Sunteți pe pagina 1din 446

A new operating range enhancement

device combined with a casing treatment


and inlet guide vanes for centrifugal
compressors
I Tomita, B An, T Nanbu
Mitsubishi Heavy Industries, Ltd, Japan

ABSTRACT

This paper describes a new device combined with a casing treatment and inlet
guide vanes that achieves both high pressure ratio and wide operating range of
centrifugal compressors. The fixed inlet guide vanes equipped at the downstream of
the recirculation bypass slit, induce a counter-rotating inflow to the impeller.
Numerical calculations including whole impeller, vaneless diffuser, scroll, casing
treatment and inlet guide vanes were conducted to analyze the detailed internal
flow field. The calculation results indicated that counter-rotating inflow induced by
the inlet guide vanes increased pressure ratio at high flow rate. On the other hand,
it could also enhance the stabilizing effects of the casing treatment at low flow rate.
Pressure measurements of impeller inlet and were conducted using high response
pressure transducers and found that unstable phenomena like rotating stall were
inhibited by the new device.

Finally, the map width enhancement effects of the new device were proved by
performance tests and over 70% wider map width was obtained where pressure
ratio was about 2.8. The newly developed map width enhancement device
combined with a casing treatment and fixed inlet guide vanes provides a wider
operating range and higher pressure ratio, without any extra complex variable
device.

1 INTRODUCTION

Recently, demands for turbochargers are growing year by year in the background of
tightening of exhaust gas and fuel consumption regulations. Especially, operating
range enhancement of centrifugal compressors is difficult because of trade-off
between pressure ratio and surging characteristics. However many points of
surging phenomena have not been cleared yet, previous studies showed that flow
separations at shroud side of impeller and unstable phenomena like rotating stall
should be controlled for surging characteristics improvement(1)(2).

Two compressors with different operating range width were investigated with
experimental and computational flow analysis. Based on the studies, we had found
one of the operating range enhancement methods utilizing a blockage effect
induced by tip leakage vortex breakdown(3)(4). Furthermore, we had achieved
developments of high efficiency compressors and wide range compressors by its
techniques(5). Only impeller developments without any additional components
were conducted in these studies. Other techniques with small trade-off between
surging characteristics and pressure ratio are needed for further operating range
enhancements.

_______________________________________
© The author(s) and/or their employer(s), 2014
79
In this study, we developed and validated operating range enhancement effects of
a casing treatment (CT) and inlet guide vanes (IGV) combination. A recirculation
type CT is one of the operating range enhancement method by eliminating
separation at low flow rate. However a variable IGV could control compressor
performances by inducing rotational inflow at compressor inlet, its complexity of
mechanical structure is critical problem.

2 TEST COMPRESSORS

The test compressor is for automotive turbocharger. The compressor has a


backswept impeller with 6 full blades and 6 splitter blades, vaneless diffuser and
scroll. The outlet diameter of the impeller is 51mm. Figure 1 shows performance
maps of a normal compressor without any devices and with a recirculation type CT.
Flow rate is normalized by choking flow rate at 178,000rpm of the normal
compressor. The efficiency is also normalized by maximum efficiency of the normal
compressor. The surging characteristics on the normal compressor are deteriorated
at pressure ratio (PR) in 2.3 or higher. Its tending is often seemed in high efficiency
or high pressure compressor especially. The recirculation type CT was equipped to
enhance the operating range by stabilizing the internal flow field with recirculation
flow at inducer at low flow rate condition. By installing CT, efficiency was about
3.8point lower than normal compressor, the pressure ratio at low flow rate rises
and was maintained good surging characteristics even 2.3 or higher pressure ratio.
There was little reduction in efficiency at low flow rate by its flow improvement
effect.

3.5 3.5

Deterioration of the Improvement of


operating range surging characteristics
3.0 3.0

2.5 2.5
Total Pressure Ratio πc

Total Pressure Ratio πc

0.99
2.0 0.98 2.0 0.95
0.97
0.95 0.93
0.93
0.90 0.90
0.85 0.85
178000rpm 178000rpm
0.80 0.80
0.75 157000rpm
1.5 0.70 157000rpm 1.5
0.75
0.70
133000rpm
133000rpm
101000rpm
64000rpm at) 20℃ 101000rpm at) 20℃
1.0 64000rpm
1.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Volume Flow Ratio Q0/Qmax[-]
Volume Flow Ratio Q0/Qmax[-]

Normal compressor CT
Figure 1: Compressor performance map (measurement)

Figure 2 shows a cutaway view of the compressor with the new device which has
recirculation type CT and IGV. The number of the IGV and its angle inducing
counter-rotating inflow to the impeller was adjusted by in-house test survey.

80
Large number of IGV increases its operating range enhancement effect, but the
efficiency is deteriorated due to friction loss increase. 7 vanes was the best number
for efficiency and surging characteristics in the case of this compressor. And an
inner ring was installed to eliminate its tip leakage vortex and decrease its
vibration. This newly developed device provides a wider operating range and higher
pressure ratio, without any extra complex variable device. In this study, flow
mechanisms on the normal compressor, CT and the new device (IGV+CT) were
investigated.

Scroll
Casing treatment

Recirculation flow

Inlet guide vanes

Flow

Rotation Impeller
Flow

Inner ring
Vaneless diffuser
Figure 2: The new map width enhancement device

3 NUMERICAL AND EXPERIMENTAL METHOD

In this study, we analyzed the flow stabilization mechanism in steady CFD. And the
unsteady phenomena was confirmed by a pressure fluctuation measurement.

3.1 CFD scheme


CFX version 14.5 was used for numerical calculation. Order to flow improvement by
recirculation flow at the casing treatment, evaluation of the recirculation flow rate is
important. The recirculation flow rate is determined by the pressure balance of the
impeller and the casing treatment. Since distortion of the flow is generated by the
circumferential pressure distributions in the scroll within the low flow rate operation
in particular, numerical calculations were performed including whole impeller, IGV,
casing treatment, vaneless diffuser and scroll applying the Frozen-Rotor in
boundary conditions.

The SST turbulence model was adopted. And the numbers of cells were about
2,440,000 for impeller, 920,000 for diffuser, 1340,000 for scroll and 7,300,000 for
other elements.

3.2 Measurements scheme


Pressure fluctuation measurement was conducted to detect the unsteady
phenomena on these compressors. As shown in Figure 3, measurement locations
are 3mm upstream from impeller leading edge. The purpose of the measurement
near impeller leading edge is to detect blade passing waves, rotating stall, and

81
rotating instability, etc. Two sensors are installed in the same meridional location to
detect propagating phenomena in circumferential direction. After passing through
the sensor 1, the impeller passes through the sensor 2. If phenomena are
propagating, a phase difference of 30 degrees is detected by sensor 1 and sensor 2.
When the rotating stall are generated, the speed and the number of disturbances
are estimated from the phase lag of these two sensors. Endevco 8510C sensors
with sampling frequency of 250kHz were applied. It corresponds to 14 points for 1
pitch on full blades at 178,000rpm.

Impeller
Rotation

Sensor 2
Pressure
transducer 3mm

30 deg Sensor 1

Figure 3: Positions of pressure transducers

4 NUMERICAL CALCULATION RESULTS

Figure 4 shows relative mach number distributions at 90% span of the impeller at
178,000rpm. Figure 4(a) shows peak efficiency point (Q/Qchoke=0.836), figure 4(b)
shows peak pressure point of the normal compressor with neither CT nor IGV
(Q/Qchoke=0.716) and figure 4(c) shows near surging point of the normal
compressor (Q/Qchoke=0.597).

4.1 Peak efficiency point (Q/Qchoke=0.836, figure 4(a))


Shock waves from suction surface and tip leakage flow were seen in figure 4(a).
Mach number distributions show that flow separations did not occur and the
impeller still had loading on these compressors because velocity differences
between pressure side and suction side on the inducer.

Two changes in flow field were observed. One was circumferential uniformity of tip
leakage flow by CT, and the other was higher acceleration on suction surface
caused by counter-rotating inflow induced by IGV.

4.2 Peak pressure point of the normal compressor (Q/Qchoke=0.716,


figure 4(b))
On the normal compressor, a large stall region was observed and eliminating of the
blade loadings was cleared from that the velocity difference between pressure side
and suction side disappeared. On the other hand, any stall did not occur on CT and
IGV+CT. These results confirmed that CT and IGV+CT could stabilize the flow
structure by recirculating flow at a non-design point.

82
Full blade
Larger
Rotation incidence angle

TE Without
IGV
Relative
velocity
LE With IGV
Full blade
High acceleration
Shock wave Splitter blade

Acceleration Tip leakage flow Tip leakage flow


was uniformed
Velocity differences by CT.
Flow by blade loading

Normal compressor CT IGV + CT


(a) Peak efficiency point (Q/Qchoke=0.836)

No stall region No stall region


Stall region
No velocity
difference
-> lost loading

Flow Large velocity


differences
-> High-loading

Normal compressor CT IGV + CT


(b) Peak pressure point of the normal compressor (Q/Qchoke=0.716)

No stall region No stall region

Stall region

No stall region but


blade loadings are
Keep high
not uniform
blade loadings
on all blades

Normal compressor CT IGV + CT


(c) Surging point of the normal compressor (Q/Qchoke=0.597)
Figure 4: Relative mach number distribution (90% span, 178,000rpm)

83
4.3 Surging point of the normal compressor (Q/Qchoke=0.597, figure 4(c))
At the surging point of the normal compressor, a large stall region could be
eliminated by CT in the same way as Q/Qchoke=0.716. Even though blade loadings
were not uniform on each blade on CT, IGV+CT could keep large blade loadings at
all blades. IGV+CT had a stable flow structure same as the design point even at low
flow rate. The obtained results suggest the operating range enhancement effect of
the new device.

4.4 Recirculation flow ratio


Figure 5 shows the ratio of their recirculation flow rate ∆Q and total flow rate Q.
Since Q/Qchoke=0.597 was an unstable point inherently, convergence of the analysis
was not fine. The recirculation flow ratio ∆Q/Q was little changed by existence of
IGV at Q/Qchoke=0.597. The recirculation flow rate on IGV+CT was 12% and that on
CT was 9% at Q/Qchoke=0.716, so the recirculation flow rate increased 1.4 times
that of CT. The recirculation flow is generated by pressure difference on inlet and
outlet of CT. Its pressure difference was increased by counter-rotating flow by IGV
because the counter-rotating flow raised the blade loading. The results obtained by
numerical calculation suggest that IGV improves the flow stabilizing effect of CT
because of its counter-rotating flow.

0.3
CT
IGV + CT
0.2
Recirculation Flow Ratio ⊿Q/Q[-]

0.1

0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2

-0.1

-0.2
Volume Flow Ratio Q0/Qmax[-]
Figure 5: Recirculation flow rate at casing treatment (178,000rpm, CFD)

5 PRESSURE FLUCTUATION MEASUREMENT

Figure 6 shows measurement results of pressure fluctuations at inlet at


178,000rpm. As figure 4, figure 6(a) shows peak efficiency point (Q/Qchoke≒0.836),
figure 6(b) shows peak pressure point of the normal compressor with neither CT
nor IGV (Q/Qchoke≒0.716) and figure 6(c) shows near surging point of the normal
compressor (Q/Qchoke ≒0.597). The solid line is pressure fluctuation measured by
the sensor 1 and dotted line is that of the sensor 2. The range of horizontal axis
corresponds to about 1.2 rotations of the impeller. In the case of stall, the blade
loading disappeared and blade-passing fluctuations are not clear.

84
5.1 Peak efficiency point (Q/Qchoke≒0.836, figure 6(a))
Blade passing fluctuations were detected clearly and any low frequency fluctuations
were not seen at this flow rate. It seems that any unsteady phenomena did not
occur and their internal flow structures was steady on the normal compressor.
Pressure fluctuations on IGV+CT were slightly higher by counter-rotating inflow.

5.2 Peak pressure point of the normal compressor (Q/Qchoke≒0.716,


figure 6(b))
The pressure fluctuations clearly showed that unsteady flow disturbances
considered being a stall occurred on the normal compressor. On the other hand,
pressure fluctuations are stable on CT and IGV+CT. The effectiveness of CT and
IGV+CT was confirmed experimentally.

5.3 Surging point of the normal compressor (Q/Qchoke≒0.597, figure 6(c))


Both Stall and surging did not occur on CT and IGV+CT even at the flow rate that
surging occurred on the normal compressor. It was also observed that the blade
loadings were not uniform on each blade on CT. And IGV+CT could keep large
blade loadings at all blades. These measurement results supported the numerical
calculations.

Sensor 1 Sensor 2
1 pitch of full blades

Pressure waves by 2 sensors Higher fluctuations by IGV

Normal compressor CT IGV + CT


(Q/Qchoke=0.834) (Q/Qchoke=0.833) (Q/Qchoke=0.839)
(a) Peak efficiency point

Unsteady flow disturbances No disturbance No disturbance

Normal compressor CT IGV + CT


(Q/Qchoke=0.691) (Q/Qchoke=0.716) (Q/Qchoke=0.713)
(b) Peak pressure point of the normal compressor

Keep large blade loadings


Blade loadings are not uniform on all blades

No data because of the surging

No disturbance No disturbance

Normal compressor CT IGV + CT


(Q/Qchoke=0.598) (Q/Qchoke=0.594)
(c) Surging point of the normal compressor
Figure 6: Pressure fluctuation (178,000rpm, measurement)

85
6 PERFORMANCE IMPROVEMENT

The compressor performance of IGV+CT and comparison are shown in figure 7.


Surging characteristics of the normal compressor were deteriorated at pressure
ratio in 2.3 or higher. However the compressor with CT had 38% smaller surging
flow rate at 178,000rpm, its pressure ratio was dropped over 0.13 m3/s. This
pressure deterioration might be occurred by loss generation at the recirculation flow
passage. On the other hand, the compressor with IGV+CT had best surging
characteristics that surging flow rate decreased 20% at PR=2.3 and 50% at PR=2.8
compared with the normal compressor.

Moreover, IGV+CT achieved higher pressure ratio at whole flow rate by counter-
rotating flow. When operating range was defined as (Qchoke-Qsurge)/Qchoke, the
operating range of IGV+CT was enhanced 76% compared with the normal
compressor and enhanced 14% compared with CT at 178,000rpm. The maximum
efficiency was not deteriorated by IGV because the maximum efficiency of IGV+CT
was almost same as CT.

3.5
Normal Compressor
CT
IGV+CT IGV+CT

CT
3.0
Normal

2.5
Total Pressure Ratio πc

2.0
0.95
0.93
0.90 178000rpm
0.85
0.80
0.75
1.5 0.70 157000rpm

133000rpm

at) 20℃
64000rpm
1.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Volume Flow Ratio Q0/Qmax[-]

IGV + CT Performance comparison


Figure 7: Compressor performance map (measurement)

7 CONCLUSION

In this study, the effects of combination of a casing treatment and inlet guide vanes
were investigated by numerical calculations and measurements.

In conclusion, the results are summarized as follows.

(1) Counter-rotating flow generated by the inlet guide vanes makes pressure ratio
high at whole flow rate. Furthermore, flow unsteadiness could be controlled by
combination with casing treatment.

86
(2) The newly developed map width enhancement device combined with a casing
treatment and fixed inlet guide vanes provides over 70% wider operating
range and higher pressure ratio compared with the normal compressor where
pressure ratio is about 2.8, without any extra complex variable device.

REFERENCES

(1) Yamada, K., Furukawa, M., Fukushima, H., Ibaraki, S., Tomita, I., 2011, “The
Role of Tip Leakage Vortex Breakdown in Flow Fields and Aerodynamic
Characteristics of Transonic Centrifugal Compressor Impellers”, ASME Paper,
GT2011-46253.
(2) Iwakiri, K., Furukawa, M., Ibaraki, S., Tomita, I., 2009, “Unsteady and Three-
dimensional Flow Phenomena in a Transonic Centrifugal Compressor Impeller
at Rotating Stall”, ASME Paper, GT2009-59516.
(3) Tomita, I., Ibaraki, S., Furukawa, M., Yamada, K., “Feature of Internal Flow
Phenomena of Centrifugal Compressor for Turbocharger with Wide Operating
Range”, Gas Turbine Congress, 2011.
(4) Tomita, I., Ibaraki, S., Furukawa, M., Yamada, K., 2012, “The Effect of Tip
Leakage Vortex for Operating Range Enhancement of Centrifugal Compressor”,
ASME Paper, GT2012-68947.
(5) Ebisu M., Shiraishi T., Tomita I., 2012, “Development of advanced centrifugal
compressor for turbocharger, applying control of internal unsteady flow
structure”, 10th international conference on Turbochargers and Turbocharging,
IMechE, pp. 135-144.

87
A one-dimensional performance model for
turbocharger turbine under pulsating inlet
condition
H Chen1, D E Winterbone2
1
National Laboratory of Engine Turbocharging Technologies, China
2
Formerly UMIST, UK

ABSTRACT

This paper describes a new 1-D model of turbocharger turbine working under
pulsating inlet condition. The flow inside the turbine volute is treated as one-
dimensional and unsteady, but with mass removal or addition to simulate the flow
into and from the rotor. The flow in each rotor passage is treated as one-
dimensional and unsteady, and is coupled to the volute flow through a sliding rotor-
stator interface. This model is solved by the method of characteristics. The results
show that the model captures important unsteady features such as hysteresis of
turbine mass flow and efficiency. Possible improvements and extensions to the
model are also discussed.

NOMENCLATURE

A Flow passage area (m2)


a Speed of sound (m/s)
b Flow passage width (m)
Cf Coefficient in disk friction torque expression (22)
CFD Computational Fluid Dynamics
Cis Isentropic expansion (spout) speed (m/s)
Cp Specific heat at constant pressure (J/kg)
f Pulse frequency (Hz)
h Specific enthalpy (J/kg)
L Wet periphery of flow passage (m)
m Index in angular momentum equation (14)
Mf Disk friction torque (N-m)

m Mass flow rate (kg/s)
N Turbine rotating speed (rev/s)
n1, n2 Indices in incidence loss equation (17)
nb Number of rotor blades
P Pressure (N/m2)
Q Heat transfer through the walls of flow passage (J/kg-s)
R Radius (m)
Re Reynolds number
s Clearance between rotor backdisc and heat shield (m)
Str Strouhal number
t Time (s)
T Temperature (oK)
U Rotor peripheral (tip) speed (m/s)

_______________________________________
© The author(s) and/or their employer(s), 2014
113
V Absolute velocity (m/s)
W Relative velocity (m/s)
x Coordinate along flow direction in the volute (m)

Greek Symbols
 Angle between absolute and peripheral velocities
 Angle between relative velocity and radial direction
 Right-hand term of compatibility conditions along characteristics
 Ratio of specific heats
 Right-running, left-running Math-lines and path-line (characteristics)
 Cycle mean, total-to-static efficiency
 Kinematic viscosity (m2/s)
 Density (kg/m3)
 Time scale, wall shear stress (N/m2)
 Azimuth angle
 Mass withdraw from/addition to the volute per unit length per unit time
(kg/s-m)
 Rotor angular speed (rad/s)

Subscripts
0 Stagnation or total state, turbine inlet
1 Centroid of turbine housing
2 Volute exit
3 Rotor inlet (after incidence)
a Ambient
f Fluid
m Mean
p Pulse
r Rotor
s Steady flow
u Unsteady flow
w Wall

1. INTRODUCTION

Turbocharging turbines for internal combustion engines often work under pulsating
flow conditions. It has long been recognised that the performance of these turbines
can be quite different from that under steady flow conditions. Yet there is still a lack
of understanding of the phenomena, and there is no simple method that can be
used with confidence to predict the performance of the turbine under such
conditions.

The relative importance of unsteady effects produced by pulsating flow might be


estimated by a Strouhal number which is the ratio of two timescales (1):

Strf-p = f/ p (1)

where f is the time for fluid particles to be transported through the turbine
components, and p is the time scale of the unsteadiness of pulsating flow. The
following rough guides hold, if

Strf-p << 1, quasi-steady effects dominate;


Strf-p >>1, unsteady effects dominate;
Strf-p ~ 1, both quasi-steady and unsteady effects are important.

114
Take for example, a four-stroke, four cylinder engine rotating at 2000rpm and a
turbine housing with averaged flow passage length of 150mm and an averaged flow
velocity of 100m/s; these would give f, p and Strf-p values of 0.0015s, 0.015s, and
0.1 respectively. In this case both unsteady and quasi-steady effects exist in the
housing. Averaged flow passage length of turbine rotor is about a quarter of that of
the housing, so quasi-steady effects should dominate the rotor flow.

Another important time scale for turbines is the time scale for the rotor to rotate
one revolution. This time scale r is largely independent of f, and there is evidence
(2-3) that with the reduction of r or the increase of turbine speed, turbine
behaviour becomes increasingly steady. So another Strouhal number may be
needed:

Strr-p= r/ p (2)

and when Strr-p is small, steady effects may dominate.

Existing low dimensional methods for predicting turbine performance under


unsteady flow fall into three categories: steady, quasi-steady and unsteady flow
models. The steady flow method uses cycle mean inlet conditions and measured
steady flow performance maps; while the quasi-steady flow method uses the same
performance maps in a quasi-steady manner to predict instantaneous performance
from the instantaneous inlet conditions. The quasi-steady flow method may produce
poorer cycle-mean performance results than the simpler steady flow method.

