CAD: (a) Exper-
imental data, (b) Fine mesh with second order schemes.
Units are in [m/s], scale is 0 (blue) - 20 m/s (red).

From Fig. 22(a) it is possible to see that, experimen-


tally, the gas jet deviation is mainly due to the circu-
lar vortex originated at the bottom of the valve and
moving downwards. Such vortex also exists in simu-
lations, but it is more elongated and for this reason it
is not able to properly deviate the incoming cylinder
flow. In the authors’ opinion this discrepancy can be
affected by two different aspects: the first is the grid
size below the valve, being 1 mm and probably not Figure 25: Comparison between computed and experimen-
refined enough to reproduce the development of the tal velocity vectors at 490 CAD: (a) Experimental data, (b)
vortex originated by the shear flow at valve exit. Turbu- Fine mesh with second order schemes.
lence model might also play a role, producing a limited
viscosity that makes the jet more straight than what When moving towards the second part of the intake
effectively it is. However, when looking to predicted stroke at 490 CAD, it is possible to see in Fig. 24(a)
turbulence levels in Fig. 23, they are comparable with that the jet enlarges due to air entrainment. Due to in-
experimental values and for this reason mesh struc- teraction with the vortex created below the valve, the

12
jet maintains its round distortion. All the tested setup located in the bottom part of the cylinder and it is also
were able to qualitatively describe the shape of the enhanced by the incoming air jet. At BDC, Fig. 27(d)
jet but the coarse grid predicts a very low penetra- shows that such vortex is almost located in the center
tion, while the fine mesh with first order schemes esti- of the cylinder and it has a major influence in the ve-
mates a reduced jet enlargement downwards. Results locity distribution.
with fine mesh and second order schemes shown
in Fig. 24(d) are in acceptable agreement with ex-
perimental data. When looking at Fig. 25, compar- (a) 450 CAD (b) 480 CAD
ing experimental velocity vectors and computed ones
with second-order schemes it is possible to see that
now the round vortex located the bottom-right part of
the window is well-established in simulations while it
starts to decay in measured data.

For what concerns the turbulence distribution in-


side the cylinder, an interesting aspect is shown in
Fig. 26(a). The large amount of turbulence found in
the bulk of the cylinder seems to be mainly originated
by the shear flow that exists the valve outlet. Turbu-
lent kinetic energy originates at the cylinder head and
valve edges, diffuses into the air jet so that a large
region of high turbulence intensity is created. In simu- (c) 510 CAD (d) 540 CAD
lations, again mesh structure seems to be responsible
for a not correct prediction of the turbulence distribu-
tion as it can be seen in Fig. 26(b). In particular, a fine
mesh in the top of the valve generates a significant
amount of turbulence while coarse mesh in the bot-
tom of the valve is not able to properly capture velocity
gradients there and the related turbulence generation.

490 CAD - Turbulence intensity


(a) Exp. (b) Fine mesh (2nd ord.)

Figure 27: Experimental evolution of in-cylinder velocity field


in the 450-540 CAD interval.

Mesh size and turbulence model are expected to play


a big role in the prediction of the in-cylinder flow at
BDC due to the presence of the large vortex inside
the cylinder. To this end, a comparison between com-
Figure 26: Comparison between computed and experimen-
puted and experimental velocity magnitude contours
tal turbulence intensity distribution at 490 CAD: (a) Exper-
imental data, (b) Fine mesh with second order schemes.
is displayed in Fig. 28. Care is necessary for a de-
Units are in [m/s], scale is 0 (blue) - 14 m/s (red). tailed analysis of the computed data since mislead-
ing conclusions can be drawn: at a first sight results
achieved using a coarse grid with 1st schemes look
To better understand experimental results at BDC,
much better than the others. To better understand the
Fig. 27 displays the measured velocity field evolution
differences between the numerical setup used, the
during the second part of the intake stroke, in the
development of the in-cylinder flow on a larger window
450-540 CAD range. Despite not completely visible in
is illustrated in Fig. 29 in the 450-540 CAD range. In
the images, the jet that enters from the other side of
particular, velocity contours are shown for the coarse
the intake valve has a significant influence on the in-
mesh with first order schemes and the fine mesh with
cylinder flow. It strongly interact with the cylinder liner
second order schemes.
and then it moves towards the piston and flows over
it already at 450 CAD (Fig. 27(a)). This is the condi-
tion for the creation of a sort of vortex which remains

