Sunteți pe pagina 1din 11

International Journal of Heat and Mass Transfer 120 (2018) 1189–1199

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Numerical investigation of boundary layer flow and wall heat transfer in


a gasoline direct-injection engine
Xueqi Fan, Zhizhao Che, Tianyou Wang ⇑, Zhen Lu
State Key Laboratory of Engines, Tianjin University, Tianjin 300072, China

a r t i c l e i n f o a b s t r a c t

Article history: Near-wall turbulent boundary-layer has a substantial impact on the wall heat transfer process in internal
Received 19 May 2017 combustion engines, and ultimately affects the engine performance. However, the heat transfer processes
Received in revised form 16 August 2017 are not well understood because of a lack of information on gas side velocity and temperature distribu-
Accepted 23 September 2017
tion. In the present study, the transient velocity and thermal boundary layers in a motored spark-ignition
Available online 4 January 2018
direct-injection engine are predicted using a Delay Detached Eddy Simulation Shear-stress Transport
(DDES-SST) model. The near-wall ensemble-averaged and fluctuation velocity and temperature fields
Keywords:
at the cylinder head are investigated throughout the compression stroke. The numerical results of the
Turbulent boundary layer
Heat transfer
flow field in engine core region agree well with high-speed Two Dimensional Three Components-
DDES Particle Image Velocimetry (2D3C-PIV) experimental data. The simulated data show that boundary-
GDI engine layer separation occurs around the closing phase of intake valves and the turbulence in the near-wall
region is anisotropic. The thickness of thermal boundary layer shows a non-monotonic variation with
time. The dimensionless velocity profiles agree well with the law of the wall in the viscous sublayer
but deviate from the log layer. The dimensionless temperature profiles tend to be uniform in the log-
layer. The maximum local wall heat flux and heat transfer coefficient are 70 kW/m2 and 200 W/(m2 K)
at top dead center (TDC), respectively. The spatial distribution of turbulent heat flux shows a strong link
between the velocity and thermal boundary layers, and the peaks in the buffer layer occupy about 70–
85% of the local wall heat flux, indicating that the forced convection heat transfer is the main heat trans-
fer mechanism in internal combustion engines.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction details of ICE boundary layer flows and wall heat transfer is crucial
for the design and optimization of ICEs.
In-cylinder turbulent flow has a substantial impact on air-fuel Despite the importance of understanding wall heat transfer in
mixture formation and flame propagation in internal combustion ICEs, only a few studies have been focused on the near-wall turbu-
engines (ICEs), and the turbulent boundary layer contributes a lot lence flow compared with the flow structures in engine core
to the wall heat transfer which ultimately affects the engine effi- region. Laser Doppler Velocimetry (LDV) was firstly applied to
ciency, pollutant formation and component thermal stresses [1– investigate the near-wall velocity filed in ICEs in 1980s [7]. Foster
3]. Therefore, quantitative prediction of local and transient heat & Witze [8] measured as close as 60 mm from the cylinder head and
transfer is important not only to accurate simulations but also to found that the boundary layer thickness was less than 1 mm.
the optimization of ICEs [4]. Due to experimental limitations in Pierce et al. [9] conducted LDV with Particle Image Velocimetry
measuring the velocity and temperature distribution in the near- (PIV) to measure the 3D velocity field on the cylinder head and
wall region, the heat transfer processes are still not well under- proved that the boundary layer in ICEs was at least a 2D velocity
stood [5]. Besides, the boundary layer flow in ICEs is influenced field. Recently, high-speed micro-PIV and Particle Tracking
by the non-equilibrium turbulence due to high-speed reciprocating Velocimetry (PTV) were applied in the near-wall region in ICEs,
motion of the piston and the periodic opening and closing of the and greatly improved the measurement accuracy. Alharbi & Sick
valves, so the current wall function models are not suitable for [10–13] measured as close as 45 mm from the cylinder head on a
the complex conditions in engines [6]. Hence, understanding the motored engine and observed reversal flows and millimeter vortex
structures. Besides, they found that the dimensionless velocity pro-
⇑ Corresponding author. files agreed well with the law of the wall in the viscous sublayer
E-mail address: wangtianyou@tju.edu.cn (T. Wang).
but deviated from the log layer. Jainski et al. [14] extended the

https://doi.org/10.1016/j.ijheatmasstransfer.2017.09.089
0017-9310/Ó 2017 Elsevier Ltd. All rights reserved.
1190 X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199