The first unsteady flow model was perhaps presented by Wallace and Adgey (4) in
1967. In this model, the nozzled turbine housing is simulated by a convergent
nozzle and the one-dimensional unsteady flow equations can then be applied to the
nozzle. The rotor is treated as a single, two-dimensional duct. Results obtained in
this work (5) showed that it is possible to trace pressure waves through the turbine
albeit in a relatively crude manner. The Wallace model was modified by Mizumachi
and his co-workers (6), who reformed the equations, and simplified the rotor flow
passage to a one-dimensional duct. The modified model was used for full
admission, and showed fairly good agreement in mass flow between predictions and
experiment. They also developed a new model for partial admission of sector
divided, nozzled, twin-entry housing. In the new model, the stator and the rotor
were each divided peripherally into six sections, a typical streamline in one sector
of the scroll was branched twice and the branching points were treated as 'constant
pressure' junctions. Rotation of the rotor was simulated by transferring the
connection of the rotor passages with the nozzles one by one at a time interval of
one-sixth of a rotor revolution. The mass flow rate and the pressure waves at the
turbine entries were well predicted, but the efficiency was overestimated.

Chen and Winterbone (7) proposed a housing model that replaces the volute by an
equivalent length of nozzle, and applied one-dimensional unsteady flow equations
to this nozzle. The flow in the rotor is treated as quasi-steady, and simulated by a
meanline model. The model was later (2) applied to a mixed flow turbine. The
results showed it captured unsteady effects better than a quasi-steady model, but it
did not show any improvement on predicted cycle-mean performance versus a
steady flow method. Baines et al. (8) described a turbine model in which the volute
is represented as a volume between the turbine inlet pipe and the turbine rotor
entry. They showed that this approach, which is effectively a zero-dimensional
model, can predict some of the measured features of unsteady flow.

115
The difficulty in modelling the unsteady flow in the volute housing is that the volute
is not a simple pipe and one-dimensional gas flow equations are difficult to apply.
Treating it as a volume ignores the gas momentum equation and will inevitably lead
to error. In the remainder of this paper, a new one-dimensional unsteady flow
model of the turbine volute housing is proposed to overcome this difficulty. The
model was originally devised between 1987-88 but has not been formally
published.

2. PHYSICAL MODEL OF TURBINE

2.1 Curved slot model of volute with mass removal/addition


Under steady flow conditions, flow continuously leaves the volute along the
circumferential direction of turbine volute and enters the rotor. The same happens
under unsteady flow condition, the only difference being that the flow is a function
of time and may reverse, entering the volute from the rotor. This flow
leaving/entering the volute may be considered as mass removal/addition to the
flow inside the volute, so the volute flow can still be treated as one-dimensional,
but with additional terms to take into account this mass removal/addition, see
Figure 1. The unsteady, one-dimensional gas flow equations (9-10), as applied to
the volute model in Figure 1, are:

Continuity

1 1V1  1V1 dA1 (3)


  
t x1 A1 A1 dx1

where  is the mass withdrawn from or added to the volute per unit length per unit
time, and is expressed as:

dR1 (4)
    2V2 sin  2 b2 R2 /( R1  0.5 )
d

Momentum

1V1 1V1  p   V dA1 L1


2 2

  V2 cos( 2   1 )  1 1  sign (V1 ) w (5)


t x1 A1 A1 dx1 A1

where w is the wall shear stress, and is expressed as:

 w  0.0025 1V1 2 (6)

Energy

1h01  p1 1V1h01   V h dA (7)


  h02  1 1 01 1  1Q
t x1 A1 A1 dx1

where Q is the heat transfer through the wall, and is expressed as:

L1 (8)
Q  0.0025 V1C p (Tw  T1  0.424V1 / C p )
2

A1

116
P1 dx1 P1+dP1
1 x 1+d1
V1 1 V1+dV1
y 1
A1 A1+dA1
x R2
R1 R1+dR1 R1
 y P2, 2, V2 +d
2, A2, R2

Fig. 1 Curved slot model of turbine volute

2.2 Method of characteristics and boundary conditions


The method of characteristics (11) is used to give necessary equations for various
boundaries:


dp1  a1 1 dV1  a1  1  a1 2   3 dt
2
 (9)

along Mach lines and ,

dx1 (10)
 V1  a1
dt
and

dp1  a1 d1   3 dt (11)


2

along path line 

dx1 (12)
 V1
dt
where

1  11  12 ,  1V1 dA1 (13a)


 11  ,  12   ,
A1 A1 dx1

L1
 2   11 V2 cos( 2   1 )  V1   sign (V1 ) w , (13b)
A1

 3  (  1)11 (h02  h01 )  V1 2  Q 1 . (13c)

For the entry region of the turbine housing from the turbine inlet flange to the
tongue ( = 0o), equations (3) to (13) are used with  = 0. The instantaneous
static pressure and time mean stagnation temperature at the inlet flange are
assumed to be known from experimental measurements, instantaneous static
temperature is then calculated using the method of (12) for inflow to the inlet. The
velocity at the inlet is calculated through the  characteristic, and for outflow from
the inlet, the density at the inlet is calculated through the path line characteristic.

The volute end ( = 360o) is simplified to a closed end with V1 = V2 = 0. Two Mach
line characteristics are used to obtain the pressure and density at this point.

117
The flow between the volute exit or station 2 and the volute central line or station 1
is assumed quasi-steady, for outflow from the volute to the rotor (W3 > 0), angular
momentum equation

V2 cos  2  V1 cos  1 R1 / R2 


m
(14)

is applied, where m is an index and may be estimated from experimental data (m =


1 implies angular momentum conservation), and following assumptions are used:

P2 = P1 (15)

 (16)

At rotor inlet, a modification of NASA's incidence loss model (13-14) is used for
inflow:

W3  W2 cos n1  2 , for  2  0 ;
2 2
(17a)

2 2
 
W3  W2 1  sin n2  2 , for  2  0 , (17b)

where W2 and W3 are the relative velocities at rotor entry before and after
incidence, as shown in Figure 2. The indices n1 and n2 were selected from test data.
Energy and mass conservations and the compatibility condition along the
characteristic running from inner rotor toward the rotor entry are also used:
2
 P3 W2  P2
2
W3 (18)
  
2  1 3 2  1 2
 3W3 A3   2W2 sin  2 A2 (19)

dp  adW 
 2 W dA L dR   L 
 a  a  sign(W ) w   2 R   (  1) Q  W sign(W ) w  dt
 A dx  A R   A 
(20)

where Q and w are the heat transfer term and wall shear stress of the rotor
passages respectively.

For reversed flow from the rotor to the volute ( W3  0 ), the same equations as (14)
to (20) are used except that equation (17) is replaced by the compatibility condition
along the path line:

 L  (21)
dp  a 2 d  (  1) Q  W sign (W ) w  dt
 A 

V2 V3
W2
W3
2<0
2>0
U2 U3
Fig. 2 Velocity triangles before and after incidence at rotor inlet

118
2.3 Rotor model and solution method
Equations similar to those in (6) are used for the rotor passages, with wall friction
and heat transfer terms included. The disc friction is estimated by the following
formula (15)

M f  0.5C f  a 2 R3 ,
5
(22)

where Mf is the friction torque, Cf is given by

/( S / R3 ) 1 / 6 , and Re  R3 /
1/ 4 2
C f  0.080Re (23)

The Lax-Friedrichs algorithm (16) has been used to solve governing equations for
internal points, and the Courant-Isaacson-Rees algorithm (17) has been applied to
solve the characteristic equations for boundary points. The volute between  = 0o
to 360o was divided into nb sections with each section extending over the same
angle  = 2/nb, where nb is the number of rotor passages. Connection between
these sections and the rotor passages is carried out in every time step.

3. RESULTS

3.1 Steady flow results


Steady flow calculations were first carried out to test the model. Turbine data used
was supplied by Holset and are shown in Table 1. The values of indices m, n1 and n2
were chosen at different rotor speeds to give agreement between the predicted and
measured performances. These steady state results are shown in Figures 3-4. The
maximum error in the predicted mass flow is about 5%. Given the fact that only a
very simple friction loss model was used for the rotor passages, and some
important losses such as clearance loss and bearing loss were not included, the
over-predicted efficiency is encouraging.

Table 1. Main input data

Housing inlet area 4.71 x10-3m2


Housing end ( =0) area 0.163 x10-3m2
Width of volute exit (station 2) 0.0171m
Number of rotor blades 12
Rotor inlet radius 0.0485m
Rotor inlet area 4.125 x10-3m2
Rotor exit radius 0.0295m
Blade angle (trailing edge shroud) 63.5o
Rotor exit area 1.73 x10-3m2
Gas constant 287.1 J/kg-oK
Gas specific heat ratio  1.4
Gas viscosity  1.5 x10-5m2/s
Turbine inlet total temperature 400 oK
Wall temperature 320 oK
Ambient pressure Pa 1.0133 x105N/m
Ambient density a 1.225 kg/m3

119
8
P00 / Pa  2
7
6
5

P00 / Pa  1.2
4
3
2

N / T00
1

0 10 20 30 40 50 0 10 20 30 40 50
Fig. 3 Steady mass flow prediction Fig. 4 Steady efficiency prediction

3.2 Unsteady flow results


Computations were carried out to investigate the influence of pulsating flow on
turbine performance. The results are shown in Figures 5-10. Figures 5-6 show the
mass flow and efficiency characteristics under a pulsating inlet condition of P00/Pa =
1.4-0.2sin(270t) and N / T00,m  30 . The turbine operating loci show hysteresis
between expansion pressure ratio and non-dimensional mass flow rate, and
hysteresis between efficiency and blade speed ratio as obtained by Dale and
Watson (20). Figure 7 shows the variation of flow angle at volute exit with time at
two different circumferential locations. Strong pulsating of the angle at the tongue
region, Figure 7a, is damped at the middle of the volute as shown in Figure 7b.
Waves with different shapes and periods were used to investigate the influence of
such parameters as frequency, amplitude and shape on the turbine performance.
The results are given in Figures 8-10, where unsteady flow results are plotted
against steady flow results. Figure 8 shows a slight increase of mass flow rate and a
slight decrease in efficiency with increasing pulse frequency. These results agree
with those obtained from experiment in (12). Figure 8b also suggests that the
steady flow method might overestimate as well as underestimate both the
efficiency and power depending on non-dimensional turbine speed N / T0 . The
figure further suggests that the steady flow method might give a better prediction
of turbine efficiency at higher turbine speeds. This is consistent with the influence
of the Strouhal number Strr-p as discussed earlier: for the highest pulse frequency
of 100Hz modelled, the values of Strr-p as expressed by equation (2), are 0.5, 0.
167 and 0.106 respectively for the three turbine speeds in the figures. Increasing
the pulse amplitude reduces the mass flow rate as shown in Figure 9a. The
relationship between turbine efficiency and the pulse amplitude, indicated in Figures
9b, shows a strong dependence on turbine speed. The influence of pulse shape is
shown in Figure 10. The stronger the pulsating is, the lower the turbine efficiency
tends to be. The figure also suggests that a steep wave-front might result in a
lower turbine efficiency.

120
1.9

1 .0 1 .1 1 .2
P00 ( t ) / Pa
1.7 1.8

0 .9
1.6
1.5

0 .3 0 .4 0 .5 0 .6 0 .7 0 .8
1.4
1.3
1.2
1 1.1

U / C is , m
2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 0.2 0.4 0.6 0.8 1.0 1.2

Fig. 5 Unsteady mass flow locus Fig. 6 Unsteady efficiency locus

28
58

2, 180 (t )

54 56

27
26
52

N / T00,m  30
50

25
46 48

24
40 42 44

23

N / T00,m  30
22

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
(a) =330deg (b) =180deg

Fig. 7 Flow angle variation with time and location

  P00 / Pa  1 . 6  0 . 6 sin( 2  ft )
1

mu/ ms  t  s ,u / t  s , s
1.05
0.98

1
0.96

0.85 0.9 0.95


0.94
0.92

0.75 0.8

f (Hz)
0.9

0 20 40 60 80 100 0 20 40 60 80 100
f (Hz)
 
(a) m u / m s (b) t  s ,u /t  s , s
Fig. 8 Effect of pulse frequency on mass flow and efficiency

121
1.1
0.98 1
t s,u /t s,s
 
m u/ m

1.0
s
0.94

0.9
0.90

0.8
0.82 0.86

0.7
X X

0.6
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
 
(a) m u / m s (b) t  s , u / t  s , s
Fig. 9 Effect of pulse amplitude X on mass flow and efficiency

 t  s ,u

0.8
m T00,m / P00,m *105
3.6 3.8 4.0

0.7
0.6
0.5
3.0 3.2 3.4

Steady flow
P00 / Pa  1.2  0.2 sin( 2 50t )
0.4
0.3
2.6 2.8

0.1 0.2

N / T 00 , m N / T 00 , m

0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35

(a) m T00, m / P00, m *105 (b) t  s

Fig. 10 Effect of pulse shape on mass flow and efficiency

4. DISCUSSION

The model presented here treats turbine housing more rationally than in other
previous models, and takes into consideration more geometrical parameters of the
housing. However, since the model was first conceived over two decades ago,
improvements are now appropriate and possible. Firstly, equations (15) and (16)
for volute exit should be replaced, perhaps by adiabatic flow assumption and mass
conservation between stations 1 and 2; secondly, the closed end treatment of the
volute end may be replaced by a junction model allowing nonzero velocity at the
end; thirdly, better rotor loss models, perhaps adopting first the steady flow loss
modelling technique, could be used to improve the accuracy of the performance
prediction.

Application of the current vaneless housing model to vaned housing is straight


forward. Station 2 is now the inlet to the nozzles which can be treated similarly to
the rotor inlet, but in the stationary frame. Extending the model to twin-entry
turbine housing, whether meridionally divided or circumferentially divided, is

122
possible. For the circumferentially divided housing, the extension is simpler and
more straightforward; for the meridionally divided housing, additional modelling is
required to allow for the interaction between the two branches of the volute, and
between the volute and the rotor. A three-way junction may be suitable for this
task.

5. REFERENCES

1. Greitzer E. M., An introduction to unsteady flows. Vol. 2 of Thermodynamics


and fluid mechanics of turbomachinery, edited by Ucer A. S., Stow P. and
Hirsch C., NATO ASI Series, Martinus Nijhoff Publishers, 1985.
2. Chen H., Hakeem I. and Martines-Botas R. F., Modelling of a turbocharger
turbine under pulsating inlet conditions. Proc. IMechE, Pt. A, Vol. 210, pp.
397-408, 1996.
3. Private communication with Honda simulation engineers in Japan, 2011.
4. Wallace F. J. and Adgey J. F., Theoretical assessment of the non-steady flow
performance of inward radial flow turbines. Proc. IMechE, Vol. 182, Pt. 3H, pp.
22-36, 1967-8.
5. Wallace F. J., Adgey J. F. and Blair G. P., Performance of inward radial flow
turbines under non-steady flow conditions. Proc. IMechE, Vol. 184, Pt. 1, No.
10, pp. 183-195, 1969-70.
6. Mizumachi N., Yoshiki H. and Endoh T., A study on performance of radial
turbine under unsteady flow conditions. Report of Inst. Indus. Sci., Uni.
Tokyo, Vol. 28, No. 1, Dec. 1979.
7. Chen H. and Winterbone D. E., A method to predict performance of vaneless
radial turbines under steady and unsteady flow conditions. IMechE 4th Int.
conf. Turbocharging and Turbochargers, Paper C405/008, Proc. IMechE, pp.
13-22, 1990.
8. Baines N. C., Hajilouy-Benisi A. and Yeo J. H., The pulse flow performance and
modelling of radial inflow turbines. IMechE 5th Int. conf. Turbocharging and
Turbochargers, Paper C484/006, Proc. IMechE, pp. 209-18, 1994.
9. Zucrow M. J. and Hoffman J. D., Gas dynamics. Vol. 2, John Wiley & Sons,
Inc. 1977.
10. Woods W. A. and Allison A., Unsteady compressible flow with gradual mass
addition and area change. IMechE Sixth Thermodynamics and Fluid Mechanics
Convention, IMechE conf. Pub. 1976-6, pp 179-186.
11. Benson R. S., The thermodynamics and gas dynamics of internal combustion
engines. Vol. I, edited by Horlock J. H. and Winterbone D. E., Oxford Uni.
Press, 1982.
12. Benson R. S., Non-steady flow in a turbocharger nozzless radial gas turbine.
SAE Meeting Milwaukee, SAE Paper 740739, Sept. 1974.
13. Wasserbauer C. A. and Glassman A. J., Fortran program for predicting off-
design performance of radial-inflow turbines. NASA Technical Note D-8063,
1975.
14. Meitner P. L. and Glassman A. J., Off-design loss model for radial turbine with
pivoting, variable-area stators. NASA Technical Report 80-C-13, 1980.
15. Glassman A. J. et al, Turbine design and application. NASA SP-290, 1972-75.
16. Lax P. D., Weak solution of non-linear hyperbolic equations and their
numerical computation. Commun. Pure and Appl. Math., Vol. 7, pp. 159-193,
1954.
17. Courant R., Isaacson E. and Rees M., On the solution of non-linear hyperbolic
differential equations by finite differences. Commun. Pure and Appl. Math.,
Vol. 5, pp. 243-355, 1952.
18. Dale A. and Watson N., Vaneless diffuser turbocharger performance. IMechE
3rd Int. conf. Turbocharging and Turbochargers, Paper C110/86, Proc.
IMechE, pp. 65-76, 1986-4.

123
The high temperature tribological
performance of turbocharger wastegate
materials
M Burkinshawa, D Blackerb
a
Materials Engineering, Cummins Turbo Technologies, UK
b
Advanced Engineering, Cummins Turbo Technologies, UK

ABSTRACT

The high temperature tribological interface between shaft and bush in a


turbocharger wastegate has been investigated using a pin-on-plate reciprocating
tribometer, which was operating at 600 °C, 850 °C and 950 °C. Contact pressure,
sliding distance and speed were kept constant throughout testing. A wide range of
shaft and bush materials were evaluated for their suitability within the previously
mentioned interface, namely; cobalt-chromium alloys and a wide range of
austenitic, martensitic and sintered powder metal alloy steels. Materials selected for
use within the interface had to possess adequate friction and wear performance
whilst in operation. The wear performance of the evaluated material combinations
have been determined and related to the topography and chemistry of the new and
worn surfaces in order to ascertain the fundamental wear behaviour of the alloys
under investigation. Wear models for the various material combinations have been
proposed and related to those experienced by the wastegate whilst in operation.

1. INTRODUCTION

Turbocharging is a fundamental technique employed by engine manufacturers in


order to minimise emissions, maximise power and reduce fuel consumption of the
internal combustion (IC) engine. Since turbocharging is used in conjunction with
diesel, gasoline and natural gas engines, the maximum exhaust gas temperature
experienced by the turbocharger turbine housing can vary significantly.

In order to ensure high levels of reliability and durability of the turbocharger, it is


important to characterise the tribological performance of particular components.
The shaft / bush interface is a critical tribosystem within the turbocharger
wastegate, itself a fundamental component within modern turbomachines. Since
the interface of interest resides within the turbine housing of the turbocharger,
temperatures exceeding 600 °C are experienced by the tribosystem. Therefore, it is
imperative that the material combination for the shaft / bush tribosystem is
extensively characterised in order to afford sufficient confidence while in operation.

Traditionally, bushings and shafts in modern turbomachinery wastegates have been


produced from stainless steel or cobalt-based alloys, in order to afford sufficient
wear behaviour and performance for the interface. The latter material is selected
due to excellent strength, corrosion and tribological performance (1) over a wide
temperature range. However, the price of such materials is prohibitive for
employment within cost-sensitive applications and therefore, alternative lower cost
alloys are under consideration (2; 3).

_______________________________ 289
© Cummins Turbo Technologies, 2014
The intention of this article is to present the innovative methods by which
conventional and novel materials intended for use within the wastegate system of a
turbocharger are characterised using a tribometer and associated surface
characterisation techniques. Tribological testing affords the acquisition of accurate
and repeatable friction and wear data from substrates of interest. Such testing is
vital in order to validate materials and test conditions prior to more complex and
expensive component-based experiments. Ultimately, effective simulation
techniques will improve the efficiency and reduce the cost of product development.

The high temperature tribological performance of a range of austenitic, martensitic


and sintered powder metal alloy steels have been presented. High magnification
topographical images of substrates were obtained using Scanning Electron
Microscopy (SEM). Secondary Ion Mass Spectroscopy (SIMS) was used to
determine the tribochemistry of worn substrates. Wear behaviour and results are
compared and related to those which would be experienced by a wastegate in
operation.

2. EXPERIMENTAL

2.1. The turbocharger wastegate shaft / bush interface


Figure 1 is a computer aided design model of a typical turbocharger wastegate,
including the valve, shaft and bush. The latter is fixed in position within the turbine
housing of the turbocharger and the shaft rotates within the bush about a given arc
length.

Figure 1: Typical wastegate shaft / bush

2.2. Materials
Numerous high temperature materials were evaluated; these are listed in Table 1.
The test materials were selected due to their suitability for a wastegate application,
together with their cost and predicted tribological and corrosion performance.
Material combinations C and D were novel selections in order to cater for high
temperature wastegate applications, whereas combinations A & B have found
previous use on turbochargers with a turbine inlet temperature < 800 °C.