13
540 CAD - Velocity magnitude (a) 450 CAD
(a) Exp. (b) Coarse Mesh Coarse mesh (1st ord.) Fine mesh (2nd ord.)

(b) 480 CAD


(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)
Coarse mesh (1st ord.) Fine mesh (2nd ord.)

Figure 28: Comparison between computed and experimen- (c) 510 CAD
tal velocity magnitude distribution at 540 CAD. Units are in Coarse mesh (1st ord.) Fine mesh (2nd ord.)
[m/s], scale is 0 (blue) - 7 m/s (red).

Compared to the fine mesh, the coarse mesh has


lower levels of turbulence and higher numerical diffu-
sivity. This last aspect is in principle responsible for a
reduced jet penetration as it is clear from Figs. 29(a)-
(b). At 510 CAD, there is a high difference in terms
of velocity distribution on the piston. In particular, the
flow in the fine mesh has a higher penetration and
moves fast towards the cylinder head, producing a
larger vortex. In the coarse mesh, instead, a vortex (d) 540 CAD
Coarse mesh (1st ord.) Fine mesh (2nd ord.)
is created in the bottom part of the cylinder and its
shape at BDC is very similar to the experimental one.
However, its generation and evolution is mainly af-
fected by numerical diffusivity and reduced jet pen-
etration that increases the vortex strength. In particu-
lar, the first one diffuses the velocity from the piston
and the liner to the cylinder bulk. For what concerns
the second order grid, again the main reason for a
non correct estimation of the velocity field seems to
be only due to the wrong distribution of the in-cylinder
turbulence intensity and related diffusivity because of
the coarse resolution adopted below the inlet valve. Figure 29: Comparison between computed velocity con-
For sake of completeness and to further clarify what tours using coarse mesh with 1st order schemes (left) and
was found so far, Fig. 30(a)-(d) finally compares ex- fine mesh with 2nd order schemes during the second part
of the intake stroke (450-540 CAD).
perimental and computed turbulence intensity distri-
butions for the three different simulation setup used.

Fig. 30(d) shows that only the fine mesh with second levels of turbulence which were experimentally found,
order schemes is able to properly reproduce the same even if with a wrong distribution due to the non cor-

14
rect position of the predicted vortex at BDC. First or- 600 CAD - Velocity magnitude
der methods instead underestimate the turbulence in- (a) Exp. (b) Coarse Mesh
tensity and this further explain how numerical diffusiv-
ity can lead to acceptable results but with the wrong
physics.

540 CAD - Turbulence intensity


(a) Exp. (b) Coarse Mesh

(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)

(c) Fine mesh (1st ord.) (d) Fine mesh (2nd ord.)

Figure 31: Comparison between computed and experimen-


tal velocity magnitude distribution at 600 CAD. Units are in
[m/s], scale is 0 (blue) - 5 m/s (red).

600 CAD - Velocity vectors


(a) Exp. (b) Fine mesh (2nd ord.)

Figure 30: Comparison between computed and experimen-


tal turbulence intensity distribution at 540 CAD. Units are in
[m/s], scale is 0 (blue) - 5 m/s (red).