above measurement to three different engine speeds and found of simple geometry like channels, pipes and plates [36]. Shadloo
that the thickness of the viscous sub-layer decreased as the engine et al. [37,38] investigated the effect of wall heat transfer on turbu-
speed increased. Koehler et al. [15] conducted time-resolved PIV lent statistics and near wall behaviors in supersonic turbulent
(TR-PIV) to measure as close as 1.5 mm from the piston crown, boundary layers under different wall conditions using DNS. The
but the experiment could only be carried out during the intake results showed that increasing the disturbance amplitude as well
stroke because of the working fluid was water rather than air. as perturbation frequency moved the laminar-to-turbulent transi-
The thermal boundary layers in ICEs can be accessed by Sch- tion upstream. Wu & Moin [39] simulated an incompressible, zero-
lieren photography, Coherent Anti-Stokes Raman Scattering (CARS) pressure-gradient flat-plate boundary layer and observed that
and Laser-Induced Fluorescence (LIF) techniques. Lucht et al. hairpin vortices developed into the downstream hairpin forests
[16,17] measured the thermal boundary layer on the cylinder head in the transitional region (800 < Re < 1900). However, DNS method
using CARS, and found that the thermal boundary layer thickness is seldom used in ICEs because of the high CPU demand. Schmitt
in fired engine increased during the expansion stroke. Cundy and et al. [40–42] conducted DNS to simulate the boundary layer flow
Sick [12,18,19] conducted LIF to measure the temperature distribu- with wall heat transfer in an engine-relevant condition. The results
tion in the combustion chamber, and the results showed that the showed that the velocity and thermal boundary layer thicknesses
thermal boundary layer thickness was thinner than the results decreased towards top dead center, and the dimensionless velocity
obtained by Lyford-Pike [20] using Schlieren photograph. They also and temperature profiles deviated strongly from the law of the
observed thermal stratification on a cooled curved metal plate wall. The DNS data showed that the turbulent heat flux occupied
from the transient temperature fields. Tran et al. [21] observed almost 60% to 80% of total wall heat flux. These conclusions were
the non-uniform temperature distribution in a Homogeneous significant but the simulation geometry was simplified without
Charge Compression Ignition (HCCI) engine by using LIF. Further- combustion chamber structure and moving valves, so it could not
more, the instant wall heat flux in ICEs was generally measured reflect the real boundary layer flow in ICEs.
by thermocouples. Nijeweme et al. [22] found that the wall heat In summary, the experiments could only access limited areas on
flux was influenced by the test location, engine speed, throttle set- wall surfaces in ICEs and there is a lack of simulation of velocity
ting and ignition time. Gingrich et al. [23] compared the instant and thermal boundary layers in realistic engines to the authors’
heat flux on the piston crown under different combustion regimes. knowledge. The objective of the present study is to analyze the
Ma et al. [24] measured the instantaneous heat flux on the cylinder velocity boundary layer and thermal boundary layer together in a
head, the peak heat flux was about 60 kW/m2 under motored con- realistic engine, and provide a further understanding of wall heat
dition at the engine speed of 500 rpm. flux and turbulent heat transfer in ICEs. The simulation setup and
Since those experiments could only provide velocity and tem- experimental validation are presented in Section 2. In Section 3,
perature distribution in limited areas, multi-dimensional simula- the evolution of velocity and thermal boundary layers during the
tions are necessary to solve the transient flow and heat transfer compression stroke are described, the dimensionless velocity and
in ICEs. However, the wall models often used in Reynolds- temperature profiles are compared against the law of the wall,
Averaged Navier-Stokes (RANS) and Large-Eddy Simulation (LES) and local wall heat flux and heat transfer coefficient are predicted.
in ICEs are too simple and need to be refined [25]. The most widely Conclusions are drawn in Section 4.
used wall model is the classic equilibrium wall-function model
derived by Prandtl [26]. Han and Reitz [4] improved this model
by considering the density and turbulent Prandtl number variation 2. Numerical method
in the near-wall region, and the improved model has also been
widely used in multi-dimensional simulations in ICEs. After that, 2.1. Governing equations and turbulence modeling
many researchers further improved the Han-Reitz heat transfer
model in ICEs by considering the pressure work term [27], the vari- Since the characteristic velocity in ICEs is on the order of 10 m/s,
ation of gas viscosity and turbulent Prandtl number [28,29] and the the in-cylinder flow can be described by the low Mach number
subgrid-scale (SGS) turbulent viscosity [30]. Ma et al. [31] con- compressible Navier-Stokes equations [31]. The compressibility
ducted k-omega turbulence model with pressure gradient to solve of the fluid is ruled by the ideal gas law. The governing equations
the 2D unsteady velocity boundary layer under the experimental for mass, momentum and energy are as follows:
condition conducted by Jainski and Sick [11,14], the gas tempera- @q
ture was about 640 K at 330 CAD under motored condition at the þ r  ðquÞ ¼ 0 ð1Þ
@t
engine speed of 800 rpm.
 
In addition to the wall-function models, hybrid LES/RANS @ui @P
q þ rðui  uÞ ¼  þ rðl  gradui Þ þ Si ð2Þ
method and Detached Eddy Simulation (DES) have been increasing @t @xi
popular to simulate in-cylinder flows. Jhavar et al. [32] solved the
 
in-cylinder flow based on Very Large Eddy Simulation (VLES) and @T @P0
qcp þ u  rT ¼ þ rðk  rTÞ ð3Þ
RANS k-e model. Hasse et al. [33] conducted Shear-Stress Trans- @t @t
port DES (SST DES) to simulate in-cylinder flow and the results
agreed better with experimental data than the RANS approach. P0 ¼ qRT ð4Þ
Hartmann et al. [34] investigated the velocity boundary layer at
where P is the hydrodynamic pressure, P0 is the thermodynamic
the intake valve seat based on SST DES model and observed flow
pressure, and the physical pressure is the sum of this two terms.
separation in the near-wall region. Buhl et al. [35] investigated
cp is the specific heat at constant pressure, l is the dynamic viscos-
the velocity boundary layer on the piston crown during the intake
ity and k is the thermal conductivity. Among them [43],
stroke using Scale Adaptive Simulation Shear-stress Transport
(SAS-SST) model, and the results showed that the boundary layer T 1:5
thickness varied strongly along the piston surface due to the l ¼ 1:46  106 ð5Þ
T þ 111
large-scale tumble flow.
Direct numerical simulation (DNS) has been widely used to T 1:5
investigate the dynamics of the coherent structures and the phy- k ¼ 2:089  103 ð6Þ
T þ 111
sics of turbulent heat transfer within the turbulent boundary layers
X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199 1191

 2  3
T T T
cp ¼ 1:05  0:365 þ 0:85  0:39 ð7Þ
1000 1000 1000
In the momentum equation, ui are the velocity components, xi
are the coordinates, Si are the generalized sources.
     
@ @ui @ @u @ @u @
Si ¼ l þ l j þ l k þ ðkruÞ ð8Þ
@xi @xi @xj @xi @xk @xi @xi

where k is the second viscosity coefficient.


The turbulence model used in this study is DDES-SST which is a
seamless hybrid URANS/LES method. A major advantage of this
model is that it only requires grid refinement in the wall-normal
direction in the near-wall region to solve wall effect. Therefore, it
has higher precision than wall function methods and less CPU
demand than LES approach. The SST RANS model is a combination
of standard k-e and k-x model. It equals the k-x model within the
boundary layer to solve attachment and the k-e model in the outer Fig. 1. Domain of the simulation.
region to solve separation. The turbulent length scale of SST RANS
model is lt ¼ k =ðb xÞ. LES approach would be applied if the grid
1=2

resolution is sufficiently fine to resolve smaller turbulent struc- Table 1


Engine parameters.
tures than lt , so DES length scale should be defined as:
Parameters values
lDES ¼ minðlt ; C DES DÞ ð9Þ
Bore 82.5 mm
where D ¼ maxðDi Þ is the maximum length scale of the computa- Stroke 90 mm
Connecting rod 168 mm
tional cell and C DES is a modeling constant. This idea successfully
Compression ratio 9.7:1
shifts the standard RANS SST model to a DES formulation. Addition- TDC clearance height 1.4 mm
ally, this work utilizes a Delayed-DES method to keep LES region out Intake valve opens 22 CAD
of the boundary layer and ease the modeled-stress depletion prob- Intake valve closes 241.9 CAD
lem. The dissipation term could be written as: Exhaust valve opens 491.3 CAD
Exhaust valve closes 706.6 CAD
 b kx  F DES
DkDES ¼ q ð10Þ Engine speed 800 rpm