Table 1: Test substrates


Material
Pin / Disc Hardness (HRC)
Combination
1 (austenitic stainless steel) / 2
A 18.9 ± 0.2 / 24.9 ± 0.2
(martensitic stainless steel)
B 3 (cobalt alloy) / 4 (cobalt alloy) 38.9 ± 0.5 / 51.5 ± 0.4
5 (austenitic stainless steel) / 6
C 20.3 ± 1.3 / 52.5 ± 0.3
(intermetallic stainless steel)
7 (nickel superalloy) / 8 (sintered
D 33.3 ± 0.1 / 23.6 ± 0.4
powder metal stainless steel)

290
Since the turbocharger wastegate shaft / bush interface operates through rotary
motion over a short arc length, it thereby induces a top and bottom dead centre
onto the wear scar. Therefore, this type of interface can be successfully simulated
using a reciprocating pin-on-plate/disc tribosystem, although with careful
consideration of external factors which may affect the correlation performance
between tribometer and real-world tests.

Disc samples simulating the wastegate bush were machined with 35.0 mm
diameter, 10.0 mm depth. Pins representing the wastegate shaft were
manufactured with dimensions 10.0 mm diameter, 30.0 mm length and 37.5 mm
radius of curvature on the end of the pin. All substrates were used in as-received
form without heat treatment, with a surface roughness value of 0.2 µm Ra.

2.3. Methods
Tribological evaluation of the test
materials was conducted using a
modified version of ASTM G133
and a Bruker UMT-3 high
temperature pin-on-reciprocating
plate / disc tribometer (Figure 2).
Samples were submerged in
acetone and ultrasonically cleaned
for 10 minutes prior to each test.
The contact conditions (Table 2)
for the tribometer were designed
to replicate those experienced by Figure 2: Schematic diagram of pin-
a wastegate in a Cummins Turbo on-reciprocating plate tribometer
Technologies D6 turbocharger.

The peak contact stress for the shaft / bush interface was calculated using equation
Equation 1-4 (4), with P determined using a free body diagram of the reaction force
of the shaft in the bush.

P
(σ ) = 1.5 =
πa +
Equation 1 (4) Equation 3 (4)

= 0.721 √( ) 1− 1−
Equation 2 (4) = +
Equation 4 (4)

σc = Contact stress (Pa) ν1, ν2 = Poisson’s ratio bush & shaft


P = Applied force (N) E1, E2 = Young’s modulus bush &
a = Contact radius (m) shaft (Pa)
Kd = Equivalent diameter (m) D1 = Bush inner diameter (m)
CE = Equivalent elastic modulus (Pa) D2 = Shaft outer diameter (m)

The relative bush / shaft displacement was determined by considering a typical


angular rotation of the wastegate shaft of 15 degrees and a shaft diameter of 10
mm. The tribometer frequency of reciprocation and time duration of test were
established by assuming the wastegate performed 216 x 103 cycles, which afforded
sufficient comparison between the tribological performance of material pairs. The
previously mentioned number of cycles would equate to 87 x 103 km on engine.
The effect of exhaust gas pulsation was considered negligible for the purpose of the
testing reported here.

291
The maximum predicted contact stress for the shaft / bush interface was replicated
for the pin-on-plate tribosystem using an applied load of 4 N. Table 2 states the
test parameters for the tribological experiments.

Table 2: Tribological test parameters

Parameter Value
Maximum contact pressure 180 MPa
Total stroke length 1.3 mm
Frequency 30 Hz
Operating temperatures 600 °C, 850 °C and 950 °C
Test duration 2 Hours
Test Medium Air

2.4. Evaluation of worn substrates


High magnification images of surfaces of interest were obtained using a Philips™
XL30 Scanning Electron Microscope. Low magnification images were obtained using
a Leica® DM LM microscope. The tribochemistry of the worn substrates was
obtained using a ToF-SIMS IV Secondary Ion Mass Spectrometer. Volumetric wear
data was obtained from three repeats per material combination using an Alicona™
InfiniteFocus® confocal microscope, which had a lateral and vertical resolution of
≤ 2.2 μm and 410 nm, respectively.

3. RESULTS

3.1. Volumetric wear data


Combination A possessed the worst wear resistance at all temperatures evaluated.
Furthermore, it was not possible to accurately measure the wear volume for this
material combination at 950 °C because of an excessive oxide formation rate. The
cobalt-based material combination B possessed the greatest wear resistance of all
four material pairs at all testing temperatures, except at 600 °C for the pin. The
wear resistance of combination D was superior to combination C at 600 °C, 850 °C
and 950 °C.

Figure 3: Wear volumes measured on (A) pins and (B) discs

3.2. Images of worn substrates

3.2.1. 600 °C
The wear scar induced into the substrates tested at 600 °C are shown in Figure 4.
The worn regions of all pin samples, namely materials 1, 3, 5 and 7, appeared to
consist of numerous abrasive wear tracks which were of non-homogenous height
distribution.

292
The wear scar generated in material 2 appeared extremely course and multiple
abrasive wear tracks were evident throughout. In contrast, the wear scars observed
on materials 4 and 8 appeared smooth, with the latter containing a number of
raised sections inside the wear scar. The wear scar of material 6 appeared uneven
and contained a number of pores. There were regions within the wear scars of
materials 6 and 8 which appeared to be indicative of material removal.

Figure 4: Experiments conducted at 600 °C. 20x magnification optical


microscope images of the wear scar on test pins (materials 1, 3, 5 & 7)
and 200x magnification SEM images of the wear scar on test plates
(materials 2, 4, 6 & 8). Arrow indicates sliding direction.

3.2.2. 850°C
As can be observed in Figure 5, at 850 °C, the wear scars generated in materials 1
and 2 were uneven and appeared to contain wear debris throughout. This was in
stark contrast to materials 3, 4, 6 & 8, in which a topographically smooth surface
was generated post-test. The wear scar induced into material 5 was dominated by
abrasive wear tracks, whereas the wear region of material 7 was smooth but
contained numerous pores.

Figure 5: Experiments conducted at 850 °C. 20x magnification optical


microscope images of the wear scar on test pins (materials 1, 3, 5 & 7)
and 200x magnification SEM images of the wear scar on test plates
(materials 2, 4, 6 & 8). Arrow indicates sliding direction.

3.2.3. 950°C
Non-uniform dimensioned oxide particles were observed in the wear scar of
material 2 when operating at 950 °C, as shown in Figure 6. The worn regions in
materials 3 and 4 possessed a smooth appearance. The wear scar of materials 6, 7
and 8 were relatively smooth, but were also non-uniform in height distribution.
Multiple abrasive wear tracks were evident in materials 1 and 5. In addition, there
was a definable lay to the wear scar of material 8.

293
Figure 6: Experiments conducted at 950 °C. 20x magnification optical
microscope images of the wear scar on test pins (materials 1, 3, 5 & 7)
and 200x magnification SEM images of the wear scar on test plates
(materials 2, 4, 6 & 8). Arrow indicates sliding direction.

3.3. Chemical nature of worn substrates

3.3.1. SIMS
Chemical images of worn substrates were obtained using secondary ion mass
spectrometry. This technique afforded direct comparison of the chemistry of regions
inside and outside the wear scar. Figure 7 shows a typical series of images obtained
from the cobalt-based material pair, combination B.

Figure 7: Typical chemical images for Co and Cr obtained using ToF-SIMS.


Total ion count is listed as “TC”. Ion intensity is represented by the scale
on the images. A & B are obtained from material 3 (pin) and C & D from
material 4 (disc).

The main elements and oxides identified within the wear scar on both pin and disc
are stated in Table 3.

294
Table 3: Chemical composition of wear scars

Pin - Disc –
Material
Pin / Disc Elements of Elements of
Combination
Interest Interest
1 (austenitic stainless Cr, Fe, Mn, Cr, Fe, Mn,
A steel) / 2 (martensitic Ni, O, CrO2, O, CrO2,
stainless steel) FeO2, MnO2 FeO2, MnO2
Co, Cr, Mo,
3 (cobalt alloy) / 4 Co, Cr, O,
B O, CrO2,
(cobalt alloy) CrO2, CoO
CoO, MnO2
5 (austenitic stainless
Cr, Fe, Mn, Cr, Fe, Mn,
steel) / 6
C O, CrO2, O, CrO2,
(intermetallic stainless
FeO2, MnO2 FeO2, MnO2
steel)
7 (nickel superalloy) /
Cr, Fe, Ni, O, Fe, Ni, O,
D 8 (sintered powder
CrO2, FeO2 CrO2, FeO2
metal alloy steel)

4. DISCUSSION

4.1. Tribological performance of test substrates


Referring to section 3 of this article, it is apparent that the wear performance of the
material pairs evaluated in this article at both 850 °C and 950 °C can be classified
in the following order, from best to worst performing:

1. Cobalt-based super alloy pair (combination B)


2. Nickel super alloy / stainless steel ( combination D)
3. Stainless steel pair (combination C)
4. Stainless steel pair (combination A)

The wear protective afforded by cobalt-based alloys is well reported in literature (1;
5) and it has been reported that nickel-based alloys possess inferior wear
protection compared to their cobalt-based counterparts (6; 3). Interestingly,
referring to section 3.3.1., the high concentration of Fe and FeO2 on the nickel-
based pin is evidence that material transfer from disc to pin occurred with material
combination D. It can therefore be inferred that such behaviour influenced the wear
performance afforded by this material combination. Indeed, further work will be
required to define whether the observed material transfer reduced wear in the
interface through steel / steel rather than steel / nickel alloy interaction.

Referring to section 3, it is apparent that the wear experienced by material


combinations A, B and C are predominantly of the abrasive type. However, due to
the evidence of material transfer with material combination D, it is hypothesised
that adhesion also contributed to the wear behaviour in this interface.

As stated by Inman et al., (7), effective high temperature wear materials must
retain strength and hardness when at temperature and the results presented here
agree with this statement. Indeed, it is apparent that the materials with the
greatest yield strength (Table 4) possessed the greatest wear resistance. In
contrast, the stainless steel material pair C possessed considerably lower yield
strengths at 850 °C, which were in the order of 200 MPa, with material combination
A possessing the lowest yield strength of all substrates evaluated at 90 MPa.

295
Table 4: Material yield strength at 850 °C

Material Yield Strength (MPa)


Cobalt alloy (3) 270
Cobalt alloy (4) 620
Nickel superalloy 680
Sintered powder metal alloy steel 400

The slight deviation from the rule is material pair D, which despite possessing the
greatest strength, only provided the second greatest wear response. Therefore, it is
apparent wear resistance was not dependent entirely on strength and additional
factors may be contributing to wear performance.

One such factor is the formation and stability of generated oxide layers. Referring
to the condition of material combination A post test, the oxide layer was
topographically rough and became increasingly unstable with increasing
temperature. It was noted by the current authors that all other material pairs
generated a more stable and smooth oxide layer post-test at all testing
temperatures, compared to material combination A.

It is clear therefore that both high strength and oxide layer stability are desirable
characteristics of a high temperature wear resistant material pair. An additional
factor that contributes to the wear resistance is glaze formation and the chemistry
of the oxide layer which is considered in the following section.

4.2. Glaze formation and the effect on wear performance


The ideal scenario for a high temperature tribological contact would be the
generation of a glaze, a composite layer generated from oxidised wear debris (8),
on the contacting surfaces. Frequency of reciprocation, normal load, and operating
temperature are fundamental parameters with regards to glaze formation for a
particular alloy (9). Glaze formation occurs through a complicated process of wear,
oxidation and mechanical mixing (8; 10) and subsequently improves tribological
performance of a given interface (8; 10).

Referring to Wood et al., (10) it would appear that the wear scar generated on
materials 3 and 4 shown in Figure 5 and Figure 6 closely resemble glaze formation
on similar cobalt based alloys. Furthermore, the glaze generated on the cobalt-
based materials was chemically similar to that identified by Wood et al, (10).
Further work is required to determine the chemistry of the glaze through depth
profiling.

Referring to the topography of the glaze generated on the cobalt based material in
Figure 5 and that reported on numerous materials (10; 9; 11; 12), it is suggested
by the current authors that glaze formation has also occurred on materials 6 and 8
at 850 °C and 950 °C.

Therefore, it is imperative that material combinations are selected for their ability
to generate successful glaze layers which ensure a sustainable wear rate and
frictional response in the tribosystem over the useful life of the component of
interest.

4.3. Tribometer testing as a simulation for wastegate wear behaviour


On the intended engine application, the wastegate shaft / bush will experience
operating temperatures of up to 950 °C. However, in order to provide a reliable and
accurate baseline by which comparisons could be made to tribometer test samples,
wastegate shaft and bush components produced from material combination A were
sourced. These substrates were retrieved from a Cummins Turbo Technologies

296
turbocharger which had operated on engine with a maximum operating
temperature of approximately 650 °C.

Abrasive wear was the dominant wear mechanism in the field return shaft / bush
tribosystem, which correlated well to the results presented for material combination
A. Therefore, this is justification to suggest that the research and methodology
reported here provides a cheap, high-efficiency tool to characterise high
temperature tribological materials for wastegate applications and identify wear
mechanisms and oxide / glaze layer formation.

Figure 8: Worn shaft and bush from field-return turbocharger

Further testing and analysis is required in order to provide additional data points for
correlation, in terms of wear rate and tribochemistry on both substrates. Such
additional information will increase confidence with regards to tribometer-based
testing and provide further evidence that simulated wastegate testing can
accurately and reliably determine the tribological performance of material
combinations for a given wastegate application. The current authors are of the
opinion that tribometer-based testing can remove the need for particular functional
tests, improving the efficiency and reducing the cost of product development.

Referring to the wear behaviour of the four material combinations, it is


hypothesised that excessive oxide formation, as experienced with material
combination A at 950 °C, could potentially reduce the clearance between shaft and
bush whilst in operation. However, more stable alloys such as combinations B & D
should not suffer such issues. The adhesive wear behaviour of the latter
combination needs to be further investigated to ensure that this does not impact
the durability of components whilst in operation.

One item of interest for the current authors is the behaviour of wear debris
produced through sliding in both the pin-on-disc and shaft / bush interfaces, since
the former testing technique may permit a greater rate of debris removal compared
to the latter. Variation in wear debris behaviour may affect tribochemistry, glaze
formation and wear protection. It is anticipated that further research into this topic
will improve the correlation of simulated and application testing.

5. CONCLUSIONS

A wastegate shaft / bush interface has been successfully simulated using


tribometer-based experiments, which were performed at 600 °C, 850 °C and 950
°C. Contact conditions were derived from a production turbocharger.

All material combinations evaluated in this research provided significantly greater


wear resistance than material combination A at all testing temperatures. Superior
wear performance was afforded by combination B, which was compiled of two
cobalt-based alloys. There was a strong correlation of material yield strength to

297
tribological performance, with further wear protection afforded by the generation of
stable glaze layers within the wear scars on tested substrates. Material transfer
from steel to nickel-based alloy was observed with material combination D.

Further work is envisaged in order to understand the effect of wear debris on glaze
formation and tribological performance. Additional work is also required to generate
further evidence for the correlation accuracy and reliability between tribometer-
based and application testing of wastegate materials.

6. ACKNOWLEDGEMENTS

Further thanks are given to David Scurr from The University of Nottingham for
obtaining the SIMS data and Leigh Fleming from Huddersfield University for
volumetric wear analysis.

7. REFERENCE LIST

1. Characteristics of Tribaloy T-800 and T-900 Coatings on Steel Substrates by


Laser Cladding. Tobar, M J, et al. 2008, Surface and Coatings Technology,
Vol. 202, pp. 2297-2301.
2. Room Temperature Mechanical Properties and Tribology of NICRALC and Stellite
Casting Alloys. da Silva, W S, et al. 2011, Wear, Vol. 271, pp. 1819-1827.
3. The Galling Wear Resistance of New Iron-Base Hardfacing Alloys: A Comparison
with Established Cobalt- and Nickel-Base Alloys. Ocken, H. 1995, Vols. 76-77,
pp. 456-461.
4. Roark, R J, Young, W C and Budynas, R G. Bodies under Direct Bearing and
Shear Stress. Roark's Formulas for Stress and Strain. s.l. : McGraw Hill, 2002,
Vol. 7, pp. 702-703.
5. Improved Mechanical and Tribological Properties of Tin-Bronze Journal Bearing
Materials with Newly Developed Tribaloy Alloy Additive. Tavakoli, A, Liu, R
and Wu, X J. 2008, Materials Science and Engineering A, Vol. 489, pp. 389-
402.
6. High Temperature Sliding Wear Behaviour of Inconel 617 and Stellite 6 Alloys.
Birol, Y. 2010, Wear, Vol. 269, pp. 664-671.
7. Studies of High Temperature Sliding Wear of Metallic Dissimilar Interfaces.
Inman, I A, et al. 2005, Tribology International, Vol. 38, pp. 812-823.
8. Microscopy of Glazed Layers Formed during High Temperature Sliding Wear at
750 C. Inman, I A, et al. Wear, Vol. 254, pp. 461-467.
9. Development of a Simple Temperature Versus Sliding Speed Wear Map for the
Sliding Wear Behaviour of Dissimilar Metallic Interfaces II. Inman, I A and
Datta, P S. 2008, Wear, Vol. 265, pp. 1592-1605.
10. Investigation into the Wear Behaviour of Tribaloy 400C During Rotation as an
Unlubricated Bearing at 600 oC. Wood, P D, Evans, H E and Ponton, C B.
Wear, Vol. 269, pp. 763-769.
11. Studies of High Temperature Sliding Wear of Metallic Dissimilar Interfaces II:
Incoloy MA956 versus Stellite 6. Inman, I A, Rose, S R and Datta, P K.
2006, Tribology International, Vol. 39, pp. 1361-1375.
12. Galling Mechanisms during Interaction of Tool Steel and Al-Si Coated Ultra-High
Strength Steel at Elevated Temperature. Pelcastre, L, Hardell, J and
Prakash, B. 2013, Tribology International, Vol. 67, pp. 263-271.
13. Improvement of the Oxidation Resistance of Tribaloy T-800 alloy by the
Additions of Yttrium and Aluminium. Zhang, Y D, et al. Corrosion Science, Vol.
53, pp. 1035-1043.

298
Development of twin-entry scroll radial
turbine for automotive turbochargers
using unsteady numerical simulation
T Yokoyama, T Hoshi, T Yoshida, K Wakashima
Mitsubishi Heavy Industries, Ltd, Japan

ABSTRACT

In recent years, responding to growing interest in environmental issues, in


particular many automotive technologies for fuel efficient vehicles and
countermeasure to emission control have been developed actively. Among such
technologies, turbochargers have been increasingly applied for downsizing of
engines. In addition, Fuel consumption at actual engine operating conditions such
as low engine speed are focused. Twin-entry scroll turbine gains increased attention
because of high efficiency at low engine speed by utilizing engine exhaust gas
pulsation. For further enhancement of fuel efficiency it is necessary to improve
efficiency of turbochargers. On the other hand, a radial turbine used for a
turbocharger operates under exhaust gas pulsation. Because large pulsation
increases flow unsteadiness, it was difficult to understand flow phenomena and
improve turbine efficiency.

In this study, Numerical fluid simulations of twin-entry-scroll with whole turbine


blade were conducted to investigate the flow phenomena under engine exhaust gas
pulsation. Two features of flow phenomena generating loss were confirmed. Exhaust
gas is induced into the front and rear scrolls alternately and leakage inflow from one
side to no-entry side is occurred. Flow separation at blade hub side is occurred by
the distorted inflow from the scroll at the partial admission of rear side. New scroll
was designed to reduce alternating leakage inflow and flow separation at blade hub.
The authors confirmed partial admission efficiency at rear side is increased by 2
point via experiment. This corresponds to improvement of cycle average turbine
efficiency.

1. INTRODUCTION

Along with growing interest in global environmental issues, regulations regarding


emission gas and fuel economy of automobiles have been tightening year by year.
An engine with a turbocharger, which feeds compressed air to the engine, can use
smaller displacement than normal aspiration engines. Therefore turbochargers can
reduce weight and friction loss of the engines, resulting in improvement of fuel
efficiency and reduction of CO2. In particular European automotive manufacturers
have been increasingly employing smaller engines for improvement of fuel
efficiency. Therefore use of turbochargers has been growing, which results in
growth of the turbocharger market.

An automotive turbocharger uses exhaust gas to rotate the turbine, and then the
compressor bladed wheel, the shaft of which is connected directly with the shaft of
the turbine, is driven. A standard turbocharger uses a single scroll turbine.
However, at lower engine speed, it cannot compress air supply sufficiently because
exhaust gas flow is small and the boost pressure is low. To raise boost pressure, a
twin scroll turbine is effective(1).

_______________________________________
© The author(s) and/or their employer(s), 2014
471
Fuel economy shown on automobile brochures is measured according to a certain
driving pattern (test cycle)(2) designated in the country or local region. In European
countries, NEDC (New European Driving Cycle)(3) has been used since its
introduction in 2000. Now European countries are planning test cycles where fuel
economy and emission gas of vehicles are measured under conditions simulating
actual driving conditions as closely as possible. They are considering use of CADA
(Common Artemis Driving Cycle)(4) etc. in the future. In such future test cycles, it
is predicted that a driving mode with repeating acceleration from lower engine
speed and deceleration will be added in consideration of driving in urban area.
Therefore it will be necessary for complying with such regulations to improve fuel
efficiency at lower engine speed further than ever.