From BDC to IVC time the main vortex inside the


cylinder evolves and, in Fig. 31 the computed flow
field immediately after the valve closure time is il-
lustrated. The vortex is still intense and fine mesh
with 2nd order schemes produces the best agree-
ment with experimental data. This is mainly due to the
better capability of such grid to reproduce flow and
Figure 32: Comparison between computed and experimen-
turbulence details in combination, predicted velocity
tal velocity vectors at 600 CAD: (a) Experimental data, (b)
magnitudes are higher than experimental ones. The
Fine mesh with second order schemes.
coarse grid with first order schemes, producing the
best flow agreement at BDC, predicts a new weaker
vortex due to numerical diffusivity that also affects the Fig. 32 provides a comparison between experimen-
results of the fine mesh in Fig. 31(c). For a proper tal and computed velocity fields with the best setup
prediction of both fuel-air mixing and combustion pro- (fine mesh and second order schemes). A rather good
cesses, a good estimation of both flow field and turbu- agreement was achieved at this point: the location of
lence at IVC time is fundamental. Within this context, the vortex is correctly placed in the bottom-right part
it is quite clear from computed results that the fine of the window. Furthermore, the simulation correctly
mesh with second order schemes produces the best predicts an almost horizontal gas flow in the region
agreement with experimental data. For this reason, immediately below the cylinder head. Finally, also the
from now, only results obtained with such mesh will vortex intensity is properly described when looking at
be analyzed and compared with experimental data. the size of the velocity vectors. However, Fig. 33 illus-

15
trates that turbulence levels remain rather high com- 660 CAD - Velocity vectors
pared to the experimental data. Still turbulence dis- (a) Exp. (b) Fine mesh (2nd ord.)
tribution is affected by what happened during the in-
take stroke. Now that compression stroke starts, tur-
bulence levels are expected to increase due to the
breakdown of the vortex. Hence, it is expected that
predicted turbulence levels inside the cylinder will re-
main higher than the computed ones.

600 CAD - Turbulence intensity


(a) Exp. (b) Fine mesh (2nd ord.)

Figure 35: Comparison between computed and experimen-


tal velocity vectors at 660 CAD: (a) Experimental data, (b)
Fine mesh with second order schemes.

660 CAD - Turbulence intensity


(a) Exp. (b) Fine mesh (2nd ord.)

Figure 33: Comparison between computed and experimen-


tal turbulence intensity distribution at 600 CAD: (a) Exper-
imental data, (b) Fine mesh with second order schemes.
Units are in [m/s], scale is 0 (blue) - 3 m/s (red).

660 CAD - Velocity magnitude


(a) Exp. (b) Fine mesh (2nd ord.)

Figure 36: Comparison between computed and experimen-


tal turbulence intensity distribution at 660 CAD: (a) Exper-
imental data, (b) Fine mesh with second order schemes.
Units are in [m/s], scale is 0 (blue) - 3 m/s (red).

Finally, when coming to TDC, still the agreement be-


tween experimental and computed data is acceptable,
mainly close to the spark-plug and above the piston.
Fig. 37(a) displays that the high velocity region is ex-
perimentally found just on left of the spark-plug in ex-
Figure 34: Comparison between computed and experimen-
periments, while in Fig. 37(a) it is possible to see that
tal velocity magnitude distribution at 660 CAD. Units are in
[m/s], scale is 0 (blue) - 7 m/s (red). predicted charge motions still involve also the right
part. However, it should be remarked that, at this time,
Halfway during compression, at 660 CAD, the flow is velocity magnitude is very small and for this reason
still characterized by a strong vortex as it can be seen velocity vector comparison for this condition is omit-
in Fig. 34. However, simulations present a stronger ted.
vortex compared to experimental data. This is much
clear from Fig. 35, where computed data show a Finally, levels of turbulence intensity at TDC are illus-
stronger horizontal component of the velocity vectors trated in Fig. 38. Predicted distribution is similar to ex-
compared to the measured ones. Fig. 36 shows that perimental data, even if turbulence levels are higher
both experiments and simulations predict an increase because of the highest intensity of the predicted vor-
of turbulence intensity due to dissipation of the vortex tex. However, the agreement is considered satisfac-
kinetic energy into turbulence. tory and the authors consider the proposed setup
(fine mesh and second order schemes) suitable also

