Anda di halaman 1dari 11

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

SAE TECHNICAL
PAPER SERIES

2008-01-0973

Conjugate Heat Transfer in CI Engine


CFD Simulations
Mika Nuutinen, Ossi Kaario and Martti Larmi
Helsinki University of Technology

Reprinted From: Multi-Dimensional Engine Modeling, 2008


(SP-2171)

2008 World Congress


Detroit, Michigan
April 14-17, 2008
400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-0790 Web: www.sae.org

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

By mandate of the Engineering Meetings Board, this paper has been approved for SAE publication upon
completion of a peer review process by a minimum of three (3) industry experts under the supervision of
the session organizer.
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise,
without the prior written permission of SAE.
For permission and licensing requests contact:
SAE Permissions
400 Commonwealth Drive
Warrendale, PA 15096-0001-USA
Email: permissions@sae.org
Tel:
724-772-4028
Fax:
724-776-3036

For multiple print copies contact:


SAE Customer Service
Tel:
877-606-7323 (inside USA and Canada)
Tel:
724-776-4970 (outside USA)
Fax:
724-776-0790
Email: CustomerService@sae.org
ISSN 0148-7191
Copyright 2008 SAE International
Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE.
The author is solely responsible for the content of the paper. A process is available by which discussions
will be printed with the paper if it is published in SAE Transactions.
Persons wishing to submit papers to be considered for presentation or publication by SAE should send the
manuscript or a 300 word abstract of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE.
Printed in USA

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

2008-01-0973

Conjugate Heat Transfer in CI Engine CFD Simulations


Mika Nuutinen, Ossi Kaario and Martti Larmi
Helsinki University of Technology
Copyright 2008 SAE International

ABSTRACT
The development of new high power diesel engines is
continually going for increased mean effective pressures
and consequently increased thermal loads on
combustion chamber walls close to the limits of
endurance. Therefore accurate CFD simulation of
conjugate heat transfer on the walls becomes a very
important part of the development. In this study the heat
transfer and temperature on piston surface was studied
using conjugate heat transfer model along with a variety
of near wall treatments for turbulence. New wall
functions that account for variable density were
implemented and tested against standard wall functions
and against the hybrid near wall treatment readily
available in a CFD software Star-CD.

conditions, by Urip, Yang and Arici [4], who studied


conjugate heat transfer in actual internal combustion
engine CFD simulations, by Tiainen, Kallio, Leino and
Turunen [5], who studied heat transfer in diesel engines
with density dependent wall functions in CFD and
combining the obtained average heat transfer
coefficients to FEM calculations of the heat transfer in
solid piston material and also by Huuhilo [6], who studied
and developed a FEM code for transient heat transfer in
the piston surface. A good motivation for this work is
figure 1 from Huuhilos Masters Thesis that shows the
effect of the piston surface material on the transient
maximum piston surface temperature.

INTRODUCTION
Accurate prediction of the piston surface temperature
and heat flux is difficult mainly because of the very large
temperature and velocity gradients of highly turbulent
flow near the wall and also because of the transient
interaction between the gas phase and solid wall
temperatures. Instead of resolving the temperature and
velocity profiles down to the wall, wall functions based on
some simplifying assumptions of the near wall turbulence
are used. Standard wall functions and hybrid near wall
treatment in the form they are usually implemented in
commercial CFD codes assume incompressible flow.
Therefore straightforward adoption of them into engine
simulations might not be appropriate, since the gas in the
cylinder is highly compressible indeed and large density
variations are likely to appear near the walls. To tackle
this problem momentum and energy equations near the
wall are integrated in their compressible form to new
modified wall functions that are sensitive to density
variation. Also variable turbulent Prandtl number near the
walls is included in the modified wall functions.
Heat transfer in combustion engines has bee widely
studied, e.g. by Schubert, Wimmer and Chamela [1],
who developed quasi-dimensional heat transfer models
for combustion ignition engines, by Han and Reitz [2],
who studied the effect of density variations on convective
heat transfer, by Urip, Liew, Yang and Arici [3], who
studied the effect of unsteady thermal boundary

Fig. 1. The simulated maximum temperature of the


piston surface layer fabricated from steel, aluminium or
zirconium oxide [6].

