You are on page 1of 15

Proceedings of the Institution of Mechanical

Engineers, Part D: Journal of Automobile


Engineering
http://pid.sagepub.com/

Unsteady in-cylinder heat transfer in a spark ignition engine: Experiments and modelling
D J Oude Nijeweme, J. B. W. Kok, C. R. Stone and L Wyszynski
Proceedings of the Institution of Mechanical Engineers, Part D: Journal of Automobile Engineering 2001 215: 747
DOI: 10.1243/0954407011528329
The online version of this article can be found at:
http://pid.sagepub.com/content/215/6/747

Published by:
http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for Proceedings of the Institution of Mechanical Engineers, Part D: Journal of Automobile
Engineering can be found at:
Email Alerts: http://pid.sagepub.com/cgi/alerts
Subscriptions: http://pid.sagepub.com/subscriptions
Reprints: http://www.sagepub.com/journalsReprints.nav
Permissions: http://www.sagepub.com/journalsPermissions.nav
Citations: http://pid.sagepub.com/content/215/6/747.refs.html

>> Version of Record - Jun 1, 2001


What is This?

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

747

Unsteady in-cylinder heat transfer in a spark ignition


engine: experiments and modelling
D J Oude Nijeweme1, J B W Kok1, C R Stone2* and L Wyszynski2
1Department of Mechanical Engineering, University of Twente, The Netherlands
2Department of Engineering Science, University of Oxford, UK

Abstract: Instantaneous heat ux measurements have shown that, in the expansion stroke, heat can
ow from the wall into the combustion chamber, even though the bulk gas temperature is higher
than the wall temperature. This unexpected result has been explained by modelling of the unsteady
ows and heat conduction within the gas side thermal boundary layer. This modelling has shown
that these unsteady eVects change the phasing of the heat ux, compared with that which would be
predicted by a simple convective correlation based on the bulk gas properties. Twelve fast response
thermocouples have been installed throughout the combustion chamber of a pent roof, four-valve,
single-cylinder spark ignition engine. Instantaneous surface temperatures and the adjacent steady
reference temperatures were measured, and the surface heat uxes were calculated for motoring and
ring at diVerent speeds, throttle settings and ignition timings. To make comparisons with these
measurements, the combustion system was modelled with computational uid dynamics (CFD). This
was found to give very poor agreement with the experimental measurements, so this led to a review
of the assumptions used in boundary layer modelling. The discrepancies were attributed to assumptions in the law of the wall and Reynolds analogy, so instead the energy equation was solved within
the boundary layer. The one-dimensional energy conservation equation has been linearized and normalized and solved in the gas side boundary layer for a motored case. The results have been used
for a parametric study, and the individual terms of the energy equation are evaluated for their
contribution to the surface heat ux. It was clearly shown that the cylinder pressure changes cause
a phase shift of the heat ux forward in time.
Keywords: heat ux, spark ignition engines, computational uid dynamics (CFD)

NOTATION
A ,B
n n
c
h
k
k
l
n
p
P

Fourier series coeYcients


heat capacity
heat transfer coeYcient
thermal conductivity, equations (3 ), (6),
(15) to (17), ( 19), (21), (22)
turbulence kinetic energy, equations (9),
(10), (12)
depth of reference temperature
measurement
harmonic number
pressure
variable in equation (11)

The MS was received on 27 September 2000 and was accepted after


revision for publication on 29 January 2001.
* Corresponding author: Department of Engineering Science, University
of Oxford, Parks Road, Oxford OX1 3PJ, UK.
D09600 IMechE 2001

Pr
q
R
t
T
M
T
0
u
v
x
y
z

Prandtl number
heat ow per unit area
specic gas constant
time
semi-amplitude of temperature variation,
equations ( 1) and (2)
mean temperature, equations (1) and (2)
tangential uid velocity
normal velocity component
coordinate into surface
coordinate in the gas away from the wall
Lagrangian coordinate
thermal diVusivity
ratio of heat capacities
viscosity
density
time
wall shear stress

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

Proc Instn Mech Engrs Vol 215 Part D

748

D J OUDE NIJEWEME, J B W KOK, C R STONE AND L WYSZYNSKI

Subscripts
i
j
l
m
p
ref
t
v
w
?
1

spatial step
temporal step
laminar
mean
constant pressure
reference
turbulent
constant volume
wall
bulk gas
INTRODUCTION

Heat transfer aVects engine performance, eYciency and


emissions. For a given mass of fuel within the cylinder,
higher heat transfer to the combustion chamber walls
will lower the average combustion gas temperature and
pressure and therefore reduce the work per cycle transferred to the piston. It is generally accepted that about
30 per cent of the energy initially in the combustion
chamber is transferred to the cooling system, and about
half of this is due to in-cylinder heat transfer. On the
other hand, cooling is necessary because during combustion the gas temperatures can reach 2500 K or so. The
wall temperatures should not exceed about 300 C for
cast iron cylinder heads and about 200 C for aluminium
heads. The lubricated walls must also be kept below
about 200 C to prevent the oil lm from oxidizing.
These temperature diVerences between the bulk gas and
the walls lead to heat uxes that can reach 10 MW/m2
during combustion. Because of these high heat ux rates
thermal stresses occur which can cause cracks in the cylinder head. Cooling of the spark plug and the exhaust
valves is especially important because overheating of
these can cause pre-ignition.
A very important aspect of heat transfer is the strong
inuence it has on exhaust emissions. The formation rate
of nitric oxide ( NO) has an exponential dependence on
temperature, so a reduction in the peak combustion temperature of 2550 K can halve the NO emissions. Wall
x
temperatures are important for emissions as well.
According to Myers and Alkidas [1] NO emissions
x
increase signicantly with increasing surface temperatures. The hydrocarbon emissions, on the other hand,
were found to decrease considerably at a given airfuel
ratio when the coolant temperature was increased from
298 to 373 K.
2