Improvement of fuel efficiency at lower engine speed requires enhancement of


turbocharger efficiency. However, because internal flow of a twin scroll turbine was
made unsteady by largely pulsing flow, it is difficult to analyze the internal flow, and
also its detail has not been well understood. This document presents work on
improvement of twin scroll turbine efficiency with use of the unsteady
computational simulation.

2. NUMERICAL SIMULATION METHOD AND SIMULATION CONDITIONS

In the past, a twin scroll turbine was analyzed under steady operating conditions as
shown in Figure 1 (i). Figure 6 (i) illustrates internal flow results of conventional
steady analysis at a constant pressure as will be described in 3.3. Because such
conventional analysis used a constant pressure that represented a mean exhaust
pulsation condition, unsteadiness of internal flow that actually occurred as a result
of exhaust pulsation flowing into the two scrolls alternatively and intermittently was
unknown. MHI, which is a manufacturer of both engines and turbochargers, has
taken its advantage to perform unsteady analysis that simulates exhaust pulsation
as shown in Figure 1 (ii).

To analyze unsteady phenomena under conditions of exhaust pulsation, MHI


performed flow analysis using CFD (Computational Fluid Dynamics). The analysis
objects were entire turbine including the scroll and the turbine rotor blade (turbine
diameter 52mm, blade number 12). As shown in Figure 2, the computational grid
used contained 2.07 million cells on the turbine rotor blade and 1.44 million cells on
the scroll, 3.51 million cells in total. The numerical flow analysis used ANSYS CFX
general-purpose 3D viscous flow simulation code. The turbulent model used was the
k-ε model. This analysis was performed at lower engine speed of 2400 rpm
(turbocharger speed of 125700 rpm).

Fig. 1 Turbine pressure Fig. 2 Grid for analysis

472
There occurs no exhaust pulsation interference from a two- or one-cylinder engine,
a little from three-cylinder, and intense from four-cylinder. It is assumed that
exhaust efficiency of a single scroll combined with a four-cylinder engine is reduced
by this interference. On the other hand, there occurs no exhaust interference for a
twin scroll, each scroll of which corresponds to a scroll connected to a two-cylinder
engine. Figure 3 illustrates exhaust pulsation in a twin scroll combined with a four-
cylinder engine (Engine simulation results at turbine inlet). Analysis of operation of
a twin scroll shows that partial inflow from one side occurs in almost all range of
engine crank angle and perfect full inflow from both sides occurs only at points
where pulsation switches. Therefore the most important factor for performance
evaluation of a twin scroll is when inflow from one side occurs.

Fig. 3 Exhaust pulsation in twin scroll


(combined with four-cylinder engine)

3. NUMERICAL SIMULATION RESULTS

3.1 Analysis of internal flow phenomena of turbine under conditions of


exhaust pulsation
For a conventional shape, it was found that mean in-cycle turbine efficiency during
inflow from the rear side was lower than that during inflow from the front side. To
understand the cause of lowering of turbine efficiency during inflow from the rear
side, MHI focused attention on formation of internal flow loss during inflow from the
rear side.

Figure 4 illustrates distribution of loss (i.e. entropy) inside a turbine during inflow of
exhaust gas from the rear side. Loss is generated in the scroll with time. The
generated loss flows down in circumferential direction of the scroll, and propagates
through the rotor blade to the outlet. As time passes, loss in the front side scroll,
which is not fed, becomes larger.

To analyze the cause of increase of loss in the front side scroll not being fed during
inflow from the rear side, MHI checked internal flow on the rear and front cross
sections. Figure 5 shows loss distribution and flow distribution on cross sections of
the rear side where main stream is being fed and the front side where no stream is
being fed during feeding of the rear side (Figure 1 (ii) (a)). It is found that there is
leakage in the non-fed side which caused large loss. It is assumed that because a
twin scroll, where exhaust pulsation flows into the front and rear scrolls
intermittently, generates pressure difference between each scroll, in-leakage to the

473
lower pressure scroll occurs. The in-leakage flow into the lower pressure side flows
in the scroll for a while, and then enters into the rotor blade, making loss and
causing efficiency lowering.

Fig. 4 Internal flow of turbine during inflow from rear side (a)

Fig. 5 Internal flow of turbine scroll during inflow from rear side (a)

474
Figure 6 (ii) (a) shows internal flow of the turbine on a cross section at a constant
position in the circumferential direction during inflow from the rear side (Figure 1
(ii) (a)). It is found that leakage flow from the throat of the rear side scroll flows to
the shroud side of rotor blade, and then flow separation occurs on the hub side
inside the rotor blade. On the other hand, Figure 6 (ii) (b) shows internal flow of the
turbine on a cross section at a constant position in the circumferential direction
during inflow from the front side (Figure 1 (ii) (b)). Unlike inflow from the rear side,
no flow separation is found on the hub side inside the rotor blade during inflow from
the front side. Leakage flow from the throat of the front side scroll flows to the hub
side of rotor blade, and then flow separation occurs on the shroud side inside the
rotor blade. However, flow separation this time is smaller than that during inflow
from the rear side. Conventional analysis under inflow from both sides at a constant
mean pressure of exhaust pulsation could not find out internal flow loss, but it is
found through unsteady analysis that exhaust pulsation flowing into the front and
rear scrolls alternatively and intermittently causes flow separation loss in the rotor
blade and this loss leads to efficiency lowering as transferred to the rotor blade
outlet.

Fig. 6 Internal flow of turbine on a cross section at a constant position in


the circumferential direction

3.2 Improvement focusing attention on flow separation inside rotor blade


and leakage flow in scroll
Among internal flow phenomena described above, MHI has focused attention on
flow separation inside the rotor blade and leakage flow to the non-fed side in the
scroll to make improvements for a new twin scroll turbine. Figure 7 compares
leakage flow in the non-fed side during inflow from the rear side (Figure 1 (ii) (a)).

475
The conventional twin scroll turbine had leakage flow at the position around 90
degrees in the circumferential direction, but the new twin scroll can achieve its
reduction and improvement of flow phenomena.

Fig. 7 Comparison of leakage flow on cross section in front side between


conventional and new twin scrolls during inflow from rear side (a)

Next, Figure 8 compares internal flow of the turbine rotor blade between
conventional and twin scrolls during feeding of the rear side (Figure 1 (ii) (a)). As
shown in Figure 8 (i), flow distribution of the new twin scroll has a smaller low flow
rate area in the hub side than that of the conventional twin scroll, which results in
suppression of flow separation during inflow from the rear side. As shown in Figure
8 (ii), loss distribution on the 90% span cross section of the new twin scroll has a
smaller loss at the posterior border of the shroud than that of the conventional twin
scroll. Loss at the trailing edge of the shroud is originated in loss generated in the
rotor blade hub side. Therefore this reduction of loss means improvement of
internal separation flow of the rotor blade.

Figure 9 compares turbine efficiency. As shown in this comparison, the new twin
scroll has achieved 1.9% enhancement of mean in-cycle turbine efficiency mainly
due to improvement of efficiency during inflow from the rear side.

476
Fig. 8 Comparison of internal flow of turbine rotor blade between
conventional and new twin scrolls during inflow from rear side (a)

Fig. 9 Comparison of turbine performance between


conventional and new twin scrolls

4. MEASUREMENTS OF TURBINE PERFORMANCE

Performance tests of new twin-scroll turbine were conducted on gas-stand (Fig. 10).
Turbine diameter is 43mm and blade number is 11. Efficiency at Front flow and full
flow condition are slightly reduced. On the other hand, turbine performance is
improved at rear flow condition by 2%. This improvement will enhance pulsation

477
cycle average efficiency of turbine. Then deviation of turbine performance between
front flow condition and rear flow condition is reduced for new twin scroll turbine.
This characteristics also reduce engine performance deviation between cylinders.

Base-Fside Modified-Fside Base-All Modified-All


Base-Rside Modified-Rside

Turbine efficiency [-]


Turbine efficiency [-]

5% 5%

1.0 1.5 2.0 2.5 3.0 1.0 1.5 2.0 2.5 3.0
Turbine pressure ratio [-] Turbine pressure ratio [-]

(a) Partial admission (b) Full admission


Fig. 10 Gas-stand test results

5. FUTURE ISSUES AND OUTLOOK

MHI has obtained prospects of improving turbine efficiency through restraining


internal flow loss of the twin scroll turbine under conditions of exhaust pulsation. In
the future, MHI will conduct performance tests on an engine test bench to verify the
effect. Because internal flow of the turbine under conditions of exhaust pulsation is
highly complicated, it is necessary to improve accuracy of simulation by measuring
the unsteady experimental data that verifies unsteady computational simulation.
Also MHI will promote optimization of the shape of a twin scroll turbine for use with
a small engine, taking into account its productivity.

6. CONCLUSION

Responding to tightening of regulations regarding emission gas and fuel economy of


automobiles, MHI has developed a high-efficiency twin scroll turbine, using
unsteady computational simulation for efficiency enhancement of turbocharger. This
twin scroll turbine can raise boost pressure even at lower engine speed, increase
engine torque, and therefore improve fuel efficiency. It is predicted that tightening
of regulations regarding emission gas and fuel economy of engines will continue in
the future, and demand for higher efficiency of turbochargers seems to be endless.
MHI is willing to develop higher-efficiency turbochargers to improvement of
environmental performance in the future.

7. REFERENCES

(1) Ebisu, M. et al., Mitsubishi Turbocharger for Lower Pollution Cars, Mitsubishi
Heavy Industries Technical Review Vol. 41 No.1 (2004-1)
(2) DieselNet, Summary of worldwide engine and vehicle test cycles, 2012
http://www.dieselnet.com/standards/cycles/#eu
(3) DieselNet, Emission Test Cycles ECE +EUDC/NEDC, 2000
http://www.dieselnet.com/standards/cycles/ece_eudc.php
(4) DieselNet, Emission Test Cycles Common Artemis Driving Cycles (CADC), 2011
http://www.dieselnet.com/standards/cycles/artemis.php

478
AUTHOR INDEX

Aghaali, H ......................................................... 179


Akehurst, S .................................... 13, 27, 149, 163
Alsalihi, Z ........................................................... 65
An, B ................................................................. 79
Andrews, D N .................................................... 253
Ångström, H E ................................................... 179
Arnau, F J ......................................................... 103
Ashtekar, A ....................................................... 361

Baar, R ................................................................ 3
Banks, A ........................................................... 241
Bargende, M ...................................................... 301
Binder, E ........................................................... 41
Blacker, D ......................................................... 289
Bolz, H ............................................................. 265
Böttcher, L ......................................................... 41
Bou-Saïd, B ....................................................... 449
Brace, C J ...................................... 13, 27, 149, 163
Burke, R D ................................................ 103, 149
Burkinshaw, M ................................................... 289

Capon, G .......................................................... 149


Casey, M V ......................................................... 55
Chen, H ............................................................ 113
Christen, C ........................................................ 189
Codan, E ........................................................... 189
Copeland, C .................................................. 13, 27
Cornwell, R ....................................................... 241
Criddle, M ......................................................... 207

Davies, P .................................................. 137, 149


Dietrich, M ......................................................... 89
Dowell, P .......................................................... 149
Duda, T ............................................................ 149
Early, J .............................................................. 89
Eynon, P ........................................................... 399

Filsinger, D .......................................... 89, 301, 349


Friedrich, I ........................................................... 3

Garrett, S ......................................................... 163


Grabowska, D .................................................... 227
Grigoriadis, P ..................................................... 41
Groves, C .......................................................... 137
Gurunathan, B A ................................................. 13

Hagemann, T .................................................... 375


Harley, P X L ...................................................... 89
Hattori, H .......................................................... 389
Hazby, H R ......................................................... 55
Heyes, F ........................................................... 125
Hirai, Y ............................................................. 217
Hoffmann, R ...................................................... 437
Hoshi, T ............................................................ 471
House, T ........................................................... 227
Hu, B ................................................................ 27

Ibaraki, S ........................................................... 65
Ikeya, N ........................................................... 217

Jackson, R .................................................. 13, 207


Jarvis, S ........................................................... 207

Kadunic, S ........................................................... 3
Kasthuri Rangan, P S .......................................... 281
Kaufmann, A ..................................................... 265
Köhl, W ............................................................ 349
Kreschel, M ....................................................... 349

Lamquin, T ........................................................ 449


Lancaster, C ...................................................... 361
Lee, D .............................................................. 207
Lewis, A G J ....................................................... 13
Liebich, R .......................................................... 437
Liu, S T ............................................................. 411
Liu, Y H ............................................................ 411
Lotz, R .............................................................. 227
Luard, N ..................................................... 13, 207
Lüddecke, B ...................................................... 301

Marques, M ....................................................... 137


Martinez Botas, R F ............................... 13, 321, 333
Matteucci, L ....................................................... 13
Morand, N ......................................................... 137
Moscetti, J ........................................................ 227
Mrazek, R ......................................................... 137

Nanbu, T ............................................................ 79
Newman, P ....................................................... 207
Numakura, R ...................................................... 55

Okhuahesogie, O F ............................................. 125


Olmeda, P ......................................................... 103

Padzillah, M H .................................................... 333


Pandian, S ........................................................ 281
Parikh, S S ........................................................ 281
Porzig, D ........................................................... 421
Pronobis, T ........................................................ 437

Raetz, H ........................................................... 421


Rajoo, S ........................................................... 333
Ramamoorthy, J M ............................................. 281
Rémy, B ........................................................... 449
Reyes-Belmonte, M ............................................ 103
Richardson, S .................................................... 207
Riley, M J W ...................................................... 125
Rinaldi, A .......................................................... 265
Roach, P ........................................................... 125
Rochette, C ....................................................... 207
Romagnoli, A ............................................... 13, 321

Sakai, M ........................................................... 321


Sato, W ............................................................ 389
Scherer, F ............................................................ 3
Schwarz, J B ..................................................... 253
Schwarze, H .............................................. 375, 421
Scott, S ............................................................ 227
Sens, M ............................................................. 41
Seume, J R ....................................................... 421
Shi, X ............................................................... 411
Shoghi, K .......................................................... 461
Smith, L ........................................................... 163
Smith, T ........................................................... 207
Spence, S W T .................................................... 89
Stewart, J ......................................................... 125
Such, C ............................................................ 241
Sugimoto, K ....................................................... 65
Suzuki, T .......................................................... 217

Tamaki, H .......................................................... 55
Tang, H ............................................................ 163
Tian, L .............................................................. 361
Tomanec, F ....................................................... 137
Tomita, I ...................................................... 65, 79
Toussaint, L ...................................................... 137
Turner, J W G ................................................ 13, 27

Van den Braembussche, R .................................... 65


Vemula, R ......................................................... 227
Verstraete, T ...................................................... 65
Vetter, D ........................................................... 375
Vlachy, D .......................................................... 137

Wakashima, K ................................................... 471


Wan-Salim, W S-I ............................................... 13
Watson, J .......................................................... 227
Winterbone, D E ................................................ 113

Yamagata, A ..................................................... 389


Yang, C ............................................................ 411
Yang, M Y ......................................................... 333
Yokoyama, T ..................................................... 471
Yoshida, T ......................................................... 471

Zangeneh, M ..................................................... 399


Zatko, M ........................................................... 137
Zhang, J ........................................................... 399
Zhang, Q .......................................................... 149
Zhao, B ............................................................ 411
Zhuge, W L ....................................................... 333
Cool2Power - Increased petrol engine
power and efficiency through an AC
driven intercooling system
S Kadunic 1, F Scherer 2, R Baar 1, I Friedrich 3
1
TU Berlin, Germany
2
Bundesanstalt für Materialprüfung und -Forschung (BAM), Germany
3
IAV GmbH, Germany

ABSTRACT

The supercharging potential of SI engines in passenger cars is mainly limited by


self-ignition and turbo charger turbine temperature. A reduced charge air intake
temperature would allow for a higher boost ratio and engine load or an increase in
efficiency. The research in this paper shows experimental results using a
turbocharged 1.4l SI-engine to prove the concept of extreme charge air cooling
using a modified air condition system. Maximum torque was increased by up to
12.5%. Assuming the original full load limit, fuel consumption could be reduced by
9% in some operating points.

1 INTRODUCTION

The downsizing of combustion engines has greatly improved the fuel efficiency of
passenger vehicles in the common driving cycles like NEDC and FTP75. There is a
trend to further decrease engine displacements to help fulfil future legislations
regarding CO2-emissions and fuel consumption [1, 2].

At the same time it can be assumed that vehicle performance must not suffer for
marketing and sales reasons. Under the constraints of passenger safety, comfort
and cost, a reduction of vehicle mass will be limited. For those reasons, rated
engine power most likely will not be reduced. As a fact, the average rated output of
passenger vehicles sold in Germany increased by 42% from 1995 to 2012 [3]. To
achieve the demanded power levels with downsized engines, superchargers (mostly
turbochargers) are used and boost ratios are continuously increased.

Unfortunately, turbocharged spark-ignition engines optimized for partial load


operation show unfavourable properties at high load levels due to limits of engine
operation. The occurrence of engine knock demands a late centre of heat release
(COHR) [4]. This directly increases brake specific fuel consumption (BSFC) due to a
decrease in thermal efficiency. Additionally, late combustion increases exhaust gas
temperature. This leads to another restriction in the operation of these engines,
which is the thermal limit of the turbocharger turbine [5]. Particularly at high
engine loads and speeds, the exhaust gas temperature has to be decreased by
running rich air/fuel-ratios. The evaporation enthalpy of the surplus fuel is
employed to cool the turbine, which significantly decreases efficiency as this
unburned fuel cannot contribute to power generation [6].

_______________________________________
© The author(s) and/or their employer(s), 2014
3
A decrease in intake air temperature reduces the knock tendency at a given engine
load. This would permit earlier COHR which decreases BSFC directly and also
reduces exhaust temperatures. Even at a given COHR, a reduction of intake
temperature reduces the exhaust temperature at high engine loads. This effect can
be employed to reduce mixture enrichment to increase efficiency at points of
engine operation where turbocharger turbine temperature is of concern. As shown
in [7] at the same time a significant reduction in exhaust emissions can be
achieved.

Regular production vehicles are equipped with intercoolers which can cool the
intake air directly or indirectly via a liquid system with ambient air. Depending on
driving situations and the design properties of the system the intake air
temperature achieved is commonly between 5K and 40K above ambient. Further
lowering of the charge temperature can be achieved by turbo cooling or early
intake valve closing (“Miller Cycle”). Both methods require an increase in boost
ratio to achieve reference load. However, this creates additional requirements for
the turbocharger compressor and increases back pressure from the turbo charger
turbine. This can lead to an increased residual gas content which is contradictive to
the idea of avoiding engine knock. Also, costly refined supercharging systems with
two stage turbochargers often become necessary.

Air conditioning (AC) systems are standard equipment in passenger vehicles sold in
the EU at the time of this publication (2013). Such a system contains most
components required to create an ultra-low temperature intercooling system. Also,
they are designed to cool down the passenger compartment quickly after start-up
with the engine at idle. During regular driving there is abundant cooling capacity
available. The compressors of these devices consume power from the combustion
engine. Regardless, a benefit in engine output or fuel efficiency is expected (at least
in certain areas of engine operation). In this study, the concept of low temperature
charge air cooling is investigated employing a modified vehicle AC system with a
1.4litre turbocharged SI engine and vehicle simulation.

2 CONCEPT AND EXPERIMENTAL SETUP

The goal of the conducted experiments was to test the concept of low temperature
intercooling with limited efforts in parts and vehicle development. A low-
temperature intercooling system should use production parts from automobiles as
much as possible and operate under regular driving conditions while requiring
minimum changes to the vehicle. Basic preliminary considerations were undertaken
regarding the layout of the system. It was found to be energetically useful to
conventionally cool the compressed intake air as far as possible after turbocharger
compressor, before entering the low temperature cooling system.

The setup shown in Fig. 1 was chosen. Compressed air from the turbo charger
compressor passes a conventional intercooler (IC). For reproducibility, an air to
water heat exchanger was used as sometimes found in regular production vehicles.
The engine dyno environment provided sufficient fluid to cool the intake air to not
more than 40°C at all operating points. This represents the performance of a well-
designed conventional intercooling system in average driving conditions, regardless
of whether an air-to-air or air-to-water system is assumed. The authors consider
this a conservative assumption, as in real world driving intake temperatures can be
well above this temperature level [8].

4
Fig. 1: Test setup

After the conventional intercooler, the precooled intake air passes the extra
intercooler (ICe) for low temperature. The same type of air-liquid heat exchanger is
used here as for the IC. It is fed with a low temperature fluid (water with
antifreeze) of around -5°C to 5°C provided by the modified AC system. This setup
permitted air temperatures of 10°C and lower after ICe, while avoiding excessive
build-up of ice from condensed intake air water. The setup was designed in a
matter for water condensing in the ICe to be directly provided into the intake
runner with no opportunities for puddles to collect. After each low-temperature test,
the system was operated without ICe-activation until regular operating
temperatures were reached to relieve the intake runners from ice and water before
further experiments or shutdown.