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

VARIABLE DENSITY WALL FUNCTIONS


In deriving the wall functions following assumptions are
made:
1) wall normal derivatives are much larger than
tangential
ones, so the tangential derivatives are
neglected.
2) near wall fluid flow is tangential to wall.
3) near wall pressure gradient can be neglected.
4) near wall heat flux consists only of laminar and
turbulent conduction.
5) gas obeys ideal gas law
6) near wall specific heat capacity is constant
7) near wall mass fractions of mixture components are
constant.
8) near wall shear stress and heat flux are constant.
Now the simplified near wall momentum and heat
equations (1) and (2) can be written as

( + t ) du = w ,

(1)

dy

dT
= qw .
c p + t
Pr Prt dy

u+ =

T+ =

y w /

u uw

w /

yu f

(2)

(3)

u uw
uf

c p w / (T Tw )
qw

(4)

c p u f (T Tw )
qw

. (5)

For turbulent viscosity Mellor [7] suggested that

t
(y + )
= + =

(y + )3 + 382.5

Prt =

0 .7
+ 0.85
Pr +

(7)

In [2] Han and Reitz made a change of variables from y


+
+
+
to y , u to u and T to T by using equations (3-5) directly,
but with the assumption

dy +
dy =
u f

(8)

which is not strictly correct, since in compressible flow


the gas density and consequently friction velocity uf are
functions of temperature and temperature in turn is a
function of the wall distance y. Only after the change of
variables was density treated as a variable but friction
velocity was still treated as a constant.
To overcome these shortcomings, new dimensionless
parameters are obtained by replacing variable density in
equations (3-5) with a reference density from point yp
that lies in the center of the near wall cell. With this
choice, the new dimensionless parameters are

y =

y p w / p

y p u f , p

Equations (1) and (2) are conveniently solved in nondimensional form. In constant density formulation, which
is the basis of the standard wall functions, the nondimensional velocity, temperature and wall distance are
written as

y+ =

Usually the turbulent Prandtl number is treated as a


constant, but Kays [8] formulated a near wall correlation
as

(6)

(9)

u f , p = w / p C 4 k p 2
u =

T =

u uw

w / p

u uw
u f ,p

c p , p p u f , p (T Tw )
qw

(10)

(11)

The dimensionless parameters are now functions of only


their dimensional counterparts and the change of
variables can be made correctly. The temperature enters
equations (1) and (2) through turbulent viscosity and
turbulent Prandtl number, equations (6) and (7), since
+
they depend on y which is a function of temperature.
*
Inserting the new dimensionless distance y and using
the ideal gas law equations (6) and (7) become

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

Tp

y
T

Now the iterative process is to repeat equations (12, 13,

Tp

y + 382.5
T

( y )
=
( y ) + 382.5 (12)

Tp

h=

T p

{ }

1
Pr
+
Pr 0.7 + 0.85 Pr

trapezoid round, since

(19)

T p

c p, p p u f , p

T p as

(20)

Dimensionless velocity can be integrated in a single

(14)

dy

c p , p p u f , p (T p Tw )

(13)

Rearranging equations (1) and (2) they become

dT =

coefficient can be computed from the obtained

dy
,
1 +

T p changes less than a

specified tolerance. Wall heat flux or heat transfer

qw =

0 .7
Prt =
+ 0.85 .
Pr

du =

17 and 18) in a sequence until

(15)

u i = u i1 + (Bi + Bi 1 )

is now known

y i
2 ,

(21)

u =0
where Bi is the integrand of eq. (14). Wall shear stress
can be computed from the obtained

u p

Equations (14) and (15) are coupled by , so their


integration is not straightforward, but has to be made
both numerically and iteratively.