EXPERIMENTAL MEASUREMENTS

2.1 Experimental equipment


The engine was a single-cylinder version of the Rover
K16, pent roof combustion chamber spark ignition
Proc Instn Mech Engrs Vol 215 Part D

engine, with a tumbling air motion. It has a bore of


80 mm, a stroke of 89 mm and a compression ratio of
10:1. Full details of the instrumentation on the engine
have already been published [2], so comment here will
be restricted to some of the less usual features.
The cylinder barrel was tted with a piezo-resistive
pressure transducer, so that when the piston was close
to bottom dead centre (BDC ) the transducer recorded
the cylinder pressure, thereby xing the pressure datum
for the piezoelectric pressure transducers in the cylinder
head. When the piston was about 20 crank angle (CA)
or further from BDC the barrel transducer recorded
the atmospheric pressure, thus enabling the absolute
pressure datum to be xed. The Kistler 4045A10 piezoresistive transducer was in a water-cooled mount (7505)
and was calibrated with a dead weight tester to give
1 V/bar. Two piezo-electric pressure transducers were
mounted in the cylinder head, a Kistler 701a in a watercooled mount and an AVL QC32D water-cooled transducer. Both transducers were dead weight calibrated
with their Kistler 5007 charge ampliers, and dynamic
testing in the engine showed no discernible diVerence
between their measurements.
Air owrate is notoriously diYcult to measure in a
single-cylinder engine, because of the unsteady nature of
the ow. For this reason a positive displacement air
owmeter has been used. The engine operating points
were dened in terms of the air owrate; the frictional
levels were high and not entirely repeatable, so this precluded the use of the brake torque for dening the engine
operating point. The manifold absolute pressure was
used as a check on the engine operating points.
A fuel injection system was used for liquid fuels (at
2 bar gauge) and methane (at 6 bar gauge) using a high
ow rate solenoid operated injector. The ratio of airfuel
to stoichiometric airfuel ratio, , was monitored by a
calibrated analyser. Over the range of airfuel ratios
tested, the airfuel ratio was known to an accuracy of
no worse than 1.5 per cent. The ignition timing was
controlled digitally, and its performance veried by an
ignition strobe and ywheel timing marks.
To be able to determine the surface heat ux accurately, a fast response surface temperature transducer is
needed to measure the wall temperature. As the Fourier
equation for the surface heat ux indicates, errors can
be caused by an inadequate response time and by the
conduction properties of the probe. Also, the fast
response probe should not disturb the gas ow in the
cylinder or the heat ow in the wall. Gatowski et al. [3]
investigated four diVerent types of fast response surface
temperature probes and concluded that eroding-type
surface thermocouples had essentially an instantaneous
response and were durable. Figure 1 shows the construction of the erodable duplex thermocouples used here.
The erodable thermocouple uses alumel and chromel
ribbon elements 25 m thick. These ribbon elements are
embedded in the probe (parallel to its axis) and separ-

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

D09600 IMechE 2001

UNSTEADY IN-CYLINDER HEAT TRANSFER IN A SPARK IGNITION ENGINE

Fig. 1

Construction of the duplex fast response surface thermocouple

ated by thin mica sheets. The end of the thermocouple


is abraded so that a tiny junction is formed by plastic
deformation of the ribbon elements. Any erosion of the
surface by the environment will form a new junction. In
order to establish the heat ux through the element a
second temperature measurement is needed. This reference thermocouple is embedded in the probe at a known
distance from the wall (4.76 mm) where the temperature
uctuations have decayed. To prove that the temperature
at that depth is constant, a worst-case scenario for heat
penetration was considered. Eckert and Drake [4] present the equations for heat transfer in a semi-innite
solid with periodic surface temperature variations:
2n
T =T cos
0

0M
0
/
n
is the period.
where
0
Eckert and Drakes solution is
T =T

0M

A S B A

exp

(1)

n
2n

x cos

0
0

S B
n
x

(2)

where is the thermal diVusivity of the probe material


and x is the distance into the probe. In this case with
aluminium, has a typical value of 7.49610 5 m2/s.
The temperature variation is of a shorter period than
the engine cycle, with many higher harmonics also present. If the fundamental period is assumed to be of equal
duration to one revolution, then the worst case is when
the cosine term becomes 1, and the engine is operating
at its lowest speed. For a speed of 750 r/min the fundamental temperature uctuation would have decayed to
D09600 IMechE 2001

749

0.76 per cent of its surface amplitude at the location of


the reference thermocouple. Twelve of these fast
response duplex thermocouples were installed around
the combustion chamber: four in the piston, six in the
cylinder head and two in the liner.
The thermocouples provide only a small voltage of
about 40 V/K each, and the surface temperature uctuations are only about 1 K when motoring, or 10 K
when ring. Great care is thus needed in the amplication and calibration of the signals. The output from
the reference thermocouple was taken to a digital display
(accuracy 1 K ), but the reference thermocouple was also
connected to the surface thermocouple to give an output
corresponding to the temperature diVerence. This way,
only the temperature diVerence was amplied, thereby
reducing errors compared with amplifying the signals
independently, and eliminating the need for cold junction compensation. The signals were taken by twisted
screened cables to precision instrumentation ampliers.
Unfortunately, because of the nature of a thermocouple junction, even thermocouples made from identical materials do not have the same output voltage
response. A potential is built up at the interface between
two dissimilar metals, owing to the diVerence in their
work functions. The lower work function metal loses
electrons to the higher, until a charged layer has built
up a potential to prevent further transfer of electrons.
The work functions depend on the band structure of the
metals, which in turn depends on their composition.
When a thermocouple junction is formed there will be
varying amounts of interpenetration of the two materials
(and possible contamination), so this aVects the band
structure and the temperaturevoltage response of the

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

Proc Instn Mech Engrs Vol 215 Part D

750

D J OUDE NIJEWEME, J B W KOK, C R STONE AND L WYSZYNSKI

thermocouple. In most applications a temperature


response diVerence between thermocouples of 0.1 K
would be negligible, but, when the temperature diVerence uctuations can be as low as 1 K, then individual
calibration is needed. Fortunately, what matters is the
diVerence in the outputs of the two thermocouples in a
duplex probe, and not their absolute valuesthis simplies their calibration. Prior to installation in the engine,
the duplex probes were mounted in a copper block and
calibrated with their corresponding leads and ampliers
by heating them in the range 50200 C in steps of about
25 K. The temperature diVerence signal (typically of
order 0.5 K ) was then curve tted against the calibration
temperature, for subsequent correction of the in-cylinder
diVerential temperature measurements.
The 12 diVerential temperature signals, along with
encoder ag and pressure signals, were logged every
degree CA by a 16-channel data acquisition card with
12-bit resolution. A nite diVerence method was applied
to calculate the temperature distribution within the walls
by solving the Fourier equation for unsteady heat
conduction and from that deducing the heat transfer.