 
lt utilized structured hexahedron grids. The grid reference size was
F DES ¼ max ð1  F SST Þ; 1 ð11Þ
C DES D selected as 1.3 mm to balance the computational accuracy and effi-
ciency. The elements on the cylinder head were refined with 10
where F SST equals the blending function in SST RANS model.
prism layers in the near-wall region. The first grid point spaced
However, the DDES-SST turbulent model requires high quality
0.1 mm from the wall, and the growth ratio was 1.1. Fig. 2(b)
grids and the first grid point should be located within the viscous
shows the dimensionless height of the first grid point at x = 0,
sublayer (y+ < 5) in RANS. The LES region in DES also requires a suf-
y = 8 mm versus crank angle degree (CAD). The red dashed line
ficiently fine isotropic grid resolution to support subgrid-scale
represents the mesh requirement of DES approach (y+ < 5), so the
model to resolve the isotropic small scale turbulent flow. It is chal-
current computational mesh satisfied the requirement in most
lenge to meet all these requirements in the complex geometries
regions. Although y+ of three piston positions close to TDC exceed
especially with moving boundaries [33]. Spalart [44] recom-
the dashed line, the current computing resource cannot support
mended to locate the first grid point at y+ = 5 and to use a ratio
further mesh refinement and the calculation accuracy is accept-
of yj+1/yj = 1.3 for the first attempt.
able. The Reynolds number (Re ¼ us  B=m) of the in-cylinder
boundary layer flow is proportional to y+ and it is raised from
2.2. Simulation setup 1500 to 5000 during the compression stroke.
A measure Mðx; tÞ [45] of the turbulence resolution is proposed
The flow field and heat transfer in a motored Direct-Injection to quantitatively validate the mesh resolution in LES region, and it
Spark-Ignition (DISI) optical engine was simulated using commer- is defined as follows:
cial CFD software FLUENT. We focused on the transient velocity
and thermal boundary layers on the cylinder head in the tumble K res ðx; tÞ
plane (x = 0 in Fig. 1) during the compression stroke. The optical Mðx; tÞ ¼ ð12Þ
K res ðx; tÞ þ K sgs ðx; tÞ
engine was simplified from Volkswagen MAGOTAN and the fuel
injector and the spark-plug were removed from the pent-roof com- K res ðx; tÞ represents the turbulent kinetic
energy calculated from
bustion chamber. The engine specifications and valve timings are 1
2
v
ðu2rms þ 2rms þ w2rms Þ, K sgs ðx; tÞ represents
the turbulent kinetic
summarized in Table 1. We only simulated a half of the symmetric energy calculated from LES subgrid-scale model. So Mðx; tÞ shows
geometry for the sake of computing resources. In this paper, 0 the ratio of resolved turbulent kinetic energy to the amount of total
crank angle degree (CAD) refers to top dead center at the beginning turbulent kinetic energy. In general, a LES calculation is considered
of the intake stroke. to be good, if the resolved structures contain at least 80% of turbu-
According to the requirements of DDES-SST model, the compu- lent kinetic energy (Mðx; tÞ > 0:8) [46]. Fig. 3 shows the Mðx; tÞ
tational mesh was constructed in The Integrated Computer Engi- phase-averaged distribution on mid-section plane at BDC. Mðx; tÞ
neering and Manufacturing code for Computational Fluid is higher than 0.8 in most areas, so the mesh resolution is fine
Dynamics (ICEM CFD) with the hybrid mesh strategy as shown in enough in LES region. Overall, the whole computational domain
Fig. 2(a). The ports and the upper part of the combustion chamber contains approximately 1.4 million cells at the bottom dead center
utilized unstructured tetrahedral grids and the cylinder area (BDC), and about 0.65 million cells at TDC.
1192 X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199

Fig. 2. Computational mesh setting (a) mesh at BDC; (b) first grid y+ at x = 0, y = 8 mm.

Fig. 3. M(x, t) phase-averaged on mid-section plane at 180CAD.


Fig. 4. Schematic of the optical engine set-up and the PIV system for tumble
measurements.
The initial temperature within the cylinder is 318 K. The initial
temperature of the intake port and the exhaust port are 313 K and
of 333 K, respectively. The wall temperature on the cylinder head is
field on the tumble plane was measured during the intake and com-
fixed at 348 K and all the other region is fixed at 318 K. The initial
pression strokes (30–330 CAD). A dual oscillator single head
pressure of the inlet and the outlet are 325 Pa (gauge pressure)
527 nm laser (LDY 300) was operated at a frequency of 2.4 kHz
and of 1325 Pa, respectively. The values of temperature and pres-
and the laser sheet thickness is 1 mm. 1 lm oil tracker particles
sure in boundary condition are the same as the initial condition.
were captured by two CCD cameras (HighSpeedStar 6) with the full
The PISO algorithm was used to solve the velocity-pressure cou-
1024  1024 pixels chip. The interrogation window in post-
pling. The time scheme is second order implicit and the time step
processing is 64  64 pixels. The final view size is 64.8  59.4 mm
varies between 13 and 5.2 ls. For momentum equation, the
and the spatial resolution of the velocity vector is 1.03 mm. The cal-
bounded center differencing is used. The discretization scheme
ibration error of CCD camera is 0.029 pixel, the maximum theoret-
applied for energy and turbulent kinetic energy is second order
ical error of three velocity components is less than 0.72 m/s at BDC.
upwind. The interpolation scheme for pressure is PRESTO!, the
The simulation results and experimental data are compared
interpolation scheme for density is second order upwind. The case
base on the macro scale flow field rather than the near-wall region,
begins at 380 CAD and runs through a total of 32 continuous
because it is challenging in the experiment to capture the unsteady
engine cycles. The first 2 cycles are discarded to ignore the impact
velocity boundary layer in a complex cylinder head structure. The
of initial conditions. The simulation was performed on a 20-core
ensemble-averaged 3D velocity vectors of PIV and DDES at 240
workstation and required approximately 96 h for each cycle.
CAD and 270 CAD are shown in Fig. 5 and the white point in the
figures represents the position of the tumble center. The red point
2.3. Comparison with experimental results in Fig. 5a represents the origin of the coordinate system. The exper-
imental data were averaged from 100 cycles to reduce the experi-
A set of two-dimensional three-component particle image mental uncertainty due to cycle-to-cycle variation and the
velocity (2D3C-PIV) experiments were made on an optical DISI simulated data were averaged from 30 cycles.
engine to validate the simulation results. The experiments were The relevance index (RI) proposed in Ref. [47] is employed to
conducted with Lavision high speed 3D-PIV system, the schematic quantitatively analyze the resemblance between PIV and DDES
of the experiment is shown in Fig. 4. The three-dimensional velocity velocity fields, and it is defined as follows:
X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199 1193

Fig. 5. Validation of the DDES velocity fields against PIV results. (a) 240 CAD DDES; (b) 240 CAD PIV, RI = 0.91 at 240 CAD; (c) 270 CAD DDES; (d) 270 CAD PIV, RI = 0.94 at 270
CAD.