INTEGRATION OF WALL FUNCTIONS


To integrate the equations (14) and (15) an initial
*
discretized profile for has to be guessed e.g.
assuming linear temperature profile

i =

Tp

Tw + (T p Tw )y / y

(16)

The discretized form of eq. (15) can be integrated with


trapezoid method

Ti* = Ti 1 + ( Ai + Ai 1 )

y
2 ,

(17)

T =0

(22)

With the assumption that Han & Reitz [2] made,


temperature variation cancels from the momentum
equation, so the dimensionless velocity can be computed
by using just the Mellor profile for turbulent viscosity.
Dimensionless temperature becomes according to Han
& Reitz [2]

T p+ =

p c p , p Tw ln (T p / Tw )
qw

( )

= 2.1 ln y +p + 2.5 (23)

COMPARISON OF WALL FUNCTION FORMULATIONS


In figure 2 is a comparison between the velocity laws of
wall in laminar, logarithmic (turbulent), incompressible
Mellor and the new variable density formulations in a
typical engine simulation situation with

y p = 200

where Ai is the integrand of eq. (15). Because the wall


*
heat flux is constant, can be re-evaluated

i =

u p uw
.
w =
u p
p

T pT

TwT + Ti (T p Tw )

(18)

Tw = 700 K
T p = 1500 K
Pr = 0.7
It can be seen that the laminar and logarithmic curves
*
intersect at y = 10.9577, which is a switching point from
turbulent to laminar law of wall in standard high Reynolds

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

In figures 3 and 4 is a comparison between the


temperature laws of the wall in constant density Mellor &
Kays, variable density Han & Reitz and the new variable
density formulations. The new formulation and Han &
*
Reitz formulation both predict smaller T values than the
constant density formulation, but the difference between
the two variable density formulations seems quite small
and this raises a question if there is any reason to use
the new, computationally more expensive formulation.
The answer is yes. Inspecting equations (11) and (23) it
is noted that the dimensionless temperatures are defined
differently in the two formulations. Comparing the
predicted heat fluxes the difference of the two variable
density formulation becomes more pronounced. In figure
5 is a comparison between the predicted heat fluxes
relative to the constant density formulation heat flux as a
function of near wall cell center dimensionless distance.
There is a huge relative difference in the predicted heat
fluxes, so there is indeed motivation to use the new,
more rigorously formulated temperature law of the wall.
The conclusion is that the constant density law of the
wall will underestimate the heat flux whereas the variable
density Han & Reitz wall function tends to overestimate
it.

18
16
14

Han & Reitz

T*
8
6
4
2
0

10

10

10

y*

Fig. 3. Temperature wall functions.


14
13

12
11

10
9
Mellor & Kays incompressible

Current formulation
Han & Reitz

6
1
10

10
y

Fig. 4. Temperature wall functions.

1.8
Laminar

1.7

Turbulent (log)
14

Current formulation

10

18
16

Mellor & Kays incompressible

12

T*

number wall treatment. The profile computed with Mellor


turbulent viscosity transits smoothly from the laminar to
logarithmic curve and there is no need to define a
switching point. The new variable density formulation
*
follows smoothly the laminar curve at small values of y
and lags the constant density logarithmic profile and the
*
Mellor profile when y gets larger in a case where the gas
is hotter than the wall. If the temperature ratio was
inversed the new formulation curve would be above
logarithmic and Mellor profiles. Interpreting eq. (22) this
means that the new formulation tends to produce larger
near wall drag than the standard wall functions when the
gas is hotter than the walls and vice versa.

Mellor incompressible

Current formulation
Han & Reitz

1.6

Current formulation
12

qw/qw0

1.5

u*

10

1.4

1.3

1.2

1.1

10

10
y

Fig 2. Velocity wall functions.

10

0.9
0
10

10

y*

Fig. 5. Effect of wall functions on heat flux.

10

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

SIMULATION PROCEDURE

COMPUTATIONAL GRID

Simulations were carried out with Star-CD on a simplified


flat piston - cylinder model of Wrtsil 20 engine with
conjugate heat transfer CFD model. Detailed CFDmodels of Wrtsil 20 engine can be found e.g. in [9] by
Antila et. all. Since the aim of this study was to test the
conjugate heat transfer model along with different near
wall treatments, only the piston was modeled as a solid
block with three dimensions, as it receives the largest
thermal load, whereas cylinder and cylinder head were
modeled just as conducting walls with constant
temperature boundary condition. The lower boundary of
the thin piston layer was modeled as a constant
temperature wall and the peripheral boundary as an
adiabatic wall. The simulations were also confined to
compression and expansion strokes, so intake and
exhaust through valves were not considered. The
situation was also considered periodic so only a 45
(periodic angle of the 8-hole nozzle) slice of the cylinder
was modeled. Piston wall temperatures and heat fluxes
as well as cylinder pressures and heat release rate were
monitored in each simulation. Between simulations only
the turbulence models and near wall treatments were
changed, except for the low Reynolds number k- model
with hybrid wall treatment, where the mesh near the
piston wall was refined to meet the requirements of the
hybrid wall model. The specifications of the modeled
engine are presented in table 1.