To solve the unsteady heat conduction equation, it is


necessary to treat the problem as one dimensional. This
is acceptable for the following reasons: the temperature
gradient normal to the wall is large compared with the
temperature gradients along the wall; temperature
diVerences along the wall decay rapidly; the silicone sealant and/or cyano-acrylate adhesive used to secure the
duplex probe and the stainless steel tube around the
probe tend to insulate the probe body; and stainless steel
has a thermal conductivity at least 10 times lower than
the aluminium alloy used.
The Fourier equation yields for the one-dimensional
case

A B

AS B
S B
S BD

x
N
T(x, t) =T (T T ) + ~ exp x
m m
ref l
n= 1
n
6 A cos nt x
n
2

+B sin nt x
n

n
2

n
2

(5)

This equation, diVerentiated with respect to x and substituted in the one-dimensional version of the Fourier equation, gives for the heat ux at the surface (where x=0 )
N
q =q +k ~
s
m
n= 1

n
[(A +B ) cos (nt )
n
n
2
+ (B A ) sin (nt)]
n
n

(6)

where
T T
ref
q =k m
m
l

2.2 Heat ux calculations and measurements

qT
1 q
qT
=
k
rc qx
qt
qx

where T is the steady state temperature at the reference


ref
point at distance l from the surface (x = 0). So the
solution of equation (4) is

(3)

The most popular method to solve equation (3) is to use


a Fourier transform, as performed by Alkidas [5]. A
sinusoidal variation of heat ux with time into a semiinnite solid produces the same frequency variation in
surface temperature, displaced in phase by 90. So the
surface temperature can be expressed as a Fourier series:

with q the steady state heat ux. So, if the reference


m
temperature and the temperature variation at the surface
are measured, a Fourier transform gives the temperature
eld T =T (x, t), and subsequently the surface heat ux.
In contrast the approach used here has been to adopt
a nite diVerence solution that has been solved explicitly,
since this is simpler and more intuitive for programming.
In nite diVerence form equation (3) becomes
1
[T(t +t ) T(t )]
t
=

x2

[T(x +x ) 2T(x)+T(x x)]

(7)

To help to visualize an iteration scheme, a suitable grid is


shown in Fig. 2 to represent the physical situation. The temperatures at the unknown grid points
[the ( j+1)th time step] are calculated from known

N
T (t) =T + ~ [A cos (nt)+B sin (nt)]
(4)
n
m
n
s
n= 1
where T is the time-averaged value of the surface temm
perature T (t), A and B are Fourier coeYcients, n is a
s
n
n
harmonic number and is the angular frequency. The
boundary conditions are
T( 0, t)=T (t )
s
T(l, t ) =T
ref

(constant)

Proc Instn Mech Engrs Vol 215 Part D

Fig. 2

Finite diVerence solution scheme

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

D09600 IMechE 2001

UNSTEADY IN-CYLINDER HEAT TRANSFER IN A SPARK IGNITION ENGINE

temperatures at other grid points (the jth time step).


Rearranging equation (7) gives
T

i,j+ 1

t
=T +
2T +T
)
(T
i,j
x2 i+ 1,j
i,j
i 1,j

(8)

where for stability the Fourier number t/x2 has to


be in the range from 0 to 0.5. As the sample rate is xed
by the encoder and engine speed, this imposes limitations
on the spatial steps in the grid. The surface temperature
and reference temperature provide boundary conditions,
but it is still necessary to make an initial guess for the
temperature distribution that is subsequently corrected.
Each cycle of data has to be taken in turn, and for the
rst cycle analysed a linear temperature variation was
assumed in the solid at time step 0. In general a diVerent
temperature prole will be predicted 720 later, and this
prole is then used as the temperature distribution at
time step 0. The solution is then repeated for the same
cycle, and the iterations within a cycle of data continue
until the changes in the temperature distribution are
negligible.
The numerical procedure was checked by using a
cosinusoidal surface temperature variation and comparing the numerical results with the exact solution derived
from the analytical solution, equation (2). For the worst
case (assuming an engine speed of only 120 r/min) the
error in the instantaneous surface heat ux was always
less than 1 per cent.
When combustion is not present, and the engine is
motored, the Reynolds and Prandtl numbers are largely
unchanged from ring operation, and thus the Nusselt
numbers for motoring and ring will be nearly the same,
so motoring is a useful operation when investigating heat
transfer. The only disadvantage of a motored engine is
that the surface temperature variation is smaller, and
hence there is likely to be a larger relative error in the
measurements. On the other hand, radiation can be