The only modification of the regular engine-driven AC compressor regarded the


refrigerant volume flow control. It was made accessible to the dyno system in order
to control the cooling output. Refrigerant used was common R134a. Compressed
refrigerant passes a condenser, which is liquid cooled by the dyno environment.
This provides consistent boundary conditions, but in a passenger vehicle this would
be the car’s air cooled AC condenser. A common thermal expansion valve is used to
control evaporation temperature. The evaporator is of the refrigerant-to-liquid type
as commonly used in commercial cooling systems. Here the low temperature fluid is
cooled after taking up heat from the intake charge. An electric pump is employed to
circulate the low temperature fluid through the evaporator and the ICe.

All engine related parts like throttle body, intake and exhaust manifolds,
turbocharger, catalytic converter, etc., were taken from a production engine
without modification (except added measurement probes). The parameters of the
test engine are shown in Tab. 1.

It is a state-of-the-art production engine, common for European midsize vehicles at


the time of this publication (see [1]).

5
Tab. 1: Technical data of test engine

Test engine data

Spark ignition, homogenous direct injection,


Turbo charged
4 cylinders inline, DOHC 16 valve
Displacement: 1390 cm³
Nominal output: 90 kW
Peak torque: 200 Nm
Nominal output from 5000 to 5500 min-1

3 EXPERIMENTAL RESULTS

Three series of experiments were run. In each series, conventional engine


operation with regular intercooling was compared to operation with the low-
temperature intercooling device engaged. For the first series, the test engine was
simulated in two vehicles representative for midsize vans and light utility vehicles.
Points of engine operation were derived assuming those vehicles performing
stationary high speed motorway driving, like common on the German Autobahn.
During the second set of experiments, the test engine was operated at its
maximum nominal effective load at various engine speeds. The goal of those two
test runs was to find out if gains in thermal engine efficiency can offset the
additional power required to run the modified AC system for low temperature
intercooling.

For the final tests, areas of engine operation were selected where an increase in
output would be of high interest. This was considered the case at low engine speed
(low-speed torque) and in the area of maximum power output. The goal was to
spread the torque band further. Here, the authors investigated if the benefits from
the cooler intake charge would permit a higher effective engine output, regardless
of the power drawn to drive the AC compressor.

After the tests performed under the operating conditions mentioned, the engine
was disassembled to investigate potential damage from water. No abnormal wear
or damage was found.

3.1 Autobahn
In Tab. 2, the data of the sample vehicles are shown. In Fig. 2, resulting points of
operation for the engine are visualized assuming typical driving situations in high
speed traffic. In all situations, constant speed on even surfaces and no wind were
assumed.

Tab. 2: Vehicle data for Autobahn simulation

Van Delivery
Front area 2,56 m² 2,8 m²
Drag coefficient 0,34 0,38
Gross vehicle weight 2165 kg 2180 kg

6
220

200

180 179 km⁄h


Brake torque in Nm
191 km⁄h
160
160 km⁄h
Delivery
140

140 km⁄h 160 km⁄h


120

100 120 km⁄h


140 km⁄h
Van

80
120 km⁄h

60
500 1000 1500 2000 2500 3000 3500 4000 4500 5000 5500 6000 6500
Engine speed in min-1
Fig. 2: Autobahn driving situations in engine map for van
and light delivery vehicle

Obviously, the heavier and less aerodynamic delivery vehicle creates larger engine
loads at given vehicle speeds. In Fig. 3 the results for engine operation with regular
and low temperature intercooling are shown.

In the case of the van, improvements from low temperature intercooling can only
be achieved at top speed. At all other vehicle speeds, the engine does not reach or
barely reaches its operating limits. The engine is operated at stochiometric air-fuel
ratio and COHR is just slightly later than thermally ideal. The additional energy
consumed by the AC compressor offsets any gains from earlier ignition timing. Only
at top speed operation does low temperature intercooling show promising results.
Here, the engine is operating in enriched mode with regular intercooling to
thermally protect the turbocharger turbine. With cooler intake air, the mixture can
be leaned out, which overcompensates the power needed to operate the modified
AC system. An improvement in BSFC of approx. 2.5% was achieved.

For the delivery vehicle, improvements in BSFC can be achieved at vehicle speeds
of 160km/h upwards. Due to the higher loads induced by the heavier and more
wind resistant vehicle, the engine is operating in enriched mode earlier than the
case in the van application. At top speed, BSFC can be reduced by approx. 3.6%.

Results from both vehicles show that serious improvements in BSFC can be
achieved as soon as the engine is required to be operated with enriched air-fuel
ratios for turbo charger protection with standard intercooling. In cases where colder
intake air “only” permits slightly advanced ignition timing there is a great chance
for any thermodynamic improvements to be overcompensated by power
consumption of the AC compressor of the low temperature intercooling system.
Essentially the heavier and bulkier the vehicle, the larger engine loads, which
increases the fuel saving potential of the low-temperature intercooling system.
Further down-speeding will enable benefits in BSFC at lower vehicle velocities. Also,
it should be considered that acceleration, head wind or uphill driving will put the
engine in enriched operation mode at lower vehicle velocities than simulated here.

7
40 40

30 30

TAir in °C

TAir in °C
20 20
Van Delivery
10 10

0 0

1500 1500
Boost pressure

Boost pressure
1300 1300
in mbar

in mbar
1100 1100

900 900

700 700
15 15
COHR in °CA

COHR in °CA
w/o AC Compressor
12 12
with AC Compressor

9 9

6 6
300 300
be in g/kWh

be in g/kWh
280 280

260 260

240 240
1.05 1.05

1.00 1.00
Lambda

Lambda
0.95 120 km⁄h 140 km⁄h 120 km⁄h 0.95
140 km⁄h

0.90 160 km⁄h 160 km⁄h 0.90

0.85 km⁄
0.85
191 h 179 km⁄h
0.80 0.80
3000 3500 4000 4500 5000 5500 3500 4000 4500 5000 5500
Engine speed in min-1 Engine speed in min-1
Fig. 3: Results during Autobahn driving

3.2 Increase of brake efficiency at full load


During this trial, the test engine was operated at its nominal full load throughout
the rpm-range. The approach was to search for an improvement in fuel
consumption at full load for all engine speeds through low-temperature
intercooling. Results are visualized in Fig. 4.

As seen in Fig. 4d, at low engine speeds (up to approximately 2000rpms) the air-
fuel ratio can be stoichiometric even with regular intercooling. However, Fig. 4g
shows that the engine has to be operated with very late COHR of up to 30° crank
angle after top dead centre (ATDC) instead of the desired 4° to 8° ATDC to prevent
engine knock. Employing the low-temperature intercooling system COHR can be
advanced by up to 5° CA. The resulting improvement in thermal efficiency
overcompensates the power consumed by the modified AC-system and an
improvement in BSFC of up to 3.8% can be achieved (see Fig. 4h). At engine
speeds over 2000rpm, the engine operates with an enriched mixture to limit
exhaust temperature for turbo charger turbine protection. A reduction in charge
temperatures lowers the temperature level during combustion, resulting in lower
exhaust temperatures too. Additionally, the reduced knock tendency and lower
pressure levels permit earlier COHR (see Fig. 4g), which provides a further

8
reduction in the exhaust temperature. Both effects combined allow reducing the
mixture enrichment while keeping the same exhaust temperatures, but low-
temperature intercooling does not permit stoichiometric mixture at those points of
engine operation. In points of operation with rich air-fuel ratio, engaging the
modified AC-system leads to a reduction in BSFC of up to 9%. Overall, during
nominal full load operation the low temperature intercooling system is
advantageous for BSFC at all engine speeds. At the same time, the air density
increase from temperature reduction lowers the required boost pressure to achieve
nominal full load at all engine speeds. This can be particularly advantageous for low
speed operation as turbo charger surge is less likely to occur.

40
Brake torque in Nm

200
∆T = 27.2 K 30

TAir in °C
190
20
180
w/o AC Compressor ∆T = 33.5 K
170 with AC Compressor 10
a b
160 0
20.0 1.05
NMEP in bar

19.0 1.00

Lambda
Auxillary load AC
18.0 Compressor
0.95

17.0 0.90
c d λ<1
16.0 0.85
950
90

TExh, Turbine in °C
pcyl, max in bar

80 Knock 900

70
850
60
e f
50 800
30 300

25 290
COHR in °CA

be in g/kWh

20 280

15 270

10 260
g h
5 250
1500 2500 3500 4500 1500 2500 3500 4500 5500
Engine speed in min-1 Engine speed in min-1
Fig. 4: Results at nominal full load

3.3 Gain in low end torque and rated power


Currently the common downsizing applications of turbocharged SI engines with
single stage turbochargers suffer from two contradicting demands. On the one hand
side, low end torque needs to be as high as possible to give the driver the
impression of a larger displacement naturally aspirated engine. This requires small
turbochargers to operate well at low gas mass flow. On the other hand side, rated
power needs to be on par with the larger reference engine, which demands decent
turbocharger operation at very high gas mass flow. This asks for large

9
turbochargers. As Fig. 5 shows, low temperature intercooling can ease the conflict
between those demands.

With an intake temperature reduction of about 30K, the brake torque at 1500rpm
was increased by 11% without increasing boost ratio. This could be achieved
through a density increase in the intake air and improved thermal efficiency by
earlier COHR. It is assumed that the torque increase can be achieved over the rpm-
range of the nominal max torque. At 1750rpm, it could even be increased by
12.5%. The nominal max torque range could be extended from 3500rpm to
4500rpm. None of these results needed turbocharger modifications. As shown in
[9], the availability of colder intake air also aids the dynamic behaviour of
downsized engines.

240 1800

Boost pressure in mbar


Brake torque in Nm

220
12.5 %
200 1600

180

160 1400

140

120 1200
30 950
900

TExh, Turbine in °C
25
COHR in °CA

850
20 800
with AC Compressor
15 w/o AC Compressor 750
700
10
650
5 600
1000 2000 3000 4000 5000 1000 2000 3000 4000 5000
Engine speed in min-1 Engine speed in min-1
Fig. 5: Results for improved output tests

4 SUMMARY OF RESULTS AND IMPROVEMENTS IN SYSTEM SETUP

Overall, it was found that low temperature intercooling employing a modified


vehicle AC system can improve efficiency by up to 9% at some high load points of
engine operation. This is particularly the case in situations where the engine is
operating with enriched air-fuel ratio for thermal component protection. If
stoichiometric combustion is possible with regular intercooling, the additional power
needed to drive the low temperature cooling system can be larger than the energy
saved by a more efficient COHR. In some of those cases, the system shows no
advantages. At full nominal engine load the system showed advantageous
properties regarding BSFC at all engine speeds. Additionally, boost pressures
required for nominal full load could be reduced.

It could be demonstrated that the concept can aid in increasing low-end torque of a
downsized turbocharged engine. Also, the torque band could be spread out further
and output could be increased. None of this needed modifications of the
turbocharger. No damage or abnormal wear was found in the engine after the test
trials.

10
In real world driving, charge air temperatures after intercooler can exceed 40°C,
particularly in hot surroundings or during high-load, low-speed driving. In those
cases, the introduced concept can show even larger potential. Also, the system
investigated was a provisional experimental setup. If the efficiency of the low
temperature intercooling system (cooling power to mechanical power consumption)
could be improved, benefits could already be gained at somewhat lower engine
loads. Potential exists when COHR is later than ideal for efficiency, due to the
danger of knock or cylinder pressure limits. For a real world application, the ICe
should be an evaporator core without an in-between-liquid-system as shown in Fig.
6.

Fig. 6: Improved system setup with intercooler / evaporator (ICev) [10]

The cold fluid circuit employed in the original experimental setup should be
eliminated. This would save cost, space and weight for additional parts and fluids.
The whole parts effort for such a low-temperature intercooling system would
consist of an evaporator-intercooler, an expansion valve and some lines. It would
also increase the efficiency of the cooling system significantly due to the elimination
of losses of one heat exchanger step. This would permit approximately 10°C charge
air temperature at R134a vaporization temperatures of around 0°C to 2°C. Those
temperatures are very similar to the vehicle’s AC system. This is, on the one hand
side, helpful to combine the systems and use synergies during parts development.
On the other hand side, the AC compressor’s efficiency is highly dependent on the
vaporization temperature and pressure. A higher vaporization temperature means a
better efficiency of the low-temperature cooling system. The system should be
optimized as a whole towards its use for intercooling. The experimental setup
consisted of parts available “off the shelf” without any specialized development. A
production system should benefit from an integrated design of the AC components
for passenger compartment AC and intercooling. Besides gains in efficiency, it is
also expected that pressure losses could be reduced in the intake manifold.

In general, the authors came to the conclusion that the basic experimental system
already showed promising potential for successful introduction into niches of the
automobile market. This potential could be increased with refined integrated
system design and development.

11
REFERENCE LIST

[1] KUBERCZYK, R.: Wirkungsgradunterschiede zwischen Otto- und


Dieselmotoren: Bewertung von wirkungsgradsteigernden Maßnahmen bei
Ottomotoren, Expert-Verlag, 2009
[2] GOLLOCH, R.: Downsizing Bei Verbrennungsmotoren: Ein Wirkungsvolles
Konzept zur Senkung des Kraftstoffverbrauchs, Springer-Verlag, 2005
[3] Dudenhöffer, F.: Wir müssen das Rennen um die höchste PS-Zahl stoppen,
http://www.zeit.de/auto/2012-10/dudenhoeffer-motor-ps, 2012
[4] WILLAND, J.: Grenzen des Downsizing bei Ottomotoren durch
Vorentflammungen, MTZ 05/2009 Jahrgang 70
[5] KADUNIC, S., SCHERER, F., BAAR, R., ZEGENHAGEN, T.: Heat2Cool-
Increased Gasoline Engine Efficiency due to Charge Air Cooling through an
exhaust Heat Driven Cooling System, MTZ 01/2014 (to be published)
[6] KADUNIC, S., BAAR, R., SCHERER, F., ZEGENHAGEN, T., ZIEGLER, F.:
Heat2Cool- Engine Operation at Charge Air Temperature below ambient
temperature, Aachener Colloquium 2013
[7] SCHERER, F.: Ladeluftkühlung durch Abgasenergienutzung - Ihr Einfluss auf
die Abgasemissionen, Dissertation TU Berlin 2014
[8] PUCHER, H.; ZINNER, K.: Aufladung von Verbrennungsmotoren, Springer
Vieweg, Berlin Heidelberg, 2012
[9] GUHR C.: Verbesserung von Effizienz und Dynamik eines hubraumkleinen
turboaufgeladenen 3-Zylinder-DI-Ottomotors durch Abgasrückführung und
ein neues Ladeluftkühlkonzept, Dissertation TU Dresden 2011
[10] KADUNIC, S.: Einfluss der Ladelufttemperatur auf den Motorbetrieb und ihr
Potenzial auf die Steigerung der Leistungsdichte des aufgeladenen
Ottomotors, Dissertation TU Berlin 2014

12
Assessment of supercharging boosting
component for heavily downsized gasoline
engines
A Romagnoli, W S-I Wan-Salim, B A Gurunathan, R F Martinez-Botas
Imperial College London, UK
J W G Turner, N Luard
Powertrain Research and Technology, Jaguar Land Rover Limited, UK
R Jackson, L Matteucci
Lotus Engineering, UK
C Copeland, S Akehurst, A G J Lewis, C J Brace
University of Bath, UK

ABSTRACT

Current trend on engine downsizing forces engine manufacturers to contemplate


powertrains with more than one boosting device. The presence of these devices
leads to complex 1-D engine models which rely on performance maps provided by
turbo/supercharger manufacturers. So far, no detailed analysis has been carried
out to understand how these maps affect engine performance simulation. As part of
the UltraBoost project (65% gasoline engine downsizing), Imperial College tested
the boosting components of a turbo-super configuration. The acquired data were
used to assess the effectiveness of 1-D engine performance prediction and to
contemplate the opportunity to exploit the boosting system and use it as engine
charge air cooler in the form of an expander.

NOMENCLATURE

Abbreviation Unit Subscripts Unit

BSFC Brake Specific Fuel OUT Outlet


Consumption INL Inlet
BMEP Brake Mean Effective T Total
Pressure S Static
C Flow velocity [m/s] SC Supercharger
cp Specific Heat TT Total-to-total
Constant Pressure [J/kg·K]
ER Expansion Ratio Greek
HP High pressure ρ Density [kg/m3]
LP Low pressure η Efficiency
m Mass Flow Rate [kg/s] γ Specific Heat Ratio
N Speed [rpm]
PR Pressure Ratio
P Pressure [Pa, bar]
T Temperature [K]
UB UltraBoost
W Power [W]

_______________________________________
© The author(s) and/or their employer(s), 2014
13
1 INTRODUCTION

The need for more efficient, more performing and less polluting engines is forcing
OEMs to significantly downsize engine capacity. The technical challenges associated
with engine downsizing (typically of up to 40%) are well understood and OEMs are
acting to develop engine concepts capable of real world fuel economy
improvements whilst delivering sufficient vehicle driveability and performance, for
example through turbocharging [1].

The UltraBoost programme is an ambitious project aiming to deliver a 2.0 Litre four
cylinder gasoline downsized demonstrator engine capable of up to 35 bar Brake
Mean Effective Pressure (BMEP), air pressure charging of up to 3.5 bar absolute and
offering approximately up to 35% potential for the reduction of fuel consumption
and CO2 emissions while still matching a JLR’s 5.0 litre V8 naturally aspirated (NA)
engine performance figures (65% engine downsizing).

The target power and torque curves 600 300

ENGINE TORQUE [Nm] llloll


are given in Fig. 1; these have
been used as baseline curves to 500 250

ENGINE POWER [kW]


build up all the analysis carried out
in this paper. In order to meet the 400 200
targets outlined above, a high level
of boosting for the engine is 300 150
required. After an assessment
phase, looking at many different 200 100
options (including electrification of
some components), a two-stage 100 50
series arrangement was selected.
Torque Power
This includes a LP stage (GT30R 0 0
turbocharger by Garrett) and a HP 0 1000 2000 3000 4000 5000 6000 7000
stage mechanical supercharger ENGINE SPEED (rpm)
from EATON (Fig. 2). More
specifically, the supercharger is a Figure 1: UB target Power and Torque
Roots type Twin Vortices Series
(EATON TVS R410) with a clutched, single speed drive. This supercharger is capable
of running with high adiabatic efficiency over a wide range of operating conditions.

The selection and matching of the boosting


system was mainly supported by 1-D engine
simulation (using GT-Power). Data for the
baseline engine platform where made
available by Jaguar Land Rover whereas for
the selection of the boosting components
(turbocharger and supercharger), the
analysis relied on the performance maps
provided by the manufacturers [3].

Engine models (1-D in nature) are being


extensively used by engine makers since
they provide simple, fast but still physically
based tools for preliminary engine design.
The theory behind engine software is not
new and the main advantage is that its
development relies on extensive database
which come from many years of
Figure 2: UB engine system experimental activity and model optimization.
layout [2] At present, the output prediction from

14
current models compares quite well with engine test bench data. However the level
of simplification intrinsically embedded in this software requires simulation
engineers to calibrate these models in order to compensate for the lack of data and
inadequate calculation routines in critical areas such as combustion, knock, surge,
turbine maps extrapolation etc. In addition to this, the rapid electrification of
powertrain also requires development of mathematical models for electric machines,
motors and generators which are not always available in standard software
packages.

Amongst the many areas which would require specific analysis, the current paper
intends to assess the impact on engine simulation output due to the presence of
boosting devices. In 1-D models, turbochargers and superchargers represent two
boundary conditions for the engine block since they determine the inlet pressure
and temperature of the charge air feeding the engine. However turbocharger
turbines present a long standing issue which is related with the width of turbine
performance maps provided by turbocharger manufacturers. These maps (usually
obtained with hot gas test stands) are narrow in range and force engine simulation
software to rely on significant extrapolated map points. In addition to this, the
advent of sequential/series installation on the intake side, forces the HP
compressors to operate at different inlet pressures and temperatures to what they
have been designed and tested for. Again this poses a question on how well the
HP compressors output parameters are predicted by current 1-D engine software.

2 AIM AND OBJECTIVES

The aim of this paper is to provide an insight into supercharging boosting


technology in heavily downsized gasoline engines. More in particular, two main
aspects of supercharging will be investigated:
- assessment of the impact of above-ambient inlet conditions on supercharger
performance prediction in current 1-D engine software. Justification: see
Section 1. Implementation: experimental analysis and comparison with
GT-Power prediction.
- assessment of supercharger performance when run as an expander (i.e. HP
stage declutched and driven by the pressurized air inlet instead of the engine
crankshaft). Justification: when not in use, the HP compressor could be
contemplated as an additional cooler for the engine charge air.
Implementation: experimental analysis.

3 EXPERIMENTAL ANALYSIS

3.1 Experimental set up


All of the Ultraboost engine development work was undertaken in one of the
University of Bath transient engine facilities. The facility features a twin
dynamometer arrangement with a 220kW AVL AC dynamometer supplemented by a
Froude eddy current brake to allow additional absorption up to the expected torque
and power rating of the Ultraboost engine. Maintaining the AC dynamometer
allowed motoring work to be undertaken for controller debugging and friction tests
as well as improved transient response during time to torque testing. All control
and data acquisition was performed using a Sierra-CP CADET V-14 control system.
In addition to standard temperatures and pressures, specific experimental hardware
was used for measuring engine speed, torque, combustion parameters, fuel flow,
emissions, intake manifold and engine blowby. Interfaces to the combustion
analysis, EMS, emissions analysers, CAN instruments, conditioning of combustion
air, fuel and cooling circuits were all achieved through the CADET v-14 software [4,
5].