16
for running combustion simulations on the same en- grid point positions. Two different windows were con-
gine. sidered: the high resolution one, 26 mm large, which
was extensively used to validate the result and the so-
720 CAD - Velocity magnitude called low resolution grid having approximately twice
(a) Exp. (b) Fine mesh (2nd ord.) the size of the first one and was used in this work in
Fig. 27. Results on the high resolution windows are
displayed in Fig. 39 (a) - (b).

400
Experimental
(a)
350

Specific kinetic energy [kJ/kg]


Coarse mesh
300
Fine mesh (1st ord.)
250
Fine mesh (2nd ord.)
200

150

Figure 37: Comparison between computed and experimen- 100


tal velocity magnitude distribution at 720 CAD. Units are in
[m/s], scale is 0 (blue) - 3 m/s (red). 50

0
720 CAD - Turbulence intensity
360 420 480 540 600 660 720
(a) Exp. (b) Fine mesh (2nd ord.)
Crank Angle [°]
140
Experimental (b)
RMS of kinetic energy [kJ/kg]

120
Coarse mesh
100
Fine mesh (1st ord.)

80 Fine mesh (2nd ord.)

60

40
Figure 38: Comparison between computed and experimen-
tal turbulence intensity distribution at 700 CAD: (a) Exper-
20
imental data, (b) Fine mesh with second order schemes.
Units are in [m/s], scale is 0 (blue) - 2 m/s (red).
0
360 420 480 540 600 660 720
To summarize the computed results and further un- Crank Angle [°]
derstand the predictive capability of the proposed ap-
proach, a comparison between computed and exper-
Figure 39: Comparison between computed and experimen-
imental kinetic energy and its rms was performed. tal evolution of specific kinetic energy (a) and rms of the
Such quantities are computed on the two experimen- specific kinetic energy (b) on the high-resolution measure-
tal grids were velocity field data were acquired, and ment window.
they are defined as:
ng
As expected, results from the fine mesh with second
X 1 order grid are the ones that better agree with exper-
Kmean = Ui2 (10)
2 imental data. The different way the vortex below the
i=1
ng valve is predicted is responsible for an underestima-
X1
rmsK = u2i (11) tion of the kinetic energy during the intake stroke and
i=1
2 an overestimation during the compression stroke due
to the higher vortex strength during that phase. Un-
where ng is the number of grid points in the measure- derprediction of the incoming jet penetration is mainly
ment window, U the velocity and u′ the turbulence in- responsible for the bad agreement with experimen-
tensity. The same procedure was carried out for com- tal data provided by both the coarse mesh and fine
puted data, which were interpolated at experimental mesh with first order schemes. For what concerns the

17
RMS of turbulent kinetic energy, satisfactory results Conclusions
were provided by the fine mesh with second order
schemes, even if during the compression stroke the This work presented a comprehensive methodol-
higher vortex strength produces more turbulence than ogy for full-cycle simulations in IC engines. To this
what was experimentally expected. end, multiple deforming grids were used and cre-
ated by a Cartesian, body-fitted mesh generator,
The comparison shown on the low-resolution window, snappHexMesh, available in the OpenFOAM code. The
shown in Figs. 40(a) - (b) provides a better idea of portion of cycle covered by each mesh is mainly
the way simulations predict the main details of the limited by prescribed quality parameters (skewness,
flow field. Results are very similar to Fig. 39, with the non-orthogonality), its topological or geometrical va-
fine mesh with second-order schemes to be the one lidity. Grid points are displaced by means of a finite-
providing the best agreement with experimental data. volume based automatic mesh motion technique. The
Remarkable, to the authors’ opinion, the capability to possibility to employ an automatic mesh generator
reproduce the turbulence decay during compression. significantly reduces the required amount of user pre-
processing time, since only a cleaned template geom-
etry has to be provided at a prescribed initial position.
The grid generator tool supports boundary layer cre-
250
ation and local refinement, allowing a better predic-
Experimental
(a) tion of incoming cylinder flow and other relevant physi-
Specific kinetic energy [kJ/kg]