Fig. 3
D09600 IMechE 2001

751

ignored, soot deposits on the thermocouples are avoided,


gas pressure and temperature are well dened and are
almost symmetrical with respect to top dead centre
( TDC ) and the signals (although small ) show little cycleby-cycle variation. In Fig. 3 the surface heat ux results
for various locations in the cylinder can be seen for
motoring operation at 1000 r/min.
Figure 3 shows that the peak heat ux occurs close to
TDC. Lawton [6 ] measured the peak heat ux at about
8 before TDC (BTDC ), and he also showed the reversal
of the heat ux during the expansion stroke. The bulk
gas temperature during the period of negative heat ux
is higher than the wall temperature, and the reversal of
the heat ux can be explained by compression work in
the boundary layer as a result of the pressure change.
These unsteady eVects will be shown to cause a phase
advance in the heat ux. This phase shift will then
explain the diVerence in timing of the peak heat ux
measured by Lawton [6 ], who used a diesel engine, compared with this work. Diesel engines have a much higher
pressure variation than spark ignition engines and therefore the pressure work term eVects are more important.
Figure 4 shows the eVect engine speed has on the heat
ux for thermocouple 1, which is at the apex of the pent
roof and 31 mm away from the spark plug. The peak
heat ux rises as expected with increasing engine speed,
owing to the higher gas velocities and turbulence.
However, during the expansion stroke the negative heat
ux also becomes more important, and this is a result
of the unsteady processes in the boundary layer.
According to Gilaber and Pinchon [7] the heat ux
varies almost linearly with the density and is heavily
inuenced by the volumetric eYciency and thus by the
throttle position. This is shown in Fig. 5.
When the engine is red, the signal levels rise by
about an order of magnitude, and the signalnoise
ratio increases dramatically. Because of combustion, the

Surface heat ux dependence on location in the cylinder when motoring ( WOT, wide-open throttle)
Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

Proc Instn Mech Engrs Vol 215 Part D

752

D J OUDE NIJEWEME, J B W KOK, C R STONE AND L WYSZYNSKI

Fig. 4

Fig. 5

Surface heat ux dependence on speed when motoring

Surface heat ux dependence on the throttle position for thermocouple 1, in the apex of the pent
roof, 31 mm away from the spark plug

timing, magnitude and form of the surface heat ux will


mainly be determined by the ame front. Bigger cycleby-cycle variations are therefore expected, as a result of
locally irregular ame shapes. Figure 6 (compared with
Fig. 3) shows that the location of the thermocouples in
the cylinder becomes more signicant when the engine
is red. The diVerence between thermocouples 2 and 9
is especially striking. Both thermocouples are mounted
in the cylinder liner 15 mm below the top of the cylinder
but diametrically opposite. This can be explained by
local diVerences in turbulence, gas velocity and gas temperature levels. Thermocouple 12 is situated in the apex
of the pent roof 31 mm away from the spark plug, like
thermocouple 1. Thermocouple 3 is situated between the
Proc Instn Mech Engrs Vol 215 Part D

exhaust valves on the pent roof but close to the squish


region; the gas velocities are likely to be high here, owing
to the initial tumble motion. Because of the higher gas
velocities, there will be a thinner boundary layer in this
area, resulting in a higher surface heat ux, as also found
by Alkidas [8].
The occurrence of the negative heat ux in the expansion stroke of Fig. 6 is signicant, and it is believed that
this is the rst time this has been reported in a ring engine.
This phenomenon has already been observed during
motoring and can also be explained by the unsteady
eVects. Because of the higher pressure variations in the
red case, the unsteady eVects are expected to become
more important. With the much higher burnt gas tempera-

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

D09600 IMechE 2001

UNSTEADY IN-CYLINDER HEAT TRANSFER IN A SPARK IGNITION ENGINE

Fig. 6

Surface heat ux dependence on location when ring

tures, and the strong dependence of the heat ux on the


ignition timing (the earlier the ignition timing, the earlier
the peak heat ux and therefore a quicker decrease of the
heat ux in the expansion stroke and therefore a greater
chance of negative heat ux) and the possible later occurrence of the heat ux, the negative heat ux is typically
observed only at some locations in the cylinder.
Figure 7 shows the heat ux measured at two diVerent
locations in the cylinder and their distances to the spark
plug. As can be seen in this gure, the ame front has
a signicant eVect on the heat ux. The gure clearly
shows the ignition at 30 BTDC, followed by an initial
rise in heat ux, due to compression, followed by a steep

Fig. 7
D09600 IMechE 2001

753

rise of the surface heat ux after the ame front arrival
at the location in the cylinder concerned. Similar observations are described by Alkidas [5]. Figure 8 shows that
the magnitude of the heat ux increases when the
ignition timing is advanced and that its maximum occurs
earlier. The magnitude of the heat ux increases both
because the combustion temperatures will be higher and
because of the higher cylinder pressure and consequently
higher gas density. These eVects have also been shown
by Gilaber and Pinchon [7].
Apart from the heat ux reversal towards the end of
expansion, these results are in broad agreement with previously published work. This phenomenon will be dis-

Surface heat ux dependence on location in the cylinder when ring


Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

Proc Instn Mech Engrs Vol 215 Part D

754

D J OUDE NIJEWEME, J B W KOK, C R STONE AND L WYSZYNSKI

Fig. 8

Surface heat ux dependence on ignition timing for thermocouple 12, in the apex of the pent roof,
31 mm away from the spark plug

cussed in Section 4, after a discussion of CFD-based


predictions of heat transfer.

COMPUTATIONAL FLUID DYNAMICS


MODELLING

3.1 Results with a commercial computational uid


dynamics code
In Section 2 the instantaneous wall heat ux in a singlecylinder version of a Rover K16 engine was measured.
For engine design variation and optimization it is interesting if the heat ux can be predicted with the use of a
CFD code. This can save time and costs in engine test
programmes. In order to investigate the accuracy of the
heat ux predicted the Rover K16 engine was simulated
with a CFD code set up specically for a reciprocating
engine. Engine geometry les provided by Rover were
used to dene the computational domain, determined by
the cylinder, piston, engine head, inlet and exhaust ports
and valves. The CFD package was selected for its ability
to adapt the mesh according to piston and valve motion.
In order to avoid large mesh distortions, a new mesh is
automatically generated during the motion of the valves
and piston when a predened mesh shape limit is
exceeded. For the simulation of 490 crank shaft rotation
between inlet valve opening and exhaust valve opening
the computational domain was remeshed 160 times. The
mesh had about 220 000 cells at BDC.
At a given engine geometry, the intake ow determines
the ow eld in the engine during the compression and
expansion strokes [8]. For that reason the inlet port was
included in the computational domain and the ow at
entry to that port was calculated with the one-dimenProc Instn Mech Engrs Vol 215 Part D