15
Figure 3: Supercharger test rig overview

As per the supercharger analysis, a bespoke test facility was set up at Imperial
College with a 100 kW electric motor (nominal speed of 1800 rpm) used as main
drive. The speed of the motor is well below the nominal speed of the EATON
supercharger which is rated with a max. speed of 24000 rpm. Hence, in order to
achieve the required speed, a system of step-up gears was included in the rig
design (Fig. 3). The power absorbed by the supercharger was measured with a
torque meter positioned after the crankshaft pulley (Figs. 3 and 4). The pulleys/belt
assembly on the supercharger side replicates the same layout as the UB engine
using the same groove type pulley and belt (green box in Fig. 4). At the inlet to the
supercharger, mass flow rate, pressure and temperature were measured using a
V-cone (DP meter from ABLE), static pressure tappings (using a Scanivalve system)
and K-type thermocouples respectively. At the exit to the supercharger, only
pressure and temperature were measured (Fig. 4).

Figure 4: Supercharger testing layout

16
The inlet to the supercharger was built in modular units allowing two types of tests.
Referring to the black box in Fig. 4, Option 1 consists of letting the supercharger
inlet exposed to ambient conditions (atmospheric pressure and temperature)
whereas Option 2 allows the inlet pressure and temperature to the supercharger to
be set by means of an inline heater and compressor. Option 1 set up was used to
generate standard supercharger performance maps (refer to Section 3.2) whereas
Option 2 was used to generate test data for 1-D engine simulation comparison
(Section 4, above ambient inlet conditions) and to assess the supercharger
performance when run as an expander (Section 5).

3.2 Experimental results


In this section the test results for the UB engine and the HP stage (supercharger)
are reported.

The UB engine was tested at full load conditions for a range of speeds going from
1000 rpm to 6000 rpm. The engine set up included the boosting components (HP
and LP stage), EGR (Exhaust Gas Recirculation) control strategy and water cooled
exhaust manifold. In Fig. 5 the measured power and torque curves for the UB
engine (experimental Torque and experimental Power) have been compared with
those obtained from the baseline 5.0L engine (Target Torque and Target Power).
The figure shows that the UB engine meets the requirements for power and torque.
The full load curves at high engine rpm have not been completed due to some
issues which arose at the time of testing (even though a full sweep for power and
torque is currently being done and it will be material for future publications). At low
engine rpm instead (from 1000 rpm to 1250 rpm), the experimental torque shows
significant deficit when compared to baseline engine (28% and 12% less torque
and power respectively). This is due to the boost system not being able to deliver
the target pressure as initially predicted by the simulation during the selection
phase of the boosting components (for more details please refer to [3]).

600 300

500 250
ENGINE TORQUE [Nm]

400 200
ENGINE POWER [kW]

300 150

200 100
Target Torque
Exp. Torque
100 Target Power 50
Exp. Power
0 0
0 1000 2000 3000 4000 5000 6000 7000
ENGINE SPEED [rpm] )
Figure 5: UB Target and Experimental Power and Torque

The HP stage was tested and its results compared with those provided by the
manufacturer. The isentropic efficiency (total-to-total) was calculated using the
thermodynamic correlations in Eq. 7. Besides the efficiency, an additional
parameter (here defined as Power Ratio) was also calculated. This is defined as the
ratio between the consumed power of the supercharger (WSC, Eq. 6) and the input
power (WSHAFT, Eq. 5) provided by the electric motor (measured with the torque
meter). Since the pulleys/belt assembly after the torque meter replicates the same
layout as that of the UB engine, the Power Ratio is intended to represent a pseudo-

17
mechanical efficiency for the SP-EB-CB (green box in Fig. 3). The Power Ratio
values were then used to generate a look-up table for the UB engine model in order
to determine the transmission losses occurring between the crankshaft and the
supercharger.

= , + (3) = + 0.5 ∙ ∙ (4)


2∙
= ∙ (5) = ∙ , − , (6)

( ) −1 ,
= ℎ = (7)
, / , −1 ,

= (8)

The measured mass flow rate and efficiency results have been plotted in Figs. 6 & 7
and compared with those provided by EATON. From Fig. 6 it can be seen that the
mass flow rate values agree well with those measured by EATON. The trend for the
mass flow rate curves is consistent with the original maps from EATON and a
difference in mass flow no larger than 4% could be found. Different considerations
have to be made for the 18krpm and 20krpm where a larger discrepancy between
the mass flow rates (black and red points) was found. This can probably be
attributed to the higher temperatures experienced within the supercharger at
higher rotational speeds; this could lead to larger clearances between the
supercharger lobes and the casing (due to thermal expansion). This explanation
was partly supported by the supercharger manufacturer (EATON) which pointed out
the fact that the unit tested at Imperial College was a one-off prototype with
potentially different coating material for the lobes and hence with different response
to thermal variations (no further indication has been provided in this direction).

2.6
8krpm 10krpm 12krpm 14krpm 16krpm 18krpm 20krpm

6krpm
2.4

2.2
PRESSURE RATIO [TT]

4krpm
1.8

1.6

2krpm
1.4

1.2

1
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
MASS FLOW RATE [kg/s] EATON IMPERIAL

Figure 6: Validation for supercharger mass flow rate measurements

18
As per the supercharger efficiency, good agreement between the EATON and
Imperial supercharger test data was also found (Fig. 7), with thermal efficiencies
beyond 70%. In the secondary axis, the Power Ratio values have also been
reported (dashed lines). The curves show that for low rotational speeds (2krpm to
6krpm) the power ratio values fall below 75%. However for rotational speeds which
are more representative of the supercharger speeds in real engine operating
conditions the power ratio climbs up to 85%.The speed ratio between engine
crankshaft and supercharger speed is 5.88. This means that for supercharger
speeds lower than 6000 rpm, the engine would be running at speeds less than
1000 rpm, operational speeds at which the supercharger will not be operating.

100% 90%
14krpm 16krpm
18krpm
90% 12krpm 20krpm
85%
8krpm
80%
2krpm 80%
4krpm 6krpm
70%
75%
60%
EFFICIENCY [TT]

POWER RATIO
50% 70%

40%
65%
20krpm
30%
18krpm
60%
14krpm
20%
2krpm 12krpm
8krpm 55%
10% 4krpm
6krpm 10krpm 16krpm
0% 50%
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18
MASS FLOW RATE [kg/s] EATON IMPERIAL Power Ratio
Figure 7: Validation for supercharger efficiency measurements
(primary axis) and plot of the Power ratio (secondary axis)

4 COMPUTATIONAL ANALYSIS

The following paragraphs analyse the output of the GT-Power simulations for the HP
stage performance. This aim is to assess how well GT-Power simulation calculates
the performance parameters when the input values are different than ambient. The
GT-Power analysis was done using a validated 1-D engine model for the UB engine
(refer to Section 1). For validation, the results for the 1-D engine model was
compared to engine test bed results as described in Sections 1 and 3.2.

4.1 Supercharger performance analysis


The supercharger was tested for rotational speeds, inlet mass flows, pressures and
temperatures equivalent to those obtained from the UB engine at full load
conditions. The control strategy for the UB engine, forces the supercharger to
operate up to 3500 engine rpm ( 20580 supercharger rpm). However for the
current testing programme, it was chosen to not exceed 2500 engine rpm on the
test bed in order to avoid damaging the supercharger as a consequence of the large

19
inlet pressure and temperature. The comparison for the supercharger inlet
conditions between the 1-D simulation and the tests is given in Table 1. From the
table it is apparent that the speed, pressure, temperature and mass flow rate have
been set accordingly, with a discrepancy no larger than 1%.

Table 1: Supercharger inlet conditions (1-D simulation/Imperial Tests)

NEngine [rpm] 1250 1500 2000 2500


NSC [rpm] 7350/7359 8820/8844 11760/11750 14700/14742
Mass flow [kg/s] 0.060/0.057 0.0939/0.09 0.180/0.172 0.265/0.254
Inlet Pres. [Pa] 1.429/1.434 1.734/1.73 2.320/2.321 2.832/2.78
Inlet Temp. [K] 303/302 320/320 333/332 347/353

4.0 385
1-D Simulation
380

SC OUTLET TEMPERATURE [K]


3.5 Tests Imperial
SC OUTLET PRESSURE [bar]

375
3.0
370
2.5
365
2.0
360
1.5 1-D Simulation 355
Tests Imperial
1.0 350
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
ENGINE SPEED [rpm] ENGINE SPEED [rpm]

12 0.9
0.8
10
0.7
8 0.6
SC POWER [kW]

EFFICIENCY

0.5
6
0.4
4 0.3
0.2
2 1-D Simulation 1-D Simulation
Tests Imperial 0.1 Tests Imperial
0 0.0
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
ENGINE SPEED [rpm] ENGINE SPEED [rpm]
Figure 8: Comparison between measured and predicted SC data

In Fig. 8 the measured outlet conditions to the supercharger have been compared
with those obtained with the simulation (SC outlet pressure, SC outlet temperature,
SC power, SC efficiency). Even though the trend of all the measurements is well
captured by the simulation, the test results show that there are some discrepancies
between the measured and predicted data. At higher engine rpm, the measured
and predicted outlet pressures diverge significantly with a variation of about 6%
(∆P=0.21 bar). As per the temperature, there is no clear pattern between the
measured and predicted values. The same occurs for the SC power which shows
larger variations as the engine speed increases (at 2500 rpm, difference of about

20
30.7%, ∆Power=3.53 kW). As per the efficiency, the results are believed to be
strongly biased by heat transfer effects. Heat transfer effects on compression
process for centrifugal machines were explained by the lead authors in [6].
Calculating the supercharger isentropic efficiency (using Eq. 7) when inlet
temperature is higher than ambient is no longer appropriated, since heat transfer
through the casing affects the SC exit temperature (which could partly explain the
disagreement observed between the predicted and measured outlet temperatures).
A variation in outlet temperature of ±5ᵒC can lead to an efficiency variation of
several percentage points.

5 SUPERCHARGER AS AN EXPANDER

5.1 Turbo-expansion: initial simulation study


In pressure-charged engines, intake air pressure is raised by a compressor in order
to increase air density and therefore, specific power output. However, the charge
air temperature is also raised by the compression process, thus making intercooling
necessary to reject this heat to the environment. However, since heat exchangers
have effectiveness below unity, the temperature of the intake air will always be
above ambient temperature in a pressure charged engine. This increased air
temperature can lower the knock limit of the engine - making it necessary for lower
compression ratios or greater spark retard. Both of these mechanisms for knock
control can lead to compromises on fuel economy for a given BMEP target. While
there are other ways to improve knock tolerance in a pressure charged SI engine
(increasing EGR rates and improved port and combustion chamber design),
lowering charge temperature at start of compression could provide a direct route to
improved fuel economy.

The concept of turbo-expansion is not new – indeed it has been used widely to
provide air conditioning to commercial aircraft for many years. The idea is a simple
one. If there is energy available to over-boost an air charge, this additional
pressure can be expanded over a turbine (expander) to remove heat. Therefore, if
the heat of compression is removed via a standard intercooler prior to expansion, it
is possible to cool the air charge below ambient temperature. Due to the potential
advantages of this approach, there have been recent attempts to apply this concept
to the air path of a SI engine. Perhaps most notably, the NOMAD engine
development project by Lotus Engineering [7, 8] sought to demonstrate the
advantages of turbo-expansion using an Opcon mechanical twin-screw expander.
The concept and the T-S diagram from this work are shown in Fig. 9.

Figure 9: Turbo expansion process [7]

21
Whelan et. al. [9, 10] proposed a slightly different turbo-expansion concept where
a compressor and turbine are both placed in the intake flow path to raise and lower
the intake pressure respectively. By removing heat between the compressor and
turbine stages with a second charge air cooler, the expansion stage will lead to
sub-ambient temperatures. This approach was investigated using the UltraBoost,
GT-Power model as part of an initial study into the possible benefits of turbo-
expansion to a highly downsized SI engine. In the model arrangement the
additional components were placed within the two stage supercharger/turbocharger
air path.

This brief study considered a single engine speed at 2000rpm where the
supercharger supplied an over-boost (pressure ratio +0.25) compared to standard
operation. The turbo-expander compressor provided final pressure ratio of 1.25
before rejecting heat through a heat exchanger (ε=0.85) and expanding through
the fresh-air turbine (expansion ratio ~1.5). The turbo-expander performance was
provided from scaled turbocharger maps which generated total to static efficiencies
of ~75%. Despite the additional parasitic loss of the supercharger (MEP= +0.8bar),
this was more than offset by the benefit of being able to advance ignition timing
due to the lower charge temperature at the end of the compression stroke. In fact,
the full-load BSFC was predicted to improve by 4-5% due the sub-ambient air
charge from the turbo-expander. Although it was not explored in this study,
diminishing returns could be expected for reduced expansion efficiencies as noted
by the work by Turner et.al. [8].

5.2 Eaton Supercharger as expander


The GT-Power simulation suggested that there could be significant benefits from
turbo-expansion air charge cooling. In addition, since the UltraBoost arrangement
places the supercharger downstream of the turbocharger, this posed an interesting
question, namely, could the Eaton supercharger be used as an effective expander?
While this would require additional boost from the turbocharger, there are
operating points where there is excess turbine energy that could be used to
generate additional boost by modulating the wastegate. In this situation, a
mechanical connected expander could present an indirect means to recover exhaust
gas energy.

Since, to the authors’ knowledge, there is no available performance test data of an


Eaton supercharger acting as an expander, the experimental facility at Imperial
College was used to test this capability. As shown in Fig. 4, the inlet to the
supercharger was pressurized with a compressor and allowed to expand from inlet
to exit. The temperatures and pressures were recorded as before and the resultant
torque and speed was measured. Figure 10 shows the expansion ratio versus mass
flow for the Eaton device operating as an expander at constant speeds of 2000rpm,
4000rpm and 8000rpm. For reference, the supercharging characteristics are also
presented as delivering a ‘pressure ratio’ for speeds up to 10,000rpm. It can be
immediately noted that to deliver a similar mass flow range to the engine, the
supercharger must spin at a slower rotational speed if it is to operate as an
expander. As an example, if an expansion ratio of 1.4 is desired, in order to deliver
a similar engine mass flow, the supercharger would have to spin at approximately
half the speed compared to the operation as a compressor. While this dual purpose
is not possible with a fixed ratio drive, model-based research into the use of a
continuously variable transmission (CVT) supercharger has been carried out by the
University of Bath in reference [11]. Thus, the possible switching between
compressor and expander operations could be envisaged using such a variable ratio
drive system.

22
2.2 2.2

2.0 2.0

1.8 1.8

EXPANSION RATIO
PRESSURE RATIO

1.6 1.6

1.4 1.4

1.2 1.2

1.0 1.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10 0.11 0.12 0.13
MASS FLOW RATE [kg/s]
Imperial_4000rpm Imperial_6000rpm Imperial_8000rpm Imperial_10000rpm

Imperial_12000rpm SC_Expander_6000rpm SC_Expander_4000rpm SC_Expander_8000rpm

Figure 10: Pressure and Expansion Ratio of Eaton Supercharger versus


Mass Flow

Figure 11 demonstrates the total-to-total efficiency of the supercharger as an


expander compared to the compressor operation. The definition of expansion
efficiency is simply taken from a turbine definition as shown in Eq. 9. It is
immediately clear from this data that a substantial drop in isentropic efficiency is to
be expected from an attempt to use the EATON as an expander. This is not
surprising considering the design intent of the device was not to operate in this way.
The values are also reasonably consistent with those reported in reference [8].

1− , / , ,
= ⁄
ℎ = (9)
1 − (1⁄( ) ) ,

Finally, Fig. 12 demonstrates the power that can be generated, and thus returned
to the engine, if it is operated as an expander. Picking up the example earlier, if an
expansion ratio of 1.4 is desired, approximately 1kW of power can be generated
while simultaneously reducing the air charge temperature.

These results represent an intriguing new use of an Eaton supercharger and show
that the device could be used as an expander in the series, turbo-super
arrangement if it were possible to change the drive ratio. However, it is also clear
that, operating as an expander, the supercharger can only deliver isentropic
efficiencies between 40-50%. It is this rather poor expansion efficiency that may
lead to diminishing returns from a sub-ambient intake air charge. To understand
the trade-off between additional turbocharger boost (turbine backpressure) and the
benefits from turbo-expansion across a supercharger, further modelling work must
be carried out using the data presented here.

23
1.800 1.800

1.700 1.700

1.600 1.600
EXPANSION RATIO

PRESSURE RATIO
1.500 1.500

1.400 1.400

1.300 1.300

1.200 1.200

1.100 1.100

1.000 1.000
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80
EFFICIENCY [TT]
SC_Expander_4000rpm SC_Expander_6000rpm SC_Expander_8000rpm
Imperial_4000rpm Imperial_6000rpm Imperial_8000rpm

Figure 11: Total to total isentropic efficiency of the supercharger operating


as a compressor and expander

1.800

1.700

1.600

EXPANSION RATIO
1.500

1.400

1.300

1.200

1.100

1.000
-2000 -1500 -1000 -500 0
POWER SHAFT[W] SC_Expander_2000rpm SC_Expander_4000rpm
SC_Expander_6000rpm SC_Expander_8000rpm

Figure 12: Power harvested from the supercharger operating as an


expander

6 CONCLUSIONS

The current article describes the outcomes of a research conducted on a heavily


downsized gasoline engine (UltraBoost project, 65% engine downsizing). The
engine boosting system layout is a two-stage series layout, with a LP stage
(turbocharger) and a HP stage (EATON TVS supercharger). The analysis focussed
on the HP stage and is aimed at understanding the effectiveness of commercial 1-D
engine software when predicting the performance of boosting elements (a
supercharger in this case) with inlet conditions other than ambient (pressure and
temperature beyond atmospheric).

24
The EATON TVS supercharger was tested with similar speed, inlet pressure and
temperature as those predicted by the engine model at the inlet to the
supercharger. The operating conditions considered for the supercharger analysis
were equivalent to the engine running at full load conditions with speeds varying
from 1000 rpm to 2500 rpm. An agreement within 1% could be achieved for the
inlet conditions between simulation and experiments. As per the supercharger
output values, the test results showed that the 1-D engine software provides a
good prediction for the supercharger outlet pressure (max. discrepancy of 6%
encountered at 2500 rpm only). As per the outlet temperature and supercharger
power, the simulation and experiments show similar trend even though, the
difference between measured and predicted values can be significant for some
engine rpm. For instance the measured supercharger power was found to be 30.7%
higher than that predicted at 2500 rpm. As per the supercharger efficiency, no valid
comparison could be brought forward since the measured efficiency was found to
be significantly biased by heat transfer due to the heated air inlet temperature.

In addition to the supercharger simulation prediction analysis, an experimental


study on the supercharger performance was also carried out in order to assess the
opportunity to use the supercharger as an additional charger air cooler (i.e.
supercharger was run as an expander). The experiments showed that there is a
substantial drop in the supercharger isentropic efficiency when used in this way. A
penalty of more than 20 percentage points could be found (efficiencies less than
50%) for the speeds under exam (supercharger speeds 4000 rpm to 8000 rpm). On
the other hand, the test results also showed that a non-negligible amount of power
(more than 1kW) can be generated from the supercharger when run as an
expander. This suggests that, besides serving as a charge air cooler, the
supercharger could also be contemplated as a power generation system. These
results are quite interesting and could potentially open the way to consider a
supercharger as an enabler for enhanced engine fuel economy.

ACKNOWLEDGEMENTS

The authors would like to thank the UK Technology Strategy Board (TSB) for
funding the UltraBoost project and all the consortium partners (Jaguar Land Rover,
Lotus Engineering, Shell Global Solutions, GE Precision Engineering, CD-Adapco,
University of Bath and University of Leeds) for the support provided.

REFERENCE LIST

[1] Improving Fuel Economy by 35% through combined Turbo and Supercharging
on a Spark Ignition Engine, C. Salamon, M. Mc Allister, R. Robinson, S.
Richardson, R. Martinez-Botas, A. Romagnoli, C. Copeland, J. Turner, 21st
Aachen Colloquium Automobile and Engine Technology, 2012.
[2] UB200 Boosting Layout, C. Carey, JLR Technical Presentation, Jaguar Land
Rover, Coventry, UK, June 2011.
[3] Copeland, C., Martinez-Botas, R., Turner, J., Pearson, R., Luard, N., Carey, C.,
Richardson, S., Di Martino, P., Chobola, P., 2012. Boost System Selection for
Heavily Downsized Spark Ignition Prototype Engine, 10th IMechE International
Conference on turbochargers and turbocharging, London.
[4] Karl. G., Andrew, L., Sam, A., Chris, B., nick, L., 2012. The effect of advanced
combustion control features on the performance of a highly downsized
gasoline engine, FISITA 2012 World Automotive Congress, Beijing, China,
F2012-A01-021.
[5] Online source, University of Bath website: http://www.bath.ac.uk – Powertrain
Vehicle & Research Centre (PVRC).