The computational grid was a 45 section of the cylinder.


The solid piston mesh was 2mm thick (about 3 times the
estimated temperature penetration depth), with 30 cells
in axial direction and the height of the first cell next to the
solid-gas interface was approximately 5m and graded
with the expansion ratio of 1.2. In radial direction there
were 25 cells and in azimuth direction 20 cells, both
evenly spaced. The fluid domain had 10 evenly spaced
0.6mm thick cells next to the gas-solid interface and the
rest of the domain had 75 deforming cells in axial
direction. Part of the deforming cells were removed from
the simulation during the compression and then added
again during expansion to maintain a desired mesh
density. The mesh in the hybrid wall treatment case was
otherwise identical but the 10 0.6mm fluid cells were
replaced by 30 graded (expansion ratio 1.2) cell layers
with approximately 5m gas-solid interface cell
thickness. The coarser computational mesh at TDC is
presented in figure 7.

Table 1. Specifications for the simulated engine.


Bore [mm]

200

Stroke [mm}

280

Connecting rod [mm]

510

Displacement [dm ]

8.8

Engine speed [revs/min]

1000

Rated power [kW]

200

Compression ratio [-]

15

Injector nozzles [-]

Nozzle hole diameter [mm]

0.37

Fuel mass/cycle [g]

1.0

Fuel (name, formula)

Dodecane, C12H26

Lambda(relative air/fuel ratio) [-]

2.46

Initial swirl number

0.5

TURBULENCE AND COMBUSTION MODELS


Simulations were made with 4 combinations
turbulence models and near wall treatments:

of

1) High Reynolds number k- model with standard wall


functions.
2) High Reynolds number k- model with the new
variable density wall functions, computed in user
subroutines.
3) High Reynolds number RNG k- model with standard
wall functions.
4) Low Reynolds number k- model with hybrid wall
treatment. In hybrid wall treatment the flow field is
resolved all the way down to the wall without any wall
functions whenever the near wall mesh density is
sufficient and when it is not the wall function solution is
blended to the solution.
In all simulations Eddy Break Up LaTCT one reaction
model without any ignition model was used to compute
the reaction rates. Thus the ignition depends solely on
the reaction rates determined by the EBU LaTCT model.
Details of the used turbulence and combustion models
can be found in [10] and their application to diesel engine
simulation in Kaarios doctoral thesis [11].
LAGRANGIAN MULTI-PHASE MODELS
Injected droplets were simulated with Lagrangian multiphase model with turbulent dispersion and standard
models for momentum, heat and mass transfer. Reitz
and Diwakar droplet break-up model [12] was used
together with Bais [13] spay impingement model
including droplet-wall heat transfer and boiling.

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

Temperature dependent droplet properties were


computed in a user subroutine. The injection profile of a
single nozzle is presented in figure 6.
0.04
0.035

mass flow rate [kg/s]

0.03

0.025

0.02

0.015

0.01
0.005

0
350

355

360

365

370

375

380

385

CA

Fig. 6. Nozzle mass flow rate.


CONJUGATE HEAT TRANSFER
Conjugate heat transfer is very simple as the energy
equation in the solid part is just a degenerate form
(velocity is set to zero) of its counterpart in the fluid
domain. The code only needs to know that there is a
step jump in the material properties (heat conductivity,
density and specific heat capacity). When solving the
energy equation the code has to adjust the interface
temperature such that the heat flux across the surface is
continuous. However, due care needs to be taken in the
discretization of the solid domain, since the heat
conductivities of metals and thus the resulting
temperature gradients differ by orders of magnitude, e.g.
comparing the heat conductivities of nitrogen and steel:
knitrogen = 0.02583W/m/K, ksteel = 43W/m/K. This means
that the solid domain mesh near the interface has to be
very dense in the normal direction. Another important
assessment is the penetration depth of the oscillating
temperature. Huuhilo [6] used an approximation based
on an analytical solution of a 1-dimensional heat
equation with sinusoidally varying temperature boundary
condition