sional unsteady-ow code PROMO. PROMO provided


the instantaneous values of mass ow, temperature and
pressure from which the instantaneous volume owrate
was calculated. Using the area at the inlet portmanifold
boundary enabled velocity to be estimated by assuming
plug ow, as an input condition to the three-dimensional
CFD code. Figure 9 illustrates the mass ow into the
inlet port at two operating conditions. The remaining
initial boundary conditions to be specied at the inlet
port are the fuel mass fraction and the inlet turbulence
characteristics. The latter are determined in the k turbulence model, by a turbulence length scale and the turbulence intensity as a fraction of the inlet velocity. These
two parameters were estimated here to be 0.50 mm and
10 per cent respectively. Kang et al. [9] claim that these
initial conditions have little eVect on the heat ux. This
statement can only be correct of course if the employed
k turbulence model is suYciently accurate to predict
the heat ux. Satisfactory comparisons of the ows were
made with data for the same engine in Jones and Junday
[10] [in which there are CFD predictions and laser
Doppler anemometry and hot wire anemometry
( HWA) data].
Figure 10 shows a comparison between the experimental data and the instantaneous heat ux averaged over
the cylinder head predicted by the CFD code for a
motored case. The data concern motored conditions at
2000 r/min. It can be observed that CFD code underpredicts the positive heat ux at TDC by a factor of about
10. The smaller (but interesting ) negative heat ux
between TDC and exhaust valve open ( EVO) is missed
completely by the CFD. Predictions (not shown here)
from the Woschni correlation [11], which is known to
underpredict the heat transfer in this type of combustion
system, are similar to the CFD results. Similar results

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

D09600 IMechE 2001

UNSTEADY IN-CYLINDER HEAT TRANSFER IN A SPARK IGNITION ENGINE

Fig. 9

Fig. 10

Inlet mass ow (0 is TDC during combustion)

Comparison between the measured and CFD-calculated surface heat ux for the cylinder head

with CFD are also shown by Han and Reitz [12] and
Diana et al. [13]. The reason for the CFD underprediction is the simplied representation of the boundary
layer processes in the models used in proprietary CFD
codes. It is expected that the underprediction will be even
larger in the red case owing to uncertainties in the CFD
combustion modelling. In the next section these simplied wall models used in the CFD code are analysed

3.2 Boundary layer modelling in computational uid


dynamics
In the turbulence model the direct eVects of molecular
viscosity on the energy containing ( large) scales of the
uctuating motion and those of the mean strain eld on
the corresponding dissipative (small ) scales are assumed
to be negligible. These assumptions are, in general, valid
D09600 IMechE 2001

755

in regions where the ow is fully turbulent. In the nearwall region, as the distinction between the large and
small scales of turbulent motion begins to diminish, the
above assumption breaks down. The simplest and still
the most popular approach to near-wall modelling has
been to employ the logarithmic law of the wall, also
known as the wall function approach [14]. CFD packages use the wall function method to interpolate between
the fully turbulent region and the wall.
The hydrodynamic boundary layer is split into a laminar viscous sublayer and a buVer layer to adapt the laminar part of the boundary layer to the turbulent core.
This way of representing the hydrodynamic boundary
layer, together with an extensive experimental investigation by Nikuradse, led to the law of the wall.
Although this law was set up just for steady, incompressible ow through smooth pipes at moderate
Reynolds numbers, it seems to hold for a wider range

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

Proc Instn Mech Engrs Vol 215 Part D

756

D J OUDE NIJEWEME, J B W KOK, C R STONE AND L WYSZYNSKI

of applications and is often used outside pipes as well.


The use of wall functions removes the need to employ a
ne mesh to resolve explicitly the ow prole in the
boundary layer, and it is therefore computationally
eYcient. However, the modelling of the boundary layer
using wall functions is based on a set of assumptions
about the ow, and the accuracy of the solution obtained
is therefore dependent on how well these assumptions
are met in any particular case. The main assumptions
implicit in the wall-function treatment are as follows:
(a) steady ow;
(b) incompressible (i.e. no change in density);
(c) the ow is essentially one dimensional, such that
gradients of velocity and scalar quantities are only
normal to the wall;
(d ) the eVects of pressure gradients are small;
(e) the turbulence is in local equilibrium;
(f ) the turbulence length scale varies linearly with
distance from the wall.
In addition to these restrictions, it has been found by
Gibson [15] that even slightly curved surfaces cause
inaccurate heat ux predictions.
The wall function formulation is

y+ ,

u+ = 1
ln(Ey+ ),
k

y+ y+
v
y+ >y+
v

(9)

where
u+ =u/u
u = tangential uid velocity
u = ( /r)1/2
w
= wall shear stress
w
y+ =rC 1/4 k1/2y/
y+ = 11.6, the value of y+ at the edge of the viscous
v
sublayer
y = distance of the cell centre from the wall
r = gas density
= laminar viscosity of the uid
k = turbulence kinetic energy
and k, E and C are empirical coeYcients in the k
model with values as follows:
k = 0.419
E = 9.793
For local equilibrium, where k = ( /r)C 1/2, the wall
w
shear stress can therefore be expressed as
rC 1/4 k1/2ku
ln (Ey+ )

isothermal boundary layer ow;


no chemical heat release in the boundary layer;
no pressure work in the boundary layer;
no temperature gradient along the wall;
the turbulent Prandtl number Pr is assumed to be
t
constant.