200 Coarse mesh


cal phenomena. Finally, the Cartesian mesh structure
strongly reduces the mesh non-orthogonality, making
Fine mesh (1st ord.) possible to employ high-order schemes.
150 Fine mesh (2nd ord.)
The proposed approach was validated by performing
RANS simulations at non-combusting conditions for
100
an optical engine whose data were available in the
ECN database. In particular, effects of mesh structure
50 and numerical schemes were evaluated by testing dif-
ferent configurations. Accurate, full-cycle simulations
are possible, however mesh structure and in particu-
0 lar size close to the valves plays a big role in com-
360 420 480 540 600 660 720
puted results, even more, at least in the authors’ opin-
Crank Angle [°] ion, than the approach used for turbulence. To this
100 end, the possibility to incorporate adaptive local mesh
90 Experimental (b) refinement in the proposed methodology appears to
RMS of kinetic energy [kJ/kg]

be of great importance to improve the quality of re-


80 Coarse mesh sults.
70 Fine mesh (1st ord.)
60 Fine mesh (2nd ord.) Acknowledgements
50
40 Authors would like to thank Prof. Volker Sick and Prof.
David Reuss for providing the detailed experimental
30
data-set of the TCC engine.
20
10
Contact Information
0
360 420 480 540 600 660 720
Dr. Tommaso Lucchini
Crank Angle [°] Department of Energy, Politecnico di Milano
Via Lambruschini, 4
Figure 40: Comparison between computed and experimen- 20156 Milano, Italy
tal evolution of specific kinetic energy (a) and rms of the tommaso.lucchini@polimi.it
specific kinetic energy (b) on the low-resolution measure-
ment window.