25
[6] Romagnoli A, Martinez-Botas RF, 2012, Heat transfer analysis in a
turbocharger turbine: An experimental and computational evaluation, Applied
Thermal Engineering, Vol:38, 1359-4311, Pages:58-77.
[7] Turner, J., Pearson, R., Bassett, M., and Oscarsson, J., “Performance and Fuel
Economy Enhancement of Pressure Charged SI Engines through
Turboexpansion – An Initial Study”. SAE paper no. 2003-01-0401.
[8] Turner, J., Pearson, R., Bassett, M., and Blundell, D. W., “The Turboexpansion
Concept - Initial Dynamometer Results”. SAE paper no. 2005-01-1853.
[9] Whelans, C., Richards, R., “Turbo-cooling applied to light duty vehicle
engines”, Congress, 2005.
[10] Whelans, C., Richards, R., Spence, S., Young, A., “Design and Development of
a Turbo-expander for Charge Air Cooling” IMECHE Turbochargers and
Turbocharging, 2010.
[11] Rose, A., Akehurst S., and Brace, C., “Investigation into the trade-off between
the part-load fuel efficiency and the transient response for a highly boosted
downsized gasoline engine with a supercharger driven through a continuously
variable transmission” Proc. of IMECHE, Part D: Journal of Auto. Eng., Sept.
2013, DOI: 10.1177/0954407013492932.

26
The effect of divided exhaust period
for improved performance in a highly
downsized turbocharged gasoline engine
B Hu1, C Brace1, S Akehurst1, C Copeland1, J W G Turner2
1
University of Bath, UK
2
Jaguar Land Rover Limited, UK

ABSTRACT

Cylinder scavenging, negative PMEP and knock sensitivity at high engine speeds are
some of the main challenges facing downsized turbocharged SI engines, while
torque deficiency is more of a problem related to engines under low speed
conditions. Traditional downsized engine can currently only address the above
problems within a very small window inevitably leaving the other situation
compromised which would either affect fuel efficiency or engine power. In order to
simultaneously enhance the already competitive advantages while mitigating
inherent deficiencies of turbocharged engines, some novel techniques need to be
investigated.

Divided Exhaust Period (DEP) is such a gas exchange concept where two exhaust
ports from each cylinder are separated into different manifolds. The blow-down
pulse is directed through one valve that leads to the turbocharger in order to boost
the intake charge, while the other valve path (termed scavenge valve) bypasses
the turbine to scavenge the remainder of the exhaust. By combining the
characteristics of the downsized turbocharged engine and the scavenging process of
a more-normal naturally aspirated engine, both the torque performance and the
gas exchange process could be enhanced.

In this paper, the DEP concept will be investigated in simulation using a validated
highly downsized 2.0 Litre SI engine model. The final results showed that the BMEP
& transient performance, BSFC and the stability of the engine were all improved
due to the fact that the DEP concept features better gas exchange process and
improved combustion efficiency.

ABBREVIATIONS

DEP: Divided Exhaust Period BSFC: Brake Specific Fuel Consumption


SI: Spark Ignition PMEP: Pumping Mean Effective Pressure
BMEP: Brake Mean Effective Pressure RGF: Residual Gas Fraction
GA: Genetic Algorithm

___________________________________________
© The author(s) and/or their employer(s), 2014
27
1 INTRODUCTION

A general trend to downsize engines means more and more SI engines feature
turbochargers to improve fuel efficiency and BMEP performance. However, there
are several obstacles such as cylinder residual gas scavenging, negative PMEP due
to turbine induced back pressure, knock sensitivity and poor transient performance
that are challenges facing this technology [1]. Conventional approaches to
addressing these issues are to improve the engine side or turbocharger side
efficiency. This paper however focuses on a novel means of gas exchange process
utilising Divided Exhaust Period (DEP) to separate the exhaust process into two
distinct stages of blow-down and scavenge.

2 TECHNICAL CONCEPTS

The Divided Exhaust Period (DEP) concept is a novel way of accomplishing the gas
exchange process in turbocharged engines with the aim to combine the positive
effects of using a turbocharger while removing some of the negative aspects by
using a Variable Valve Timing (VVT) system on the exhaust system. Figure 1 shows
the schematic view of the DEP concept. Unlike the standard layout of the engine
exhaust system, two valves are separated to feed the turbine and the exhaust pipe
respectively. The blow-down valve functions just like the standard exhaust valve
discharging the first portion of exhaust gas into the turbine to boost the intake air.
While the scavenge valve which evacuates the latter exhaust portion directly into
the exhaust pipe (lower backpressure) aims at better gas exchange. By combining
the characteristics of standard turbocharged and naturally aspirated engines, it is
anticipated that the gas exchange process could be improved while keeping or
exceeding the required BMEP. Specific reasons from the perspective of better BSFC
and enhanced BMEP are as follows:

BSFC benefit
As the high backpressure between the turbine inlet and the exhaust port is isolated
from the engine piston during the latter portion of the exhaust stroke, an increased
PMEP could be attained. For fixed BMEP condition, increased PMEP indicates less
Gross IMEP thus better BSFC.

Meanwhile, by bypassing the turbine a positive pressure gradient across the intake
and exhaust system could be achieved. This is beneficial for the cylinder scavenging
process resulting in a large decrease for Residual Gas Fraction (RGF). It is known
that RGF is strongly related to the knock intensity and engine test variability. Some
author also stated that at knock-limited points, less RGF would cause the engine
less prone to knock which then would lead to the advance of the spark timing [2].
Hence, the DEP concept could also benefit from improved combustion efficiency.
Improved combustion phasing from reduced RGF also results in reduced exhaust
gas temperature and thus may reduce the need for other exhaust gas temperature
control actions such as over-fuelling, the adoption of water cooled exhaust
manifolds or increased exhaust gas recirculation (EGR).

BMEP benefit
Enhanced BMEP could be achieved for the DEP concept engine if power density is a
more important concern. This is due to the fact that DEP concept engine generally
features a smaller turbocharger and for most of operating points, it is also
attributed to the improved PMEP [3][4][5][6][7][8]. It should be noted that better
gas exchange process and improved combustion phasing are the two major reasons
to keep knock intensity under appropriate constraints for higher BMEP target.

28
Figure 1: Schematic view of the exhaust system layout in DEP concept

3 METHODOLOGY AND CONTROL STRATEGY

In this paper, seven degrees of freedom in the actuation of the DEP concept engine
including the timing of the intake, blow-down and scavenge valves and the lift &
duration of the blow-down and scavenge valves were investigated. In order to
understand the underlying gas exchange process, the timing of the scavenge valve
(one degree of freedom mechanism) effect on BSFC and BMEP were investigated
first. Optimization of the fully flexible mechanism was then undertaken and
presented later in the paper (only for BSFC optimization situation). For one degree
of freedom study, the profile of the intake and the exhaust valves were retained to
investigate the BSFC or BMEP improvement potential with minimum change of the
whole system.

12.00
10.00
Valve Lift (mm)

8.00
6.00
4.00
2.00 A
B
0.00
90.0 180.0 270.0 360.0 450.0 540.0 630.0
Crank Angle [aTDC]

Blowdown Scavenging Intake

Figure 2: Exhaust valve timing

It should be noted that in this paper, the timing of the scavenge valve is also the
control parameter to maintain or achieve the target BMEP, the function similar to a
traditional wastegate. The value of the other parameters will be based on the
optimization for different purposes and will form a map in the ECU, which is more
like a traditional calibration. The reason for selecting the timing of scavenge valve
to be the control parameter is due to the following reasons:

29
 Variable valve timing (VVT) mechanism is more easily manufactured than
variable valve lift and variable valve duration mechanisms.
 The scavenge valve is subjected to less thermal and pressure shock than
the blow-down valve

One degree of freedom mechanism


By adjusting the timing of the scavenge valve, the relationship between the blow-
down and the scavenge valve varies to feed a specific amount of the exhaust mass
flow into the turbine. Meanwhile, the overlap between the scavenge valve and the
intake valve accordingly alters which influences the amount of blow-through or
short circuit air flow. Overlap A and overlap B as can be seen in Figure 2 is
necessary to avoid choked flow and to control the blow-through/or back flow
respectively.

The operation strategy depends on the requirements of the system*. If fuel


efficiency is pursued, the scavenge valve should be set to a maximum advance
that attains BMEP target. By advancing the scavenge valve, less pumping work is
needed thus fuel efficiency is increased. This strategy can also benefit from less
blow-through which leads to less fuel injected. However, because of RGF doesn’t
improve significantly as in the situation when the timing of the scavenge valve
retards, the benefit from the combustion stability is small.

If achievable BMEP is preferred, then the scavenge valve generally should be


retarded. To be more specific, the profile of the scavenge valve needs to be shifted
right to allow more overlap as shown by B in figure 2. This would accordingly cause
more mass flow into the turbine which increases BMEP and less RGF due to the
larger amount of blow through, but the corresponding anchor angle (CA50) retards
and the larger amount of fuel mass flow rate would adversely affect the fuel
economy.

*This strategy only applies to the high and some of the low to medium conditions. Strategy at the low speed is
detailed in the RESULTS.

Multiple degrees of freedom mechanism


Due to the possible large blow through and the limitation of only one controllable
degree of freedom, the results might not be optimized as can be seen in the
following section. Fully flexible mechanisms of the intake and exhaust valves are
therefore necessary to be introduced to exploit the maximum potential of the DEP
concept engine.

As all the degrees of freedom are strongly interrelated, changing one would cause
the others to change their action on the engine system. Some mathematic
approaches for optimizing multi-variables systems and multi-objective responses
thus need to be investigated. A genetic algorithm has the capability to mimic the
process of natural evolution and is thus considered to be one of the most useful
approaches to solve this kind of problem. The schematic diagram for optimizing the
results is shown below. To be more specific, each degree of freedom represents a
gene in the chromosome and BSFC forms the fitness function. More detail about the
structure of the GA system can be referred from figure 3.

It should be noted that potential PMEP gains due to the implementation of the DEP
concept could offset some of the turbocharger power loss due to the smaller
amount of exhaust mass flow into the turbine. However, for some operating points,
the initial BMEP or desired higher BMEP is still not attainable. Thus, a smaller sized
turbocharger may need to be selected to optimize the whole system. The matching
of the IC engine and the turbocharger was performed by scaling the mass flow axis
of the standard engine’s turbocharger maps. Without considerable effort, a smaller
sized turbocharger was selected by using mass multiplier in GT-Power [9].

30
Figure 3: Genetic Algorithm Procedure

4 RESULTS

This study was carried out on a validated 2 Litre extremely downsized SI engine
model using GT-Power. It is worth mentioning that the target BMEP of this engine
can be in excess of 32bar and it is anticipated that the gas exchange process is
potentially significantly different to that of the engines other literatures investigated
given their lower BMEPs. The diameter of the valves was increased by 4mm to
avoid choking when only one valve is involved in the gas discharge process. The
wastegate control strategy was replaced by the timing of the scavenge valve to
maintain or exceed BMEP for strategy 1 and 2 below. The Douand & Eyzat knock
model was implemented in GT-Power in an effort to predict knock onset for
individual cycles [10].

In this paper, firstly we aimed at minimizing the fuel consumption while matching
the BMEP performance of the base engine. After that, the potential of exceeding the
original engine performance was investigated. It should be noted that the
comparable results are from the original engine model in WOT condition. For
strategy 1, one degree of freedom effects were presented first with detailed
explanation. Then multiple degrees of freedom interactions were presented with
some explanations of the process occurring. For strategy 2, only one degree of
freedom effect on BMEP was investigated.

Strategy 1 for BSFC purpose


In this strategy, the aim is to reduce fuel consumption, residual gas content and to
improve boost control. Steady state operating points of 6500RPM/26.2bar,
4500RPM/30.9bar, 2500RPM/17.9bar and 1500RPM/12.7bar were presented. The
reason why these four sets were chosen is because they represented four different
gas exchange process detailed in the following.

31
1) One degree of freedom mechanism

Table 1: Averaged results for 6500RPM


6500RPM/26.2bar:

BMEP (bar) BSFC (g/kW*h) RGF CA50 PMEP (bar) Turbine Power
Original 26.2 258.7 2.74 17.3 -2.2 43.9
DEP 26.2 254.3 1.75 17.5 -1.2 39.2

From table 1, it can be seen that the BSFC may be reduced by 1.72% while keeping
BMEP fixed. This was largely due to the improvement of the pumping work as
negative PMEP was increased by 44.7%. The combustion phasing on the other hand
didn’t change too much which did not correlate with the reduction of RGF [9].
Possible reason could be poorer choking condition when only one exhaust valve is
involved which adversely affects the temperature and pressure in cylinders.
Invalidated knock model could also influence the relationship between knock
sensitivity and combustion phasing.

Lower pumping work can also be implied from the reduced turbine power, DEP
concept engine only needed39.2kW power to boost the intake pressure in order to
attain the required BMEP while the base engine required as large as 43.9kW, which
indicated that more power was wasted to overcome the pumping resistance for the
baseline engine. As a wastegate was not required for boost control, the saved
pumping loss during the wastegate open also contributed the improved PMEP.

Table 2: Averaged results for 4500RPM


4500RPM/30.9bar:

BMEP (bar) BSFC (g/kW*h) RGF Anchor Angle PMEP (bar) Turbine Power
Original 30.9 249.0 3.49 23.3 -0.450 35.0
DEP 30.9 246.7 0.94 19.0 0.074 29.9

At 4500RPM, as can be seen in table 2, BSFC dropped a lot from 246.7 to 242.9.
Unlike 6500RPM situation, both the reduction of the pumping work and the
improvement for combustion efficiency contributed the decrease of BSFC. The
spark-timing advanced by about 4 degrees from 23.3 to 19.0 which contributed
more than the improvement of PMEP although it was not verified quantitively. This
correlated with RGF very well with RGF decreased from 3.49 to 0.94. It is also
anticipated that the DEP concept engine would also benefit from more stability due
to the significant RGF decrease. This would further improve engine efficiency [3].

BSFC at 2500RPM and 1500RPM did not change too much which indicated that the
DEP concept did not have significant BSFC benefit within lower engine speed. At
1500RPM, the increased BSFC was attributed to the decrease of the pumping work.
It should be noted here, the original model is equipped with a supercharger as well
but in order to investigate the DEP concept on turbocharged SI engines, the
supercharger was removed.

32
Table 3: Averaged results for 2500RPM and 1500RPM
2500RPM/17.9bar:

BMEP (bar) BSFC (g/kW*h) RGF Anchor Angle PMEP (bar) Turbine Power
Original 18.1 239.4 4.36 14.7 -0.105 5.6
DEP 18.2 239.4 3.23 13.3 -0.036 5.2

1500RPM/12.7bar:

BMEP (bar) BSFC (g/kW*h) RGF Anchor Angle PMEP (bar) Turbine Power
Original 12.7 246.2 5.14 11.4 0.045 0.9
DEP 12.7 247.3 4.31 10.9 0.006 0.7

Pumping Loop
Fuel consumption reduction between the standard and the DEP model can also be
presented using the pumping loop in P-V diagram below.

Figure 4: Pumping loop in P-V diagram at 4500RPM

It can be seen from figure 4 that for the end of the exhaust stroke, the original
engine continued to push the remaining exhaust gas from the cylinder to the
turbine which increased the backpressure further and as the exhaust valve started
to close there was a large peak in the cylinder pressure. However for DEP concept
engine, as the blow-down valve was closed while the scavenge valve started to
open, the cylinder pressure dropped dramatically to nearly atmospheric level which
removed the backpressure to the piston. Pumping work can also be implied from
the enveloped area, it is clear that the DEP concept benefit from less pumping loss
than the original model.

33
From the results above, it can be summarized that for fixed BMEP condition, BSFC
improvement depends on the trade-off between the benefits of the reduction in
pumping work and the combustion efficiency gain.

2) Multiple degrees of freedom mechanism

Table 4: Averaged results for multiple degrees of freedom mechanism

6500RPM 4500RPM 2500RPM 1500RPM


Original DEP Original DEP Original DEP Original DEP
BMEP (bar) 26.2 26.2 30.9 30.9 18.1 18.1 12.7 12.7
BSFC (g/kW*h) 258.7 252.9 249.0 239.3 239.4 235.6 246.2 245.6
RGF 2.74 1.19 3.49 1.50 4.36 2.72 5.14 4.93
Anchor Angle 17.3 16.2 23.3 19.3 14.7 11.4 11.4 12.25
PMEP (bar) -2.2 -1.2 -0.450 0.26 -0.105 0.156 0.045 0.062
Turbine Power (kW) 43.9 38.1 35.0 29.0 5.6 5.0 0.9 0.85

Table 4 shows the averaged results for multiple degrees of freedom mechanism.
Due to the improved gas exchange process and better combustion phasing, the
BSFC performance for all speed range was improved with large decrease of BSFC
occurred around the high speed.

However, due to the complexity of utilizing seven degrees of freedom mechanism


for series production engine, it is almost impossible to achieve such improvement.
Engines equipped with DEP concept in the future is most likely to be in the between
of the one degree freedom and the seven degree freedoms, thus the behaviour or
the performance of the engine should also be in the between.

Strategy 2 for BMEP purpose


This strategy targets higher BMEP, thus the BMEP curve is more important than the
corresponding BSFC. Only the results of 4500RPM and 1500RPM were presented for
simplicity purpose.

Figure 5 shows the relationship between the scavenge valve timing and BMEP &
BSFC & RGF & PMEP. The upper two plots indicated maximum BMEP potential at
higher engine speed whilst the lower representing the lower speed BMEP trend. The
performance of both situations improved by around 10% maximally. But it should
be noted that even though BMEP trend is similar for these two sets of speeds and
loads the curve of BSFC & RGF & PMEP is totally different.

The reason why the BSFC trend in fixed BMEP was not the same is attributed to the
fact that at high speed with increased scavenge valve timing, there was larger blow
through across the cylinder. However, at low speed as the overlap between the
scavenge valve and the intake valve was relatively small, blow through effect was
not noticeable. The decreased trend of BSFC for lower speed was basically due to
the gain of PMEP, which can be seen from the last plot of figure 5.

34
Figure 5: BMEP, BSFC, RGF and PMEP for 4500RPM and 1500RPM

RGF curve difference can be explained by the following points:

 Timing relationship difference between the intakes, blow-down and


scavenge valve
 Pressure difference across the cylinder

As with increased scavenge timing, the overlap between the scavenge valve and
the intake valve is increased. Two different gas exchange process could occur here.
When the intake pressure is larger than the exhaust pressure when the scavenge
valve opens, the blow through effect causes RGF decrease dramatically which is the
case for higher speeds. However when the intake pressure is smaller as at the
lower speed, back flow effect would dominate [11].

In order to explore the gas exchange process in lower speed range, the following
figure was drawn. For comparative purposes, only the scavenge valve timing of 270
and 310 degree representing the lowest and largest RGF were selected.

Figure 6 indicates the backflow for larger scavenge and intake valve overlap (lower
blow-down and scavenge overlap). This can explain the RGF trend for lower engine
speed well. It can therefore be concluded RGF for DEP concept model not only
depends on the blow through but also on the possible back flow. For higher speed
the first effect contributes more while for lower speed the latter one dominates.

35
Large Back Flow

Figure 6: Gas exchange process for different scavenge valve


timing at 1500RPM

Strategy 3 for transient performance purpose


Transient performance is one of the major challenges to address in order to see
widespread adoption of turbocharging in most production engines. Although further
research work in this area is needed, it is expected that DEP concept could benefit
for transient performance of turbocharged engine. This is majorly due to the fact
that smaller inertia of the modified turbocharger is adopted.

36
5 DISCUSSIONS

It should be noted that even though the base engine model is validated using
prototype engine testing data, the modified model for the DEP concept does not
calibrate. The potential of this novel theory could be better proved with some
testing results. In addition to that, the knock model for running the simulation
which decides the anchor angle (CA50) might not be suitable for the DEP concept,
which would be verified in the future.

Higher blow-down exhaust temperature and choking are the two major drawbacks
for DEP concept engines. As the majority of high enthalpy mass flow is evacuated
during the first portion of the exhaust stroke and a relatively low enthalpy mass
flow which directly links to the exhaust doesn’t balance some of the blow-down
pulse temperature. Higher blow-down but lower scavenge pulse temperature should
be anticipated. It should be noted that for this study, the limit of the inlet turbine
temperature was not considered. High temperature-resistant material should be
used or fuel-enrichment strategy needs to be involved. This would either increase
the cost of the system or affect the fuel efficiency. Figures 7 shows the comparison
of the original, blow-down and scavenge exhaust temperature when one degree of
freedom involved in fixed BMEP condition.

1500

1000
Temperature (K)

Original
Blow-down
500
Scavenge

0
6500 4500 2500 1500
Engine Speed (RPM)

Figure 7: The comparison of the original, blow-down and


scavenge exhaust temperature

From figure 8, it can be seen that even though the diameter of the exhaust valves
were increased, choking condition still existed and even poorer than the original
engine. This is due to the fact that during some of the exhaust stroke and intake
stroke, only one of the exhaust valves was involved. This can also partly explain
why PMEP drops for higher speed range when the timing of the scavenge valve
retards. The choking condition could be largely improved by optimization of the
timing, lift and duration of the exhaust valves.