ðu; v Þ the upward piston. After 320 CAD, the velocity gradually decreases
RIu;v ¼ ð13Þ
kuk  kv k as the piston slows down. At the same time, the gas momentum
dissipates because of the large-scale tumble flow breaks down to
RIu;v ranges from 1 to 1.RIu;v = 1 means that the two velocity
small-scale flow structures at the end of the compression stroke.
fields are identical; RIu;v = -1 indicates the two velocity fields have
The velocity magnitude is about 2 m/s at the TDC.
the same magnitude but the opposite flow direction; and RIu;v = 0
Affected by the large-scale tumble flow, the flow direction near
means the two velocity field are completely unrelated.
the cylinder head surface should be negative as shown in Fig. 5(a).
The DDES and PIV results show good agreement in the aspects
However, a positive velocity appears at 220 CAD as shown in Fig. 6
of the position of the tumble center, the velocity magnitude distri-
(a). Analyzing the reason, there would form a small gap between
bution and the flow direction. The RI is 0.91 at 240 CAD and 0.94 at
the intake valve and the valve seat before the complete close of
270 CAD.
the intake valve, and this small gap will lead to a high-speed
low-pressure region which results in a local reversal flow. Besides,
boundary layer separation occurs within 0.5 mm spacing from the
3. Results and discussion wall at 210 CAD and 230 CAD, the near-wall transient velocity field
at 210 CAD in the 24th cycle is shown in the inset of Fig. 6(a). The
3.1. Velocity boundary layer top boundary refers to the cylinder head surface and the white dot
on the wall indicates the separation point of the boundary layer.
Wall-bounded flow is influenced by the fluid viscosity in the Therefore, the stability of the boundary layer is significantly influ-
near-wall region, which is defined as a 3 mm-thick layer from enced by the moving valves.
the wall in this paper. The typical point is selected at x = 0 mm, In addition to the flow separation, sub-millimeter vortex struc-
y = 8 mm on the cylinder head surface as shown in Fig. 5a. The tures randomly occur in the near-wall region. Fig. 7 shows the
ensemble-averaged velocity fields and the velocity fluctuation movement of a near-wall vortex from 195 to 205 CAD in the
fields are analyzed below. 28th cycle. The top boundary refers to the cylinder head surface.
The counterclockwise vortex generated in the boundary layer
3.1.1. Temporal evolution of the velocity boundary layer around 195 CAD and moved towards the wall. It finally disap-
Fig. 6 shows the evolution of the ensemble-averaged velocity peared after hitting the wall at 205 CAD so the lifetime lasted for
profiles of stream direction during the compression stroke. The more than 10 CAD. Similar vortex structures were observed in
velocity decreases from 180 to 240 CAD because of the intake valve the micro-PIV experiment in ICEs [11,14] and in plate boundary
closing. After the intake valve fully closed at 241.9 CAD, the veloc- layer flow [48,49]. These vortex structures can contribute substan-
ity rapidly increases to around 3 m/s and the momentum is due to tially to heat and mass transfer in the boundary layer.
1194 X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199

Fig. 6. Ensemble-averaged velocity profiles of stream direction at selected CADs. (a) Velocity profiles from 180CAD to 240 CAD; (b) Velocity profiles from 260 CAD to 360 CAD.

Fig. 7. Near-wall vortex movement from 195 to 205 CAD in the 28th cycle. (a) 205 CAD; (b) 210 CAD; (c) 215 CAD.

3.1.2. Anisotropic turbulence in the near wall region fluctuation intensities are on the same order as the ensemble-
The flow in the near-wall region of ICEs is a complex wall tur- averaged velocities and almost twice as that in the normal
bulent flow and the turbulence intensity is quantified by the fluc- direction.
tuation of the ensemble-averaged velocity. Fig. 8 shows the Reynolds stresses represent the mean momentum fluxes by tur-
evolution of the stream and normal fluctuation intensities during bulence and generally are larger than viscous stress. Fig. 9 shows
the compression stroke. Generally, the stream and normal fluctua- the Reynolds stress in the near-wall region at 180 and 300 CAD.
tion intensities gradually increase as the piston moves upward, but The two normal stresses gradually increase with the increase of
there is another peak at 240 CAD. The reason is that the closing of the distance from the wall while the shear stress has no significant
the intake valve leads to local reversal flow and momentum dissi- change. More specifically, Reyy is much larger than Rezz and Reyz
pation. In the normal direction, the velocity fluctuation increases throughout the compression stroke. Reyz is always close to zero
almost linearly as a function of the distance from the wall and and could be negative or positive. The results reflect that the turbu-
the maximum is about 1 m/s at TDC. In the stream direction, the lence in the near-wall region is anisotropic, and this is a major dif-
tendency of the velocity fluctuation profiles is similar to that of ference from the isotropic turbulence in the engine core region.
the mean velocity profiles, which grows rapidly within 0.5 mm Hence the mesh strategies should be different in the near-wall
spacing from the wall and grows gently after that. The stream region and engine core region, also known as zonal approach [50].

Fig. 8. Velocity fluctuation profiles at selected CADs. (a) Fluctuation intensity of the stream direction; (b) Fluctuation intensity of wall-normal direction.
X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199 1195

Fig. 9. Reynolds stresses distribution in near-wall region: the normal stress in the stream direction (y): Reyy ¼ v 02 , the normal stress in the normal direction (z): Rezz ¼ w02 , the
shear stress on the yoz plane: Reyz ¼ v 0 w0 . (a) Reynolds stresses at 180 CAD; (b) Reynolds stresses at 300 CAD.