Of particular signicance is the no pressure work in the


boundary layer assumption, since it is this term that
leads to the reversal of the heat ux. In the CFD code,
the temperature variation equation is modied to take
account of the Prandtl number and the so-called
sublayer resistance factor, P:
T+ =Pr (u+ +P )
t
where

( 11)

T+ =c r(T T )u /q
p
w
c = isobaric specic heat capacity of the uid
p
T = uid temperature
T = wall temperature
w
q=heat ux
Pr = turbulent Prandtl number, whose value is
t
taken as 0.9
P=9.0(Pr /Pr 1) (Pr /Pr ) 1/4
l t
l t
Pr = laminar Prandtl number
l
The heat ux can then be written as
c y+ (T T )
p
w
( 12)
Pr [(1/k) ln (Ey+ ) +P ]y
t
This expression for heat ux appears as a source term
in the enthalpy equation for near-wall cells.
A local heat transfer coeYcient, h, is calculated and
written to the post-processing le according to

(10)
h=

This expression for shear stress appears as a source term


in the momentum equations for near-wall cells.
In addition to assumptions for modelling the boundProc Instn Mech Engrs Vol 215 Part D

(a)
(b )
(c )
(d )
(e)

q=

C = 0.09

=
w

ary layer ow, there are also assumptions in the determination of the heat ux. As the equations for heat transfer
and momentum transfer in the boundary layer are analogous, and their boundary and initial conditions are
similar, the solutions must be analogous. This is only
strictly valid under some conditions, of which no pressure gradient is the most important one. In the viscous
sublayer, the uid ow is laminar and heat transfer is
mainly dominated by heat conduction. In the turbulent
core the turbulence is the driving force behind the
heat transfer. The buVer layer provides a transition,
by accommodating both of these mechanisms. The
Reynolds analogy formulation is the simplest way to link
the momentumheat transfer and is used to obtain
the temperature variation in the boundary layer. The
Reynolds analogy introduces assumptions that will
impose some more conditions on the boundary layer,
including:

q
T T

( 13)
w
It is stressed that this is a local heat transfer coeYcient,
i.e. it is based on a local uid temperature for each sur-

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

D09600 IMechE 2001

UNSTEADY IN-CYLINDER HEAT TRANSFER IN A SPARK IGNITION ENGINE

face patch. For a given heat ux, the uid temperature


in a small cell will be closer to the wall temperature than
that in a large cell. As explained earlier in this section,
the wall function treatment is not exact and is based on
a number of assumptions about the boundary layer. The
formulae above are generally held to be applicable in
the approximate range 30y+ 300, in which the calculated values may be expected to be independent of y+ .
However, in practice this is not necessarily the case, and
even within the above range of y+ variations of surface
quantities may be observed.
In the CFD mesh, the cell size (and hence y+) near
the walls is subject to variations. This is a consequence
of the mesh structure, which is in turn dictated by the
automatic mesh generation method used in the CFD
package. The result is that surface quantities such as
shear stress and heat ux may display stripes or a patchiness, which reects the underlying variation in near-wall
cell size. This situation is compounded for heat transfer
coeYcients which, as described above, are based on local
cell uid temperatures; therefore the stripes in heat
transfer coeYcient are more pronounced than those for
shear stress or heat ux.
The law of the wall and the Reynolds analogy give in
a steady isobaric ow a prediction of heat transfer with
modest accuracy. When the pressure is changing rapidly,
as in a piston engine, heat transfer processes become
important but are not taken into account by the law of
wall and Reynolds analogy. This is clearly illustrated by
the measurements presented in Section 2. The measured
reversal of the heat ux is evidence of the invalidation
of the Reynolds analogy. This means that the heat transfer modelling requires more processes to be taken into
account, with greater modelling resolution in the boundary layer. Several approaches can be followed. A threedimensional CFD method along the lines of Section 3
with a so-called k turbulence model can be used [16 ].
This is at present being implemented in a CFD code at
the University of Twente. A diVerent approach is presented in Section 4 of this paper. This method uses a
one-dimensional energy equation for the boundary layer.
This equation retains all the relevant terms and is solved
in the boundary layer by a numerical method.
4

ENERGY EQUATION MODELLING IN THE


BOUNDARY LAYER

Lawton [6 ] modelled the energy equation in the thermal


boundary layer of a motored diesel engine, and despite a number of assumptions provided evidence for
the reversal of the heat ux. Using a one-dimensional
approach (since properties can be assumed to vary only
in a direction normal to the wall ), requires solution of
rc

qT
qT
+rvc
=
p qy
p qt
convection

D09600 IMechE 2001

qq
y
qy
heat flux

dp
dt
work

(14)

757

In prior work, Diana et al. [13] ignored the work term,


while Han and Reitz [12] and Angelberger et al.[17] also
ignored the convective term (which accounts for ow
normal to the wall as a result of density changes). It will
be seen in the analysis to be presented here that both
these terms are important.
The heat ux term can be determined by using the
concept of a turbulent conductivity (k ) to supplement
t
the laminar conductivity (k )
l
qT
q= (k +k )
( 15)
l
t qy
Thus the energy equation ( 14) becomes
rc

qT
qT
qT
q
dp
+rvc
=
+
(k +k )
t qy
p qt
p qy qy
dt

(16)

The solution of equation (16) is facilitated by conversion


to Lagrangian coordinates and normalization of the temperatures. The one-dimensional energy equation for a
pure ideal gas can be expressed in Lagrangian coordinates as

C A B D

k qT
r q r
qT
dp
=
+
k 1+ l
(17)
p qt
r qz r
k qz
dt
0
0
The relation between the Lagrangian, z, and Eulerian,
y, coordinates is
rc

y r
dy
( 18)
r
0 0
which satises the continuity equation. The subscript, 0,
means the evaluation of the property at initial conditions
such as IVC. The Lagrangian coordinate is now attached
to each uid particle, which is why the second term on
the left-hand side is made redundant.
Since for a semi-perfect gas r=P/RT, R =c c and
p
v
=c /c , equation (17) becomes
p v
1 T dp
k qT
kr
r q
qT
=
+
k 1+ t
p dt
r c qz k r
k qz
qt
0 p
0 0
( 19)
r dz =r dy
0

or

z=

A B D

Far away from the wall, the gas is assumed to be compressed isentropically
T /( 1)
p
= 2
(20)
p
T
0
0
Assuming that the gas conductivity is proportional to
absolute temperature, following Isshiki and Nishiwaki
[18], gives

A B

kr
p
=
( 21)
k r
p
0 0
0
The relation between the thermal conductivity, k, and
the gas viscosity, , can be described by
k
Pr
t=
t
k Pr
t