T ( z , t ) = T0 e z / cos t

2k
=
c p

SIMULATION RESULTS
Maximum piston surface temperature, total piston
surface heat transfer and cylinder pressure were
extracted from all 4 simulations. To avoid
misinterpretation, standard wall functions refer to those
constant density wall functions that are by default in StarCD. The maximum piston surface temperatures are
presented in figure 7. It is noted that the new modified
variable density wall functions (MWF) predict clearly
higher peak temperatures than the other standard (SWF)
and hybrid (HWF) wall functions. This information would
be invaluable when designing pistons for extreme
thermal loads. Total piston heat transfer is presented in
figure 8. The new wall functions also predict higher heat
transfer from the charge to the piston surface as
expected. Cylinder pressures are presented in figure 9.
The new modified wall treatment produces marginally
(approximately 0.2bar) higher peak pressure than the
corresponding simulation with standard wall functions,
which is surprising because more heat is flowing through
the piston with the new wall treatment. However the
relative difference is very small and can be explained by
inspecting the energy balance of the charge gas. In
figure 10 is the cumulative energy difference (heat
release - heat loss through boundaries) between
simulations with the new modified wall treatment and the
standard wall functions. It is noted that while the heat
loss is greater with the new wall treatment, also the heat
release from the burning charge is more rapid. The total
of these is positive between approximately CA 369 and
375, where the pressure also peaks and consequently
the maximum cylinder pressure is higher with the new
modified wall treatment. In figure 11 is a snapshot from
the piston surface temperature distribution and spray at
12 after the TDC, when the piston surface temperature
has its maximum value, approximately 656K. An order of
magnitude comparison can be made with figure 1, where
a same engine was modeled, although the shape of the
piston and running speed were different.
The CPU times of the simulations are listed in table 2.
Simulation with the hybrid wall treatment was the most
time consuming, but it also had a denser mesh than the
other three. There is no big difference between the
standard and RNG k-, but with the new modified wall
treatment the CPU time is increased some 20%. Of
course the increase in the CPU time depends on the
number of points used in numerical integration and the
tightness of the stopping criterion (now 200 and 0.5%).

(24)

where is the angular frequency of the temperature


variation and is the penetration depth where the
-1
temperature fluctuation is e times the amplitude. The
thickness of the modeled solid layer has to be at least
few times the penetration depth in order to avoid
disturbances from the bottom constant temperature
boundary condition.

Table 2. CPU times.


Turbulence model/wall treatment/cells

CPU time [s]

k- / SWF / 57500 cells

11744

RNG k- / SWF / 57500 cells

10844

k- / HWF / 67500

20109

k- / MWF / 57500 cells

14186

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

Cylinder pressure [Pa]

1.75

x 10

1.7

k SWF
RNG k SWF

1.65

k HWF
k MWF

1.6
360

364

366

368

370
CA

372

374

376

378

380

400
CA

410

420

430

440

450

Fig. 9. Cylinder pressures.

Fig. 6. Computational mesh.

50

Cumulative difference in (heat productionheat loss) [J]

660
650
Piston surface temperature, Tmax [K]

362

640
k SWF
630

RNG k SWF
k HWF

620

k MWF

610
600
590
580
570
360

380

400

420

440

460

480

500

CA

50

100

150

200

250
350

360

370

380

390

Fig. 10. Cumulative difference in energy balance in


simulations performed with the new modified and
standard wall functions (k- model).

Fig. 7. Maximum piston surface temperature.

250
k SWF

Total piston heat transfer [kW]

RNG k SWF
k HWF

200

k MWF

150

100

50

0
355

360

365

370

375

380

385

390

395

400

405

410

CA

Fig. 8. Total piston heat transfer.


Fig. 11. Piston surface temperature and spray.

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014

CONCLUSION
To conclude, the new variable density formulation for
wall functions works well and gives reasonable and
expected results compared to constant density wall
functions. That the new variable density wall functions
predict higher heat loss and higher cylinder pressure
simultaneously seems inconsistent, but the difference is
so small that it can be explained by the changes caused
by the new wall functions in turbulent kinetic energy and
dissipation rate that control the reaction rate in the used
EBU LaTCT combustion model. One could do even
better by including the effect of temperature dependent
specific heat capacity to the formulation, but this is left
for future work. However the quality of the modified wall
functions can not be verified without careful comparison
with experimental data, but that is beyond the scope of
this research.