3.2.1. Dimensionless velocity profiles deviate in the log layer (y+ < 30). The buffer layer (5 < y+ < 30) was
The dimensionless ensemble-averaged velocity profiles are not taken into account because of irregularity. The thickness of the
compared with the commonly used velocity law of the wall, which viscous sublayer is defined as y+ = 5 according to the law of the
was derived under the condition of two-dimensional steady and wall, but the thickness of the viscous sublayer length increases
fully developed flow over a plate. The law of the wall is usually as the CAD increases. The viscous sublayer thickness is about
non-dimensionalized using the viscous scale dv and the friction y+ = 5 at 220 CAD, but it exceeds y+ = 30 close to the TDC (360
velocity us : CAD). In the log layer, the dimensionless velocity increases as the
mw piston moves upward. So there is a negative offset between the
dv ¼ ð14Þ simulated curves and the standard wall function at the beginning
us
of the compression stroke, and the offset changes to positive later.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 In addition, the trend of the u+ profile is similar to the log law, but
lw @ u
us ¼ ð15Þ there is a decline tendency on the outside at several crank angle
q @y w w degrees due to the complex in-cylinder flow.
The standard wall function was derived under ideal condition
where qw and lw are the gas density and dynamic viscosity at the
such as wall-parallel flow, quasi-steady flow, zero-pressure gradi-
wall, mw ¼ lw =qw : Then, the distance from the wall and velocity
ent. The in-cylinder flow is three-dimensional unsteady with a
can be expressed according to the parameters above yþ ¼ y=dv ,
complex combustion chamber structure and a moving piston and
uþ ¼ u=us . The law of the wall consists of two layers: the viscous
valves, and these differences results in the inconsistency between
sublayer and the logarithm layer:
( the in-cylinder boundary layer velocity profiles with the law of
þ
yþ if yþ < 11 the wall. Moreover, the in-cylinder velocity boundary layer is
u ¼ ð16Þ formed by the shear effect of the large-scale tumble flow, hence,
1
j lnðy Þ þ B if yþ P 11
þ
it is affected by the evolution of tumble flow including the formu-
where j ¼ 0:41 is the von Kármán constant and B = 5.2 is the log- lation, moving and breaking process. Fig. 11 shows the evolution of
law constant. the viscous scale and the friction velocity. The viscous scale
Fig. 10 shows the dimensionless velocity profiles at selected increases and the friction velocity decreases at the beginning of
CADs during the compression stroke. In this study, the length of the compression stroke, which is mainly caused by the decrease
the near-wall velocity profiles is fixed at 3 mm, so the starting of the velocity gradient at the wall before flow reversing. Both
point and the length of the dimensionless curves are different at parameters show strong irregularity at the instant of the intake
different piston positions. Overall, the simulated results agree well valve closing. After the near-wall flow is relatively stabilized, the
with the standard wall function in the viscous sublayer (y+ < 5) and in-cylinder gas is compressed as the piston moves upward and

Fig. 10. The dimensionless velocity profiles at selected CADs. The dashed line refers
to the widely used standard wall function based on Eq. (4). Fig. 11. Evolution of the viscous scale and the friction velocity.
1196 X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199

leads to a rapid decrease in the kinematic viscosity, which finally


results in a gradual decrease of the viscous scale and the friction
velocity at the end of the compression stroke.

3.2. Thermal boundary layer

The following discussion focuses on the thermal boundary


layer. The complex local wall heat transfer process is analyzed
based on the ensemble-averaged and fluctuation temperature
and velocity distribution in the near-wall region. The same region
as the previous discuss for velocity boundary layer is used.

3.2.1. Temporal evolution of the thermal boundary layer


Fig. 12 shows the ensemble-averaged temperature profiles dur-
Fig. 13. Evolution of the thermal boundary layer thickness.
ing the compression stroke. The intake temperature is 313 K and
the wall temperature on the cylinder head is fixed at 348 K. The
gas temperature is lower than the wall temperature at the begin- curve has a peak value within 0.5 mm spacing from the wall, indi-
ning of the compression stroke, and becomes higher after 270 cating that the temperature fluctuation firstly increases and then
CAD because the gas is heated up. The temperature gradient at decreases with the increase of the distance from the wall, and
the wall increases towards TDC. The temperature profile varies the fluctuation intensity is almost constant after 2 mm spacing
non-monotonically from the wall at 260 CAD. The gas temperature from the wall. As the piston upwards, the peak value increases
slightly increases within 0.2 mm from the wall and then decrease. and the position of the peak moves towards the wall. The maxi-
The temperature variation is caused due to the advection by the mum is around 65 K at 0.2 mm spacing from the wall at TDC.
velocity field. This behavior indicates that the large-scale flow The temperature fluctuation is associated with the convective heat
structures have a great impact on the temperature distribution in flux in the in-cylinder flow, which will be discussed in
the near-wall region. The classic standard wall function cannot Section 3.2.3.
reflect non-monotonic variation of the temperature boundary layer
[31]. For all piston positions, the temperature gradient tends to 3.2.2. Dimensionless temperature profiles
zero with the increase of the distance from the wall and the tem- The ensemble-averaged temperature profiles were non-
perature stabilized at around 1 mm spacing from the wall. dimensionalized and compared with the universal thermal law of
The thermal boundary layer thickness is defined as the distance the wall. The thermal law of the wall was derived under the same
between the wall and the point where the gas temperature reaches ideal conditions as the momentum law of the wall with a constant
to 99% of the main stream gas temperature (calculated in the range inlet temperature and wall temperature. The friction temperature
from 1 mm to 3 mm spacing from the wall in this study). Fig. 13 is defined as:
shows the evolution of thermal boundary layer thickness during 

the compression stroke. The thickness increases as the excess tem- kw @@yT 
perature trends to zero and reaches to a peak value of 1.7 mm at Ts ¼ w
ð17Þ
qw cp us
260 CAD. After that, the thermal boundary layer thickness drops
sharply around 270 CAD because of the gas temperature and the where kw is the thermal conductivity at the wall and cp is the speci-
wall temperature are almost the same. Then the thickness fic heat capacity of the gas. The dimensionless temperature is
increases again as the temperature gradient reverses and the T þ ¼ ðT  T w Þ=T s . The near-wall temperature profiles can be
excess temperature increases. At around 310 CAD, the thickness described by the following formulas:
decrease gradually because of the local wall temperature gradient (
increases. Pryþ if yþ < 11
Tþ ¼ þ
ð18Þ
Although the mean temperature profiles are much more regular Pr t ðj lnðy Þ þ BÞ þ P
1
if yþ P 11
than the mean velocity profiles, the temperature fluctuation in the
near wall region could not be ignored as shown in Fig. 14. Each