18
REFERENCES 10. R. Dahms, T.D. Fansler, M.C. Drake, T.-W. Kuo,
A.M. Lippert, and N. Peters. Modeling ignition
1. D. L. Reuss, R. J. Adrian, C. C. Landreth, D. T. phenomena in spray-guided spark-ignited en-
French, and T. D. Fansler. Instantaneous Planar gines. Proceedings of the Combustion Institute,
Measurements of Velocity and Large-Scale Vor- 32(2):2743 – 2750, 2009.
ticity and Strain Rate in an Engine Using Particle-
Image Velocimetry. SAE Paper, 890616, 1989. 11. A. D. Gosman. State of the Art of Multi-
Dimensional Modeling of Engine Reacting Flows.
2. D. L. Reuss, T. Kuo, B. Khalighi, D. Haworth, and Oil and Gas Science and Technology, 54(no. 2),
M. Rosalik. Particle Image Velocimetry Measure- 1999.
ments in a High-Swirl Engine Used for Evaluation
of Computational Fluid Dynamics Calculations. 12. H. Barths, C. Hasse, and N. Peters. Computa-
SAE Paper, 952381, 1995. tional fluid dynamics modelling of non-premixed
combustion in direct injection diesel engines. In-
3. S. Gaikwad, K. Arora, V. Korivi, and S. Cho. ternational Journal of Engine Research, 1 (3):pp.
Steady and Transient CFD Approach for Port Op- 249–267, 2000.
timization. SAE Int. J. Mater. Manuf., 1:754–762,
2009. 13. J. P. Duclos, M. Zolver, and T. Baritaud. 3d Model-
ing of Combustion for DI-SI Engines. Oil and Gas
4. J. W. Campbell, G. Hardy, and F. Personeni. Science and Technology, 54:pp. 259–264, 1999.
Diesel Engine Intake Port-Flow Stability and ap-
14. K. Y. Huh and A. D. Gosman. A Phenomeno-
plication of CFD using a Hybrid-Low-Reynolds
logical Model of Diesel Spray Atomization. Pro-
new wall method. THIESEL 2006 Conference on
ceedings of the International Conference on Mul-
Thermo- and Fluid Dynamic Processes in Diesel
tiphase Flows, Tsukuba, Japan, 1991.
Engines, 2006.
15. C. Bai and A. D. Gosman. Mathematical Modeling
5. O. Colin, A. Benkenida, and C. Angelberger.
of Wall Films Formed by Spray Wall Interaction.
3d Modeling of Mixing, Ignition and Combustion
SAE Paper, 960626, 1996.
Phenomena in Highly Stratified Gasoline Engines
. Oil and Gas Science and Technology, 58:pp. 16. T. Lucchini, G. D’Errico, and D. Ettorre. Numerical
47–62, 2003. investigation of the spray-mesh-turbulence inter-
actions for high-pressure, evaporating sprays at
6. S. Singh, R. D. Reitz, and M. P. B. Musculus. engine conditions. International Journal of Heat
Comparison of the Characteristic Time (CTC), and Fluid Flow, 32:pp. 285–297, 2011.
Representative Interactive Flamelet (RIF), and
Direct Integration with Detailed Chemistry Com- 17. R. N. Dahms, M. C. Drake, T. D. Fansler, T.-
bustion Models against Optical Diagnostic Data W. Kuo, and N. Peters. Understanding ignition
for Multi-Mode Combustion in a Heavy-Duty DI processes in spray-guided gasoline engines us-
Diesel Engine. SAE Paper, 2006-01-0055, 2006. ing high-speed imaging and the extended spark-
ignition model SparkCIMM. Part A: Spark chan-
7. G. D’Errico, T. Lucchini, R. Di Gioia, and G. Bo- nel processes and the turbulent flame front prop-
nandrini. Application of the CTC Model to Pre- agation. Combustion and Flame., 158 (11):2229–
dict Combustion and Pollutant Emissions in a 2244, 2011.
Common-Rail Diesel Engine Operating with Mul-
tiple Injections and High EGR. SAE Paper, 2012- 18. H. Jasak, J. Y Luo, B. Kaludercic, A. D. Gos-
01-0154, 2012. man, H. Echtle, Z. Liang, F. Wirbeleit, M. Wierse,
S. Rips, A. Werner, G. Fernstrm, and A. Karls-
8. T. Lucchini, G. D’Errico, A. Onorati, G. Bonan- son. Rapid CFD Simulation of Internal Combus-
drini, L. Venturoli, and R. Di Gioia. Development tion Engines. SAE Paper, 1999-01-1185, 1999.
of a CFD Approach to Model Fuel-Air Mixing in
Gasoline Direct-Injection Engines. SAE Paper, 19. H. Si, J. Fuhrmann, and K. Gartner. Boundary
2012-01-0146, 2012. Conforming Delaunay Mesh Generation. Com-
put. Math. Phys., Vol. 50:38–53, 2010.
9. G. Bonandrini, R. Di Gioia, D. Papaleo, and
L. Venturoli. Numerical Study on Multiple In- 20. J. Schoberl. NETGEN An advancing front 2D/3D-
jection Strategies in DISI Engines for Particulate mesh generator based on abstract rules. Com-
Emission Control. SAE Paper, 2012-01-0400, puting and Visualization in Science, Vol. 6:41–52,
2012. 1997.