ACKNOWLEDGMENTS

8. W. M. Kays, Turbulent Prandtl number where are


we?, ASME J. Heat Transfer 116, 284 (1994).
9. E. Antila, O. Kaario, P. Kilpinen, T. Lahtinen, M.
Larmi, H. Pokela, A. Saarinen, K. Stalsberg-Zarling,
P. Taskinen, J. Tiainen, O. Toivanen, Mastering The
Diesel Process, Publications of the Internal
Combustion Engine Laboratory, Helsinki University
of Technology 79, ISBN 951-22-6998-8, (2004).
10. CD-Adapco, METHODOLOGY, Star-CD version
3.26.
11. O. Kaario, The Influence of Certain Submodels on
Diesel Engine Modeling Results, Doctoral Thesis,
Publications of the Internal Combustion Engine
Laboratory, Helsinki University of Technology,
(2007).
12. R. D. Reitz et. R. Diwakar, Effect of drop breakup on
fuel sprays, SAE Technical Paper Series 860469,
(1986).
13. C. Bai and A. D. Gossman, Development of
methodology for spray impingement simulation, SAE
Technical Paper Series 950283, (1995).

The authors wish to thank the Helsinki University of


Technology Internal Combustion Engine Laboratory for
the opportunity and facilities to carry out this work.

CONTACT

REFERENCES

Helsinki University of Technology, Internal Combustion


Engine Laboratory, PO BOX 4300, FIN-02015 HUT,
Finland

1. C. Schubert, A. Wimmer and F. Chmela, Advanced


Heat Transfer Model for CI Engines, SAE Technical
Paper Series 2005-01-0695, (2005).
2. Z: Han et. R. D. Reitz, A temperature wall function
formulation for variable-density turbulent flows with
application to convective heat transfer modeling, Int.
J. Heat Mass Transfer. Vol 40, No. 3, 613-625,
(1997)
3. E. Urip, K. H. Liew, S. L. Yang and O. Arici,
Numerical Investigation of Heat Conduction with
Unsteady Thermal Boundary Conditions for Internal
Combustion Engine Application, Proceedings of
IMECE04 2004 ASME International Mechanical
Engineering Congress and Exposition, IMECE200459860, (2004).
4. E. Urip, S. L. Yang and O. Arici, Conjugate Heat
Transfer for Internal Combustion Engine Application
using KIVA code, Mechanical EngineeringEngineering Mechanics Department, Michigan
Technological University Houghton, Michigan 49931.
5. J. Tiainen, I. Kallio, A. Leino and R. Turunen, Heat
Transfer Study of a High Power Density Diesel
Engine, SAE Technical Paper Series 2004-01-2962,
(2004).
6. P. Huuhilo, Finite Element Analysis of Transient Heat
Transfer in the Piston Surface of Combustion
Engine, Masters Thesis, Helsinki University of
Technology, Department of Engineering Physics and
Mathematics, (2006).
7. G. L. Mellor, Proc. Symp. Fluidics Internal Flow,
Pennsylvania State University (1968).

Mika Nuutinen, Research Scientist

mika.nuutinen@tkk.fi

Downloaded from SAE International by Kwang Yang Industry Co Ltd, Wednesday, July 16, 2014
+

dimensionless velocity

dimensionless velocity with reference density

turbulence model constant = 0.09

distance to the wall

heat transfer coefficient

heat conductivity, turbulent kinetic energy

kt

turbulent heat conductivity

Pr

Prandtl number = cp/k

Prt

turbulent Prandtl number = tcp/kt

qw

wall heat flux

temperature

Tw

wall temperature

NOMENCLATURE

cp

specific heat

dimensionless distance to the wall

dimensionless distance to the wall with reference


density

Greek symbols

dimensionless temperature parameter (Tp/T)

von Karman constant = 0.419

molecular viscosity

turbulent viscosity

dimensionless temperature

dimensionless
density

temperature

with

magnitude of gas velocity

uf

friction velocity

uw

magnitude of tangential wall velocity

reference

dimensionless viscosity t/

density

wall shear stress

Anda mungkin juga menyukai