Fig. 14. Temperature fluctuation profiles at selected CADs during the compression
Fig. 12. Ensemble-averaged temperature profiles at selected CADs. stroke.
X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199 1197

where j ¼ 0:41 is the von Kármán constant and B = 5.2 is the log-
law constant. Pr is the laminar Prandtl number, Pr t is the turbulent
Prandtl number, and the term P is a function of Pr and Prt .
Fig. 15 shows the dimensionless temperature profiles, y+ was
plotted in logarithmic scale and T+ was divided by the laminar
Prandtl number. The dimensionless temperature profiles agree
well with the thermal law of the wall in the viscous sublayer,
although the curve has an obvious positive offset from the dashed
line at 360 CAD because the first grid point exceeds y+ = 5. In addi-
tion, all the curves tend to be horizontal in the log layer and grad-
ually increase as the piston moves upward. T+/Pr reaches to a peak
value of around 12 at TDC.
Fig. 16 shows the temporal evolution of the friction tempera-
ture and Prandtl number. The friction temperature is inversely pro-
portional to the friction velocity according to Eq. (17), so it
Fig. 16. Temporal evolution of the friction temperature and Prandtl number.
increases at the beginning of the compression stroke mainly due
to the decrease of the friction velocity while the gas temperature The heat transfer coefficient is defined as
is almost constant. The temperature gradient direction reverses qw
around 260 CAD because the gas is compressed by the moving pis- h¼ ð19Þ
Tv  Tw
ton, so the friction temperature decreases in this stage. After that,
the friction temperature gradually increases because of the where T v is the cylinder-averaged temperature and T w is the wall
increase of the local wall heat flux and reaches to a peak of about temperature of 348 K. As shown in Fig. 17(b), the heat transfer coef-
42 K at TDC. The Prandtl number reflects the ratio of momentum ficient decreases from 100 W/(m2 K) at BDC to 30 W/(m2 K) around
diffusion and heat diffusion in the near-wall flow, and it is calcu- 260 CAD. The reason is that the excess temperature and the local
lated from the dynamic viscosity, thermal conductivity and specific wall heat flux decrease concurrently at this stage. After 270 CAD,
heat capacity of the gas at wall-bound temperature using the gas temperature is higher than 348 K, the excess temperature
Pr ¼ lcp =k. It can be seen that the Prandtl number rises from and local wall heat flux increase simultaneously which results in
0.705 to 0.727 during the compression stroke in Fig. 16. The an increase of heat transfer coefficient after 280 CAD. The maximum
Prandtl number was often set to be a constant in most flow and heat transfer coefficient is about 200 W/(m2 K) at TDC. As men-
heat transfer simulations in ICEs and care should be taken in using tioned earlier (in Section 3.2.1), the direction of temperature gradi-
such simplification in future simulations. ent reverses at around 260 CAD, so another peak at 270 CAD is
caused by a large wall temperature gradient and the small temper-
ature difference between the engine core region and the wall.
3.2.3. Wall heat transfer
Heat transfer mechanism in the near wall region of ICEs includes
A main purpose of studying transient velocity and thermal
gas-phase convection and fuel film conduction if the influence of
boundary layers in ICEs is to understand the complex heat transfer
high-temperature gas and soot radiation is ignored [4]. Therefore,
process. In this work, the local wall heat flux is calculated based on
the one-dimensional heat flux equation is written as [51]
the Fourier law, thus, thermal conductivity is calculated from Eq.
(19) and the temperature gradient at the wall is estimated accord- @T
q ¼ qcp hv 0 T 0 i þ k ð20Þ
ing to the ensemble-averaged temperature profiles. Fig. 17(a) @z
shows the evolution of local wall heat flux during the compression
where the first term in Eq. (20) is the convective term which is also
stroke. A negative value means the in-cylinder gas absorbs heat
known as the turbulent heat flux and the second term is the con-
from the cylinder head and a positive value is just the opposite.
ductive term. More specifically, v 0 and T 0 represent the wall-
The in-cylinder gas release heat after 260 CAD because the excess
normal direction velocity fluctuation and the temperature fluctua-
temperature changes to positive. The local heat flux reaches to a
tion. v 0 T 0 reflects to the correlation between the velocity and ther-
peak around 70 kW/m2 at TDC, and the maximum heat absorption
mal boundary layers. q refers to the gas density and cp refers to
is about-5 kW/m2 at BDC.
the specific heat capacity, these two parameters are treated as a
function of the distance from the wall.
Fig. 18 shows the spatial distribution of turbulent heat flux at
several piston positions after 300 CAD. The turbulent heat flux
increases as the piston moves upward and the local wall heat flux
increases at the same time. For any piston position, the turbulent
heat flux profile has a peak value near the wall, and the peak value
increases and moves towards the wall as the piston moves upward.
The black dashed line connects the peak point of all the curves and
shows the moving tendency of the peak location. According to Eq.
(20), the turbulent heat flux distribution depends strongly on the
correlation between the wall-normal velocity fluctuation and tem-
perature fluctuation [52]. Compared with the temperature fluctua-
tion peak point in Fig. 14, the peak location of the turbulent heat
flux is farther from the wall. Generally, most peaks are located
within the buffer layer which reflects that the buffer layer is the
main region for turbulence heat transfer. The maximum turbulent
Fig. 15. The dimensionless temperature profiles at selected CADs during the
compression stroke. The bold dashed line refers to the viscous sublayer according to
heat flux is about 70%85% of the local wall heat flux, and this data
Eq. (18). The buffer layer and the log layer are not plotted because of a lack of proper proved that forced convection heat transfer is the main heat trans-
experience parameters. fer mechanics in ICEs.
1198 X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199

Fig. 17. Evolution of local wall heat flux and heat transfer coefficient. (a) Temporal evolution of heat flux during the compression stroke, (b) Temporal evolution of heat
transfer coefficient during the compression stroke.