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

(22)
Proc Instn Mech Engrs Vol 215 Part D

758

D J OUDE NIJEWEME, J B W KOK, C R STONE AND L WYSZYNSKI

Finally, on following Isshiki and Nishiwaki [18], normalizing in terms of T(?, t) with
U(z, t ) =

T(z, t)
T(?, t)

(23)

equation (19) becomes


p q2U
p q Pr qU
qU
t
=
+
0 p qz2
0 p qz Pr qz
qt
0
0
t

(24)

laminar
turbulent
where =k /r c . The boundary conditions and initial
0
0 0 p
condition for equation (24) are
U(0, t) =

T(0, t)
= f (t )
T(?, t)

U(?, t)=1
U(z, 0 ) =

T(z, 0)
= f (0)g(z)
T(?, 0)
(25)

where g(z) is a function of z, corresponding to the initial


boundary layer existing at the beginning of the compression process. Equation (24) has been divided into a
laminar and a turbulent contribution to the heat transfer. The laminar part is equivalent to the conduction
equation that has previously been solved for the
unsteady heat ow in the solid, for obtaining the surface
heat ux from the measured wall temperature.
As a reminder, the following assumptions have been
made:
1. Pressure is a function only of time: p =p (t).
2. Radiative heat transfer and heat release in the boundary layer are neglected.
3. The temperature and velocity gradients in the direction parallel to the wall are neglected.
4. Gas transport properties are proportional to the
absolute gas temperature.
5. The gas is semi-perfect.
To be able to solve this equation, the turbulent thermal
conductivity, k , or the turbulent Prandtl number, Pr ,
t
t
and the turbulent viscosity, , need to be known across
t
the boundary layer. Usually empirical expressions are
used, obtained from measurement in incompressible
turbulent ows, like those of Mellor [19] for / and
t
Kays [20] for Pr/Pr . Because the compressibility of the
t
boundary layer is recognized as an important parameter
when determining the surface heat ux, this seems
contradictory.
Thus a diVerent method has been adopted here. The
turbulent thermal conductivity, k , is tuned until equat
tion (24) gives a temperature eld close to the wall, such
that the heat ux matches the measured surface heat
ux for that location in the cylinder. Turbulence intensities can be calculated with CFD and so the thermal
conductivity expression can be tuned by location,
Proc Instn Mech Engrs Vol 215 Part D

greatly improving the spatial diVerences in heat ux.


Temperature measurements close to the wall have been
obtained by Lucht et al. [21], using a CARS (coherent
anti-Stokes Raman spectroscopy) temperature measurement system. Although the engine in that study was run
under diVerent conditions from those for which the
temperature eld has been calculated, a comparison
( Fig. 11) none the less shows good general agreement
for the shape of the temperature prole at TDC in the
boundary layer. It is unfortunate that the CARS temperature measurements were only in the region of TDC
and did not continue far enough into the expansion
stroke to show a reversal in the temperature gradient
adjacent to the wall. As well as providing evidence of
the negative heat ux, the solution of equation (14) can
illustrate the relative contributions of the convective and
work terms, and this is shown in Fig. 12.
Figure 12 clearly shows the eVects of the individual
terms of the energy equation. The convection term is the
most important for the magnitude of the heat transfer.
The pressure work term is very important for the phasing
of the heat transfer. Even with a low pressure variation
of about 5 bar, as used here, the heat ux is advanced
by 10 CA in the cycle, compared with the heat ux
without the work term.
In conclusion, it can be stated that, if the surface heat
transfer is modelled, both the work term and the convective term need to be taken into account. This is because
the convective term largely determines the magnitude of
the heat ux, and the work term is responsible for the
phase advance of the heat ux. This phase shift enhances
the peak heat ux and is also partly responsible for the
negative heat ux. To be able to model the boundary
layer, for both modied wall functions and the direct
solution of the one-dimensional energy equation, more
information is needed about the turbulent Prandtl
number and the turbulent gas viscosity in the boundary
layer, for the compressible ow eld, of an internal combustion engine. If these turbulent boundary layer parameters are modelled more accurately, the energy equation
can then predict the surface heat ux more accurately.
This is especially true if modelling the energy equation
in the boundary layer were to be implemented into a
CFD code, using the main ow features, such as gas
temperature, turbulence intensity, gas density and pressure data, to calculate the local instantaneous surface
heat ux.
5

CONCLUSIONS

Instantaneous heat ux measurements from fast


response thermocouples have shown that in the expansion stroke, heat can ow from the wall into the combustion chamber, even though the bulk gas temperature is
higher than the wall temperature. It is believed that this
is the rst time such results have been reported from a

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

D09600 IMechE 2001

UNSTEADY IN-CYLINDER HEAT TRANSFER IN A SPARK IGNITION ENGINE

Fig. 11

Temperature prole comparison with CARS measurements by Lucht et al. [21] at TDC

Fig. 12

The relative importance of the individual terms of the energy equation

ring engine. This unexpected result has been explained


by modelling the unsteady ows and heat conduction
within the gas side thermal boundary layer. This modelling has shown that these unsteady eVects change the
phasing of the heat ux, compared with that which
would be predicted by a simple convective correlation
based on the bulk gas properties. In other words, it is
impossible for the widely used correlations (of the form
Nu=ReaPrb ) to give the correct phasing when the bulk
gas temperature is being used. It has thus been shown
that the CFD heat ux calculations need to include the
work term within the boundary layer and the convection
term if accurate results are to be obtained.
The heat ux probes comprised a pair of thermocouples inside a 5 mm diameter probe in a tapered housing, one thermocouple being formed at the surface and
D09600 IMechE 2001

759

the other (the reference junction) being 5 mm away. The


temperature measurements were used as boundary conditions for solving the unsteady heat conduction equations; a nite diVerence approach was used in contrast
to the more usual Fourier analysis. The heat ux probes
have been installed throughout the combustion chamber
of a pent roof, four-valve, single-cylinder spark ignition
engine. The surface heat uxes have been reported here,
showing the eVects of location, speed and throttle setting
for both motoring and ring, and the eVect of ignition
timing when ring.
To make comparisons with these measurements, the
combustion system was modelled with CFD. Engine
geometry les were used to dene the combustion
chamber, the inlet port and the piston and valve motion.
Because the intake ow is highly important to the ow