19
21. N. Sinha, P. Cavallo, R. Lee, A. Hosangadi, D. C. 33. P. M. Knupp. Algebraic mesh quality metrics for
Kenzakowski, S.M. Dash, H. Affes, and D. Chu. unstructured initial meshes. Finite Elements in
Novel CFD Techniques For In-Cylinder Flows On Analysis and Design, Vol. 39:217–241, 2003.
Tetrahedral Grids. SAE Paper, 980138, 1998.
34. J. H. Ferziger and M. Peric. Computational Meth-
22. D. Schmidt, S. Toninel, S. Filippone, and ods for Fluid Dynamics. Springer, 2002.
G. Bianchi. Parallel Computation of Mesh Mo-
tion for CFD of IC Engines. SAE Paper, 2008-01- 35. A.Y. Alharbi and V. Sick. Investigation of bound-
0976, 2008. ary layers in internal combustion engines using
a hybrid algorithm of high speed micro-PIV and
23. P. K. Senecal, K. J. Richards, E. Pomraning, PTV. Experiment in Fluids, 49 (4):pp. 949–959,
T. Yang, M. Z. Dai, R. M. McDavid, M. A. Pat- 2010.
terson, S. Hou, and T. Shethaji. A New Paral-
lel Cut-Cell Cartesian CFD Code for Rapid Grid 36. D. C. Haworth. Large-Eddy Simulation of in-
Generation Applied to In-Cylinder Diesel Engine Cylinder Flows. Oil and Gas Science and Tech-
Simulations. SAE Paper, 2007-01-0159, 2007. nology, 54 (2):pp. 175–185, 1999.
24. X. Yang, A. Solomon, and T. Kuo. Ignition and 37. K. Liu and D. C. Haworth. Large-eddy simula-
Combustion Simulations of Spray-Guided SIDI tion for an axisymmetric piston-cylinder assembly
Engine using Arrhenius Combustion with Spark- with and without swirl. Flow, Turbulence & Com-
Energy Deposition Model. SAE Paper, 2012-01- bustion, 85:pp. 279–307, 2010.
0147, 2012.
38. P. Abraham, K. Liu, D. Haworth, D. Reuss, , and
25. Z. Wang, R. Scarcelli, S. Som, and S. McConnell V. Sick. Evaluating Large-Eddy Simulation (LES)
et al. Multi-Dimensional Modeling and Valida- and High-Speed Particle Image Velocimetry (PIV)
tion of Combustion in a High-Efficiency Dual-Fuel with Phase-Invariant Proper Orthogonal Decom-
Light-Duty Engine. SAE Paper, 2013-01-1091, position (POD). Oil and Gas Science and Tech-
2013. nology, proceedings of LES4ICE 2012 - Large
Eddy Simulation for Internal Combustion Engine
26. R. Scarcelli, N. Matthias, and T. Wallner. Numer-
Flows, 2013.
ical Investigation of Combustion in a Lean Burn
Gasoline Engine. SAE Paper, 2013-24-0009,
2013.
27. T. Lucchini, G. D’Errico, H. Jasak, and Z. Tukovic.
Automatic Mesh Motion with Topological
Changes for Engine Simulation. SAE Paper,
2007-01-0170, 2007.
28. T. Lucchini, G. D’Errico, and M. Fiocco. Multi-
dimensional modeling of gas exchange and fuel-
air mixing processes in a direct-injection, gas fu-
eled engine. SAE Paper, 2011-24-0036, 2011.
29. N. Nordin. Complex Chemistry Modeling of
Diesel Spray Combustion. PhD thesis, Chalmers
University of Technology, Department of Thermo
Fluid Dynamics, 2001.
30. T. Lucchini, G. D’Errico, F. Brusiani, and
G. Bianchi. A Finite-Element Based Mesh Motion
Technique for Internal Combustion Engine Simu-
lations. COMODIA 2008, MS2-3, 2008.
31. H. Jasak. Error Analysis and estimation for the
finite volume method with applications to fluid
flows. Ph.d thesis, Imperial College of Science,
Tecnology and Medicine, London, 1996.
32. H. Jasak and Z. Tuković. Automatic mesh motion
for the unstructured finite volume method. Trans-
actions of FAMENA, 30(2):1–18, 2007.

20

View publication stats

You might also like