moves upward, but the evolution of the thermal boundary layer


thickness is non-monotonic under the combined impact of excess
temperature and temperature gradient, the peak thickness is
1.7 mm at 260 CAD. The dimensionless temperature profiles agree
well with the thermal law of the wall in the viscous sublayer and
tend to be uniform after that. In addition, the friction temperature
and Prandtl number increase concurrently after the near-wall flow
is relatively stabilized.
The in-cylinder gas firstly absorb heat from the cylinder head
and release heat after 260 CAD, the peak local wall heat flux is
around 70 kW/m2 at TDC. The heat transfer coefficient decreases
before 260 CAD and gradually increases to a peak around 200 W/
(m2 K) at TDC. The spatial distribution of turbulent heat flux shows
a strong link between the velocity boundary layer and thermal
boundary layer, and the peaks located in the buffer layer occupy
Fig. 18. Turbulent heat flux profiles at selected CADs.
about 70%–85% of the local wall heat flux, which indicates that
the forced convection heat transfer is the main heat transfer
mechanic in internal combustion engines.
4. Conclusions
Acknowledgments
In this work, the transient velocity and thermal boundary layers
with wall heat transfer during the compression stroke in a motored The study is financially supported by the National Science Fund
SIDI-engine were predicted using DDES-SST model. The simulated for Distinguished Young Scholars (No. 51525603).
results of the in-cylinder macro-flow field agreed well with the
high-speed 2D3C-PIV experimental data in the tumble plane. We Conflict of Interest Statement
firstly analyzed the temporal evaluation of the velocity and ther-
mal boundary layers, compared the dimensionless curves with We declare that we have no conflict of interest to this work. The
the standard wall functions, and predicted the evolution of local manuscript has been read and approved by all named authors
wall heat flux and the turbulent heat transfer in the near-wall and there are no other persons who satisfied the criteria for
region. From these results, the following conclusions can be drawn: authorship but are not listed. We confirm that the order of authors
The temporal evolution of the ensemble-averaged stream veloc- listed in the manuscript has been approved by all of us.
ity in the near-wall region is significantly influenced by the moving
piston and valves during the compression stroke. Flow separation References
and sub-millimeter vortex structures randomly occur in the veloc-
ity boundary layer. The significant changes of velocity fluctuation [1] M. Schmitt, C.E. Frouzakis, A.G. Tomboulides, Y.M. Wright, K. Boulouchos,
and Reynolds stress in different directions proved that the turbu- Direct numerical simulation of the effect of compression on the flow,
temperature and composition under engine-like conditions, Proc. Combust.
lence in the near-wall region is anisotropic. Inst. 35 (3) (2015) 3069–3077.
The dimensionless velocity profiles agree well with the law of [2] J.B. Heywood, Internal combustion engine fundamentals, McGraw-Hill, New
the wall in the viscous sublayer and deviate from the log layer, York, 1988.
[3] J. Chang, O. Güralp, Z. Filipi, D. Assanis, T.-W. Kuo, P. Najt, R. Rask, New heat
these differences come from the temporal complex flow and the transfer correlation for an HCCI engine derived from measurements of
variable fluid properties in ICEs. The thickness of the viscous sub- instantaneous surface heat flux, in: SAE Paper, 2004.
layer gradually extends and the dimensionless velocity in log- [4] Z. Han, R.D. Reitz, A temperature wall function formulation for variable-
density turbulent flows with application to engine convective heat transfer
layer increases as the piston moves upward. In addition, the vis- modeling, Int. J. Heat Mass Transf. 40 (3) (1997) 613–625.
cous scale and the friction velocity decrease concurrently after [5] G. Borman, K. Nishiwaki, Internal-combustion engine heat transfer, Progress in
the near-wall flow is relatively stabilized. Energy and Combust. Sci. 13 (1) (1987) 1–46.
[6] I. Celik, I. Yavuz, A. Smirnov, Large eddy simulat ions of in-cylinder turbulence
The ensemble-averaged temperature and the temperature gra-
for internal combustion engines: a review, Int. J. Engine Res. 2 (2) (2001) 119–
dient in the near-wall region gradually increase as the piston 148.
X. Fan et al. / International Journal of Heat and Mass Transfer 120 (2018) 1189–1199 1199