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

Proc Instn Mech Engrs Vol 215 Part D

760

D J OUDE NIJEWEME, J B W KOK, C R STONE AND L WYSZYNSKI

in an IC engine, the inlet port was modelled as well as


the combustion chamber. The ow at entry to the inlet
port was modelled with a one-dimensional unsteady ow
package, and the calculated values of mass ow, temperature and pressure were then used as inlet boundary
conditions to the three-dimensional CFD. The CFD predictions were found to give very poor agreement with
the experimental measurements, so this led to a review
of the assumptions used in boundary layer modelling
within CFD. The discrepancies were attributed to
assumptions in the law of the wall and the Reynolds
analogy, so instead the energy equation was solved for
the boundary layer.
The one-dimensional energy conservation equation
has been linearized and normalized and solved in the gas
side boundary layer for a motored case. As well as heat
conduction, the energy equation incorporates work
being done within the boundary layer (as a result of the
pressure changes) and a convective ow normal to the
surface (due to density changes arising from both pressure and temperature changes in the boundary layer). The
results have been used for a parametric study, in which
the individual terms of the energy equation were evaluated for their contribution to the surface heat ux. It
was clearly shown that the work term causes a phase
shift of the heat ux forward in time and that the convective ow term contributes signicantly to the magnitude
of the heat ux.

ACKNOWLEDGEMENTS
Support from the EPSRC, BMW/Rover Group and
Shell Global Solutions is gratefully acknowledged.

REFERENCES
1 Myers, J. P. and Alkidas, A. C. EVects of combustion
chamber surface temperature on the exhaust emissions
of a single-cylinder spark-ignition engine. SAE paper
780642, 1978.
2 Ball, J., Raine, R. and Stone, R. Combustion analysis and
cycle-by-cycle variations in spark ignition engine combustion. Part 2: a new parameter for completeness of combustion and its use in modelling cycle-by-cycle variations in
combustion. Proc. Instn Mech. Engrs, Part D, Journal of
Automobile Engineering, 1998, 212(D6), 507524.
3 Gatowski, J. A., Smith, M. K. and Alkidas, A. C. An experimental investigation of surface thermometry and heat ux.
Expl Thermal Fluid Sci., 1989, 2, 280292.
4 Eckert, E. R. G. and Drake, R. M. Heat and Mass Transfer,
2nd edition, 1959 (McGraw-Hill, New York).

Proc Instn Mech Engrs Vol 215 Part D

5 Alkidas, A. C. Heat transfer characteristics of a sparkignition engine. J. Heat Transfer, 1980, 102(2), 189193.
6 Lawton, B. EVect of compression and expansion on instantaneous heat transfer in reciprocating internal combustion
engines. Proc. Instn Mech. Engrs, Part A, Journal of Power
and Energy, 1987, 201(A3).
7 Gilaber, P. and Pinchon, P. Measurements and multidimensional modelling of gaswall heat transfer in a S.I. engine.
SAE paper 880516, 1988.
8 Alkidas, A. C. The eVects of intake-ow congurations on
the heat-release and heat-transfer of a single cylinder fourvalve S.I. engine. SAE paper 910296, 1991.
9 Kang, Y. H., Chang, I.-P. and Martin, J. K. A comparison
of boundary layer treatments for heat transfer in IC
engines. SAE paper 90052, 1990.
10 Jones, P. and Junday, J. Full cycle computational uid
dynamics calculations in a motored four valve pent roof
combustion chamber and comparison with measurement.
SAE paper 950286, 1995.
11 Woschni, G. A universally applicable equation for the
instantaneous heat transfer coeYcient in the internal combustion engine. SAE Trans., 1967, 76, 30653083.
12 Han, Z. and Reitz, R. D. A temperature wall function formulation for variable-density turbulent ows with application to engine convective heat transfer modeling. Int. J.
Heat Mass Transfer, 1997, 40, 613625.
13 Diana, S., Giglio, V., Police, G., Bella, G. and Cordiner, S.
Heat transfer evaluation in 3D computations of premixed
SI engines. SAE paper 972876, 1997; Diesel and SI engine
modelling, SP-1306, pp. 5770.
14 Patankar, S. V. and Spalding, D. B. A calculation procedure for heat, mass and momentum transfer in threedimensional parabolic ows. Int. J. Heat Mass Transfer,
1972, 15.
15 Gibson, M. M. EVects of the surface curvature on the law
of the wall. Near-Wall Turbulence, Zoran Zaric Memorial
Conference, 1988, pp. 157171 (Hemisphere, New York).
16 Wilcox, D. C. Simulation of transition with a two-equation
turbulence model. Am. Inst. Aeronaut. Astronaut. J., 1994,
31(8), 14141421.
17 Angelberger, C., Poinsot, T. and Delhay, B. Improving nearwall combustion and wall heat transfer modelling in SI
engine computations. SAE paper 972881, 1997; Diesel and
SI engine modelling, SP-1306, pp. 113130.
18 Isshiki, N. and Nishiwaki, N. Study on laminar heat transfer of inside gas with cyclic pressure change on an inner
wall of a cylinder head. In Proceedings of the Fourth
International Heat Transfer Conference, Versailles, Paris,
1970, paper FC3.5, pp. 110.
19 Mellor, G. L. In Proceedings of Symposium on Fluidics
Internal Flow, Pennsylvania State University, 1968.
20 Kays, W. M. Turbulent Prandtl numberwhere are we?
Trans. ASME, J. Heat Transfer, 1994, 116, 284.
21 Lucht, R. P., Dunn-Rankin, D., Walter, T., Dreier, T. and
Bopp, S. C. Heat transfer in engines: comparison of CARS
thermal boundary layer measurements and heat ux
measurements. SAE paper 910722, 1991.

Downloaded from pid.sagepub.com at National Cheng Kung University on August 20, 2014

D09600 IMechE 2001

You might also like