[7] M.J. Hall, F.V. Bracco, Cycle-resolved velocity and turbulence measurements [30] M. Jia, E. Gingrich, H. Wang, Y. Li, J.B. Ghandhi, R.D. Reitz, Effect of combustion
near the cylinder wall of a firing S.I. engine, in: SAE paper, 1986. regime on in-cylinder heat transfer in internal combustion engines, Int. J.
[8] D.E. Foster, P.O. Witze, Velocity measurements in the wall buondary layer of a Engine Res. 17 (3) (2015) 331–346.
spark-ignited research engine, in: SAE paper, 1987. [31] P.C. Ma, T. Ewan, C. Jainski, L. Lu, A. Dreizler, V. Sick, M. Ihme, Development and
[9] P.H. Pierce, J.B. Ghandhi, J.K. Martin, Near-wall velocity characteristics in analysis of wall models for internal combustion engine simulations using
valved and ported motored engines, in: SAE paper, 1992. high-speed micro-PIV measurements, Flow, Turbul. Combust. (2016) 1–27.
[10] A.Y. Alharbi, High-speed high-resolution vector field measurements and [32] R. Jhavar, Using Large Eddy Simulation to Study Diesel DI-HCCI Engine Flow
analysis of boundary layer flows in an internal combustion engine, Structure, University of Wisconsin - Madison, Wisconsin, Mixing and
University of Michigan, 2010. Combustion, 2007.
[11] A.Y. Alharbi, V. Sick, Investigation of boundary layers in internal combustion [33] C. Hasse, V. Sohm, B. Durst, Numerical investigation of cyclic variations in
engines using a hybrid algorithm of high speed micro-PIV and PTV, Exp. Fluids gasoline engines using a hybrid URANS/LES modeling approach, Comput.
49 (4) (2010) 949–959. Fluids 39 (1) (2010) 25–48.
[12] V. Sick, High speed imaging in fundamental and applied combustion research, [34] F. Hartmann, S. Buhl, F. Gleiss, P. Barth, M. Schild, S.A. Kaiser, C. Hasse, Spatially
Proc. Combust. Inst. 34 (2013) 3509–3530. resolved experimental and numerical investigation of the flow through the
[13] L. Lu, V. Sick, High-speed particle image velocimetry near surfaces, J. intake port of an internal combustion engine, Oil & Gas Science and
Visualized Exp. 76 (2013). Technology – Revue d’IFP Energies nouvelles 71 (1) (2016) 1–129.
[14] C. Jainski, L. Lu, A. Dreizler, V. Sick, High-speed micro particle image [35] S. Buhl, F. Gleiss, M. Köhler, F. Hartmann, D. Messig, C. Brücker, C. Hasse, A
velocimetry studies of boundary-layer flows in a direct-injection engine, Int. Combined Numerical and Experimental Study of the 3D Tumble Structure and
J. Engine Res. 14 (3) (2012) 247–259. Piston Boundary Layer Development During the Intake Stroke of a Gasoline
[15] M. Koehler, D. Hess, C. Brücker, Flying PIV measurements in a 4-valve IC engine Engine, Flow, Turbul. Combust. (2016) 1–22.
water analogue to characterize the near-wall flow evolution, Meas. Sci. [36] J.M. Wallace, Highlights from 50 years of turbulent boundary layer research, J.
Technol. 26 (12) (2015) 125302. Turbul. 13 (53) (2012) 1–70.
[16] R. Lucht, M. Maris, Cars measurements of temperature profiles near a wall in [37] M.S. Shadloo, A. Hadjadj, F. Hussain, Statistical behavior of supersonic
an internal combustion engine, in: SAE paper, 1987. turbulent boundary layers with heat transfer at M1=2, Int. J. Heat Fluid Flow
[17] R. Lucht, D. Dunn-Rankin, T. Walter, T. Dreier, S. Bopp, Heat transfer in engines: 53 (2015) 113–134.
comparision of cars thermal boundary layer measurements and heat flux [38] M.S. Shadloo, A. Hadjadj, D.J. Bodony, F. Hussain, S.K. Lele, Effects of heat transfer
measurements, in: International Congress & Exposition, 1991. on transitional states of supersonic boundary layers, center for turbulence
[18] M.E. Cundy, Development of high-speed laser diagnostics for the investigation research, Stanford University, Proc. Summer Program (2016) 175–184.
of scalar heterogeneities in engines, University of Michigan, 2012. [39] X. Wu, P. Moin, Transitional and turbulent boundary layer with heat transfer,
[19] M. Cundy, P. Trunk, A. Dreizler, V. Sick, Gas-phase toluene LIF temperature Phys. Fluids 22 (8) (2010) 1–8.
imaging near surfaces at 10 kHz, Exp. Fluids 51 (5) (2011) 1169–1176. [40] M. Schmitt, C.E. Frouzakis, Y.M. Wright, A.G. Tomboulides, K. Boulouchos,
[20] E.J. Lyford-Pike, J.B. Heywood, Thermal boundary layer thickness in the Direct numerical simulation of the compression stroke under engine-relevant
cylinder of a spark-ignition engine, Int. J. Heat Mass Transf. 27 (10) (1984) conditions: evolution of the velocity and thermal boundary layers, Int. J. Heat
1873–1878. Mass Transf. 91 (2015) 948–960.
[21] K.H. Tran, P. Guibert, C. Morin, J. Bonnety, S. Pounkin, G. Legros, Temperature [41] M. Schmitt, C.E. Frouzakis, Y.M. Wright, A.G. Tomboulides, K. Boulouchos,
measurements in a rapid compression machine using anisole planar laser- Investigation of wall heat transfer and thermal stratification under engine-
induced fluorescence, Combust. Flame 162 (10) (2015) 3960–3970. relevant conditions using DNS, Int. J. Engine Res. 17 (1) (2015) 63–75.
[22] O. Nijeweme, J. Kok, C. Stone, L. Wyszynski, Unsteady in-cylinder heat transfer [42] M. Schmitt, C.E. Frouzakis, Y.M. Wright, A. Tomboulides, K. Boulouchos, Direct
in a spark ignition engine: Experiments and modelling, Proc. Inst. Mech. numerical simulation of the compression stroke under engine relevant
Engineers, Part D: J. Automobile Eng. 215 (6) (2001) 747–760. conditions: Local wall heat flux distribution, Int. J. Heat Mass Transf. 92
[23] E. Gingrich, J. Ghandhi, R.D. Reitz, Experimental investigation of piston heat (2016) 718–731.
transfer in a light duty engine under conventional diesel homogeneous charge [43] M. Bellec, A. Toutant, G. Olalde, Large Eddy Simulations of thermal boundary
compression ignition, and reactivity controlled compression ignition layer developments in a turbulent channel flow under asymmetrical heating,
combustion regimes, SAE Int. J. Engines 7 (1) (2014) 375–386. Comput. Fluids (2016) 1–18.
[24] P.C. Ma, M. Greene, V. Sick, M. Ihme, Non-equilibrium wall-modeling for [44] P.R. Spalart, young person’s guide to detached-eddy simulation grids, 2001.
internal combustion engine simulations with wall heat transfer, Int. J. Engine [45] F.D. Mare, R. Knappstein, M. Baumann, Application of LES-quality criteria to
Res. (2017) 1–11. internal combustion engine flows, Comput. Fluids 89 (2) (2014) 200–213.
[25] C.J. Rutland, Large-eddy simulations for internal combustion engines - a [46] S.B. Pope, Turbulent Flow, Cambridge University Press, UK, Cambridge, 2000.
review, Int. J. Engine Res. 12 (2011) 1–31. [47] K. Liu, D.C. Haworth, Development and assessment of POD for analysis of
[26] H. Schlichting, K. Gersten, boundary layer theory, 8th revised and, enlarged turbulent flow in Piston Engines, in: SAE paper, 2011.
edition., Springer, Berlin, 2003. [48] D. Lengani, D. Simoni, M. Ubaldi, P. Zunino, F. Bertini, Turbulent boundary
[27] C. Rakopoulos, G. Kosmadakis, E. Pariotis, Critical evaluation of current heat layer separation control and loss evaluation of low profile vortex generators,
transfer models used in CFD in-cylinder engine simulations and establishment Exp. Thermal Fluid Sci. 35 (8) (2011) 1505–1513.
of a comprehensive wall-function formulation, Appl. Energy 87 (5) (2010) [49] R. Dekou, J.M. Foucaut, S. Roux, M. Stanislas, J. Delville, Large scale
1612–1630. organization of a near wall turbulent boundary layer, Int. J. Heat Fluid Flow
[28] S. Keum, H. Park, A. Babajimopoulos, D.N. Assanis, D. Jung, Modelling of heat 61 (2016) 12–20.
transfer in internal combustion engines with variable density effect, Int. J. [50] U. Piomelli, Wall-layer models for large-eddy simulations, Prog. Aerosp. Sci. 44
Engine Res. 12 (6) (2011) 513–526. (6) (2008) 437–446.
[29] F. Berni, G. Cicalese, S. Fontanesi, A modified thermal wall function for [51] F. Nicoud, numerical study of a channel flow with variable properties, Center
the estimation of gas-to-wall heat fluxes in CFD in-cylinder simulations of for Turbul. Res. (1998) 289–310.
high performance spark-ignition engines, Appl. Therm. Eng. 115 (2017) [52] H. Zhao, A. Wei, K. Luo, J. Fan, Direct numerical simulation of turbulent
1045–1062. boundary layer with heat transfer, Int. J. Heat Mass Transf. 99 (2016) 10–19.

S-ar putea să vă placă și