37
Figure 8: Mach number for Original and DEP concept engine at 6500RPM

6 CONCLUSIONS

The general trend under fixed BMEP condition is that at lower engine speed, the
backpressure is low and there is less to gain from the DEP concept whilst at higher
speeds the backpressure is higher, the potential to reduce this backpressure is
higher. Thus the BSFC reduction is more noticeable at higher speeds.

Using the DEP concept for entire speed range due to the better cylinder scavenging
can increase BMEP. However, the trend for RGF under high and low speed range is
different. Blow through is dominating for high speed operating point while backflow
effect is large enough for low speed situations.

Transient performance for the DEP concept engine is considered to be better than
the baseline engine. It is attributed to the PMEP reductions and more importantly
due to the smaller inertia of the DEP turbocharger.

The DEP-based engine can achieve improved BSFC and BMEP performance due to
better gas exchange and combustion processes without changing the settings of the
intake and exhaust valve profile. However, introducing fully flexible intake and
exhaust valve mechanism could further optimize the results above.

7 ACKNOWLEDGEMENTS

The authors would like to thank Nick Luard in Jaguar Land Rover Limited for his
kind support in the GT-Power modelling.

We also wish to appreciate the China Scholarship Council to partly sponsor my


living expenses.

38
8 REFERENCE LIST

[1] Möller, C. E., et al. (2005). Divided Exhaust Period - A Gas Exchange System
for Turbocharged SI Engines, SAE International.
[2] Westin, F., et al. (2000). The Influence of Residual Gases on Knock in
Turbocharged SI-Engines, SAE International.
[3] Ayala, F. A., et al. (2006). Effects of Combustion Phasing, Relative Air-fuel
Ratio, Compression Ratio, and Load on SI Engine Efficiency, SAE International.
[4] David Roth, D. M. B. D. P. (2012). “Valve-Event Modulated Boost System:
Fuel consumption and Performance Potential”.
[5] Gundmalm, S. (2013). Divided Exhaust Period on Heavy-Duty Diesel Engines.
Stockholm, KTH Royal Institute of Technology: vi, 72.
[6] Gundmalm, S., et al. (2013). “Divided Exhaust Period: Effects of Changing the
Relation between Intake, Blow-Down and Scavenging Valve Area”. SAE Int. J.
Engines 6(2).
[7] Roth, D. B. and M. Becker (2012). “Valve-Event Modulated Boost System: Fuel
Consumption and Performance with Scavenge-Sourced EGR”. SAE Int. J.
Engines 5(2): 538-546.
[8] Roth, D. B., et al. (2010). Valve-Event Modulated Boost System, SAE
International.
[9] Gamma Technology.
[10] Douaud, A. and Eyzat, P., “Four-Octane-Number Method for Predicting the
Anti-Knock Behavior of Fuels and Engines”, SAE Technical Paper 780080, 1978,
doi:10.4271/780080.
[11] Hong, H., et al. (2004). Review and analysis of variable valve timing
strategies - Eight ways to approach, Professional Engineering Publishing.

39
Advanced boosting technologies for future
SI engine concepts
P Grigoriadis, L Böttcher, E Binder, M Sens
IAV GmbH, Germany

ABSTRACT

This paper examines the influence of future boundary conditions for boosting
systems. It employs engine process simulation to select different intake valve
timings on a four-cylinder 1.4-l spark-ignition engine with a view to illustrating the
extent to which a single-stage boosting system is capable of providing the level of
boost pressure that is demanded. Single-stage boosting systems show a number of
benefits over two-stage systems in terms of cost, package and catalyst light-off.
This was the reason for examining how far adapting or advancing a single-stage
boosting system can meet the rising demands on boost pressure. Variabilities on
the compressor and turbine side were used to increase boosting system spread.
Comprising VTG, wastegate and VTC, the most promising configuration makes it
possible to attain the basic full-load curve using Miller timings (∆IVC=40°). This
reduces consumption in the region of 3 to 6%. Despite demanding a higher level of
boost pressure, the supercharging system provides even further potential for
increasing mean effective pressure by 10% at same low end torque speed.
Measurements on a single-cylinder engine were able to confirm the fuel-saving
trends identified in simulation.

NOMENCLATURE

BMEP Break mean effective pressure n speed in rpm


pcyl Cylinder pressure in bar m
 mass flow kg
s
PFI performance index P Power in kW

Greek Letters
η efficiency
Π pressure ratio

Subscripts
C compressor d displacement
exh exhaust F friction
Rel related Rel related
red reduced RP rated power
is isentropic s static
Spec specific T turbine
TC turbocharger LET low end torque
measured measured value m mechanical
max maximum WG wastegate

_______________________________________
© The author(s) and/or their employer(s), 2014
41
1 INTRODUCTION

Engine developers move between the poles of fuel consumption, emissions and
response behavior while keeping a focus on the applicable exhaust emission testing
and consumption cycles. In current cycles the time-averaged engine operating
range tends to concentrate on the part-load section. For future demands, such as
Real Driving Emissions (RDE), the range considered will extend to the entire engine
operating range. This is why discussion today is looking closer at technologies that
can produce a positive effect on the entire engine operating range. With exhaust-
gas turbocharged SI engines, the focus is on avoiding full-load enrichment and
minimizing throttling losses. The latter can be achieved in the broadest sense by
means of downsizing, downspeeding and lean combustion processes. For this
purpose and as a technology element now firmly established in engine
development, the boosting system must be adapted to accommodate the changed
engine boundary conditions. Single-stage boosting, made up of a rigid radial
compressor and a rigid radial turbine with wastegate, is seen as a standard. In
recent years aerodynamic measures have made it possible to increase throughput
spread and efficiency of such radial machines. However, the question must be
asked as to whether aerodynamics can deliver further potential at reasonable cost.
If not, has the single-stage boosting system reached its limit?

Engine process simulation is a tool that is suitable for examining a question of this
type. The methods for reproducing a boosting system in engine process simulation
have been reconsidered in recent years. For this reason, new methods are briefly
explained in the next section.

2 METHODS AND TOOLS IN THE MODERN DEVELOPMENT PROCESS

2.1 Modern Development Process


The modern development process integrates 1-D engine process simulation to an
even greater extent into configuring an engine concept. This means that engine
concepts undergo preliminary evaluation on the basis of simulation results, with
transient engine behavior carrying greater weight in the evaluation process.
Transient engine behavior is understood to mean both an engine’s responsiveness
(time to torque) as well as catalyst light-off, these being increasingly influenced by
exhaust gas turbochargers.

Figure 1 shows the general path taken in a modern development process. The
simulation results are examined to establish whether they reach target values, such
as drivability at full load, fuel consumption and responsiveness. If applicable,
promising parameters, such as the aerodynamic behavior of a turbocharger, are
adjusted until they reach the target set.

These initially synthetic and iteratively obtained maps then provide the basis either
for selecting an appropriate turbocharger or for configuring new rotors. A key
aspect in this process is the advanced turbocharger model that is explained in more
detail in the next section.

2.2 Advanced Turbocharger Model


The advanced turbocharger model basically comprises the following elements:

- Aerodynamics of compressor and turbine on the basis of maps obtained


from experiments
- Friction on the basis of a map obtained from experiments
- Heat flow and heat storage capability on the basis of semi-empirical
approaches

42
Figure 1: Modern development process for configuring a boosting
system for supercharged engines

Aerodynamic behavior is ascertained directly on IAV’s turbocharger test bench. The


heat transfer normally occurring are avoided as far as possible here by adapting the
measurement boundary conditions. The procedure required for this is described in
detail in [1]. In essence, however, it is based on the following methods and
assumptions:

- Closed-loop system for extending turbine maps and for increasing


compressor mass flow to reduce the flow of heat
- Exhaust gas flap for increasing turbine mass flow to reduce the heat
transfer
- Diabatic criterion χ to assess heat transfer behavior in qualitative terms
- Calculation of turbine power output PT by use of temperature upstream and
downstream of the turbine

Proceeding from the power outputs available for compressor and turbine,

PF  PT  PC and (Eq. 1)

PC
m  (Eq. 2)
PT

can be used for determining frictional power PF or the level of mechanical efficiency
ηm. The latter has already been used with success in [2]. A further measurement on
the turbocharger test bench under normal “hot” boundary conditions permits
validation of the heat model. The heat-storage capability of a turbocharger is also
simulated and is essentially based on turbocharger mass. To validate the model, the
turbocharger is operated in a transient state both on the real-world test bench as
well as in a model test rig. The model’s quality is then determined on the basis of
how well the simulated and measured transient turbocharger speed curve match
up. Semi-empirical approaches are then used for extrapolating map ranges that
cannot be measured. Details on this are provided in [3]. The range to the left of the
surge limit and the range in the fourth quadrant of a compressor map are of
particular relevance to acceleration cycles and to the low-end torque operating point
(LET).

43
2.3 Turbocharger Matching
Classic turbocharger matching is based on the ability to place an engine’s target
full-load curve into a compressor map in such a way as to ensure a safety margin to
the compressor’s surge limit and throughput limit. Furthermore, compressor output
and turbine output are to be equal at each steady-state operating point. As a rule,
the underlying, simplified engine air mass flow curves do not take account of any
valve overlap or recirculated exhaust gas. In future, allowance will be made for
engine boundary conditions of this nature. Combustion process limits, such as
engine knock, will also be included (see Figure 2).

Figure 2: Schematic diagram of the interaction between turbocharger


and engine in the turbocharger nomogram

3 CURRENT AND FUTURE BOUNDARY CONDITIONS FOR BOOSTING


SYSTEMS

3.1 Torque Spread, Power Density and Responsiveness


Usually, the demands on an engine are such that maximum attainable engine
torque is even reached at relatively low engine speeds. For a better understanding,
IAV has defined a characteristic value that can be calculated as follows

BMEPLET  1000
BMEPLET,Re l  (Eq. 3)
nLET

Over recent years power density has constantly increased and averages at approx.
80 kW/l, but can also reach peak levels of up to 133 kW/l [4]. The relevant
characteristic value is calculated as follows

PRP
PRP, Spec  (Eq. 4)
Vd

44
Both characteristic values can be combined to produce a relative performance index
PFIRel which is shown in Figure 3 in the form of an isoline. Today’s volume-produced
engines come with an average relative performance index of 0.35 rising to a
maximum of 0.7. The basis for these data is taken from IAV’s Engine Knowledge
Database. It has to be mentioned, that the maximum values for the attainable
torque at low engine speeds (BMEPLET,Rel) and the the power density (PRP,Spec) are
given by the maximum known values in the database.

PRP,Spec BMEPLET,Re l
PFIRel   (Eq. 5)
PRP,Spec,max BMEPLET,Re l,max

On top of this, the naturally aspirated engine component of an engine map becomes
increasingly smaller as power density rises. This means that a boosting system has
more and more of an influence on any requested sudden load increase from part to
full load. As a result, the impact of the mass moment of inertia associated with the
turbocharger’s response becomes increasingly pronounced, diminishing the
responsiveness of an engine with high power density on account of the design
principle involved.

Figure 3: Relative performance index PFIRel for different engines;


data taken from IAV’s Engine Knowledge Database

3.2 Effective Engine Displacement


Throttling losses can be reduced by using variabilities in the valvetrain on the basis
of various technological approaches. The point at which the intake valves close is
selected in a way that appears to reduce an engine’s displacement. Engine power
output is prevented from falling by opening the throttle, ultimately diminishing
throttling losses. Advancing the closing cycle is usually referred to as Miller timing,
retarding it as Atkinson timing. Apart from dethrottling the engine, the cylinder
charge is also relaxed with Miller timing, producing a reduction in the final
compression temperature. This has a positive effect on the engine’s knock

45
characteristics at high loads, in the region of natural aspirated full load. As a result,
ignition timing can be moved into ranges that provide greater efficiency. Miller
timing tends to impede the formation of intake-port-induced charge motion in the
cylinder which dissipates into turbulent kinetic energy (TKE) in the further course of
piston movement, producing a negative influence on combustion. Lacking
turbulence lengthens the inflammation phase and significantly increases the
duration of combustion. Countermeasures to oppose this, however, have already
been described [5]. Any endeavor to reduce effective engine displacement across
the entire engine operating range for the above-mentioned reasons demands a
higher level of boost pressure – particularly above natural aspirated full load –
which must then come from the boosting system.

3.3 Lean Combustion Process


The lean combustion process involves providing an air-fuel ratio in the combustion
chamber that is leaner than stoichiometric. Depending on the lean concept
implemented, the value of 1, usual for spark-ignition engines, is then globally raised
to as much as 4-5, producing a fall in throttling losses. The lean combustion process
requires a higher level of ignition energy and places elevated demands on the
exhaust gas aftertreatment systems on account of the mixture being lean. Because
of the higher airflow rate required, particularly if lean-burn operation is also to be
used above the naturally aspirated engine load range, the lean combustion process
also demands a higher level of boost pressure from the boosting system [6]. Added
to this, a lean air-fuel ratio leads to lower exhaust gas temperature, which means
that the turbine can ultimately be provided with less exhaust gas energy.

3.4 Number of Cylinders


Reducing engine displacement while maintaining effective engine output is known
as downsizing and, among other aspects, gains its advantage from shifting the
operating point in an engine’s operating map. Downsizing is also used for reducing
friction losses, particularly if reducing displacement involves a smaller number of
cylinders [7], [8]. For the same engine power output, the mass flow characteristics
in the exhaust gas and air path with high time resolution then exhibit higher
amplitudes. This means that a highly fluctuating supply of energy is admitted to a
turbine. In the high engine power output range, this can lead to turbine choking,
ultimately producing relatively low levels of turbine efficiency. The decrease in
airflow rate also fluctuates widely on the fresh-air side. Particularly in surge-limit
vicinity, this can result in unstable compressor operation and consequently limit
achievable engine load [9].

4 INTERACTION BETWEEN ENGINE AND BOOSTING SYSTEM

4.1 Basic Engine Model


Engine process simulation provided the basis for examining engine behavior, in
principle following the procedure described in Section 2. The engine model used is
an exhaust-gas turbocharged 1.4 l four-cylinder spark-ignition engine with an
effective power output of 96 kW and an LET of 19 bar. Applied in equation 5, these
two values produce a relative performance index of 0.36. The valvetrain provides
variabilities at the intake. The simulation environment is GT-Power, and the engine
model provides predictive combustion development computation – SI Turb – as well
as a knock model. The main center of heat release is set to produce maximum
efficiency (8 °CA ATDC). If knock probability increases, ignition timing is
automatically retarded. Boost pressure is regulated by a wastegate.

46
4.2 Boost Pressure Demanded from Changed Engine Boundary Conditions
By way of example, this section examines the influence of Miller timing (see Section
3.2) on the demand for boost pressure. Initially, this involves no changes to the
basic boosting system. Proceeding from basic timing, the intake valve timing was
advanced by 40° CA. Figure 4 shows the full-load characteristics for the basic
configuration (unfilled circle symbol) and torque reduced by 30% at 1,500 rpm for
the basis configuration with Miller timing (filled circular symbol). The drop in torque
is attributable to the reduced engine charge from advanced “intake closes”. The
basic turbocharger was then scaled down in size as shown in Figure 5 (rectangular
symbols), with the full-load curve once again being determined for the engine.
Despite Miller timing, it is now possible to achieve the basic full-load curve again. As
from a speed of 2,000 rpm, however, the turbocharger is pushed to its speed limit.
At 5,500 rpm, mean effective pressure falls from 14.7 bar to 9.4 bar, equating to a
power loss of 32kW.

Figure 4: Full-load curves for the 1.4-l spark-ignition engine from different
valve-train strategies and for different boosting systems

4.3 Selecting and Adapting the Boosting System


To set Miller timings without having to accept the above-described losses in engine
load, further adjustments must be made to the boosting system. In principle, higher
boost pressure is required across the entire rpm range, this being accompanied by
higher boosting system spread. Although a serial-sequential two-stage boosting
system consisting of two exhaust gas turbochargers would deliver this spread [10],
it cannot be considered for the present on the following grounds:

- Additional mass in the exhaust gas path (2nd turbine) acts as a heat sink,
particularly in the heating phase, adversely affecting catalyst light-off
- High demands on packaging
- High demands are placed on ensuring a tight seal at the exhaust gas flap
additionally installed in the exhaust gas path as it needs to switch between
high-pressure and low-pressure turbine
- High cost

47
Figure 5: Turbocharger maps used in simulation for
producing different configurations

However, a two-stage serial-sequential system could be restricted to the fresh-air


side alone. Combinations consisting of an exhaust gas turbocharger compressor and
an electrically or mechanically driven compressor are conceivable. However, it
should be remembered that in the engine concept considered here an additional
compressor of this kind must be in a position to deliver boost pressure over a
prolonged period. High load on the vehicle electrical system, relatively high cost,
packaging and, ultimately, less potential to save fuel all work against the electrical
system. The last three points also apply to the additional mechanical system.

For the above reasons it would make sense to get the single-stage boosting system
to meet the underlying demands. This is done by fitting both compressor and
turbine with a variable guide vane system. The basic compressor is equipped with a
VTC system (Variable Trim Compressor) that can alter the effective inlet cross

48
section of a rotor [11]. As a result, the trim appears capable of changing the ratio
between rotor inlet and rotor outlet diameter. Reducing the inlet cross section by
40% shifts the map towards smaller mass flows, thereby resulting in a shift in the
surge limit (up to 33%). As this is a variable system, the initial cross section can be
reproduced without any additional pressure loss in the air path.

On the hot gas side, the wastegate turbine currently used is replaced with a VTG
turbine familiar from diesel engines. Although a VTG turbine has also been used in
the SI engine segment since 2005 [12], it is still a niche application. In just the
same way as the wastegate turbine, the VTG turbine provides a wider spread in the
turbine map range. This, however, comes with relatively minor loss in turbine
efficiency. As the entire exhaust gas mass flow has to be fed through the rotor of a
VTG turbine, this must be made larger than the rotor of a wastegate turbine. In the
studies conducted here, rotor diameter is therefore increased from 37 to 40 mm.
This geometric modification is always accompanied by an increase in the mass
moment of inertia and, on the basis of available data, is put at about 30%. In
addition to the previously mentioned turbocharger maps for the basic (rigid
compressor, wastegate turbine) and the scaled turbocharger Figure 5 shows the
turbocharger maps for the advanced boosting system (VTC compressor, VTG
turbine). The latter were synthesized on the basis of experience gathered from
experiments and by employing analytical approaches. IAV’s own matching tool was
used to match the VTG turbine to the engine in such a way as to achieve the
engine’s LET operating point.

As shown in Figure 4, the basic full-load curve cannot be completely reproduced


with this configuration (triangular symbol). Proceeding from the LET operating
point, the VTG system continually opens and, at an engine speed of 3,500 rpm, is
open all the way. As the VTG system cannot be opened any further, the
turbocharger reaches its speed limit above 3,500 rpm with throttle position
unchanged (100%). This necessitates further increasing turbine throughput spread
and turbine efficiency which is done by adding a wastegate flap. This measure
ultimately makes it possible to achieve the basic full load curve (see Figure 4,
diamond symbols). Below, this configuration is referred to as “single-stage
advanced+”. The nomogram in Figure 6 presents the full-load curves for the “basic
configuration” and “single-stage advanced+” options. The compressor map shows
the higher level of boost pressure required across the full-load curve. At the LET the
boost pressure differential is approx. 300 mbar, at 4,000 rpm even 490 mbar. As
the compressor map is wider for the “single-stage advanced+” configuration,
enabled by the VTC system, the engine operating points are within the stable map
range. On account of its throughput spread and relatively high levels of efficiency,
the turbine is able to provide the compressor with the necessary drive power. At
maximum engine power output, the turbine pressure ratio and, with this, the
exhaust gas backpressure, are shown to remain at about the same level in spite of
boost pressure being higher. Specific fuel consumption can be lowered by between
3 and 6% along the engine full-load curve. This is because it was possible on
average to advance the main center of heat release by 3 °CA and reduce pumping
work by an average of 130 mbar. In the investigations presented hitherto, an
attempt was made to achieve the same mean effective pressure. Using the
supercharging unit’s full potential to increase power output does, however, make it
possible to increase maximum mean effective pressure from an initial 19.9 to
21.9 bar and maximum effective power output from 96 to 107kW. This corresponds
to an increase of 10 %, taking the relative performance index to 0.44. The extent to
which avoiding scavenging can influence the boosting system was also examined.
For the engine under study here, this influence is relatively slight which probably
has to do with the basic configuration’s valve overlap being low to start with.

49
Figure 6: Turbocharger nomogram for the basic configuration (black)
and the "1-stage advanced+" configuration (red)

5 EXPERIMENTS ON THE SINGLE-CYLINDER ENGINE

To validate the above-described simulations, experiment-based studies were


conducted on a single-cylinder research engine already used in [5]. This has a fully
variable valve train and provides the capability of giving the injector a centered or
lateral position. In a test series at constant load (n=1,500 rpm, full load), “intake
closes” timing was gradually advanced while adjusting boost pressure and taking
the air-fuel ratio to 1. Figure 7 shows the savings measured in specific fuel
consumption. In general, an increasing scavenging–pressure differential as well as
increasing Miller timing are both seen to have a positive influence on specific fuel
consumption. Further simulations were carried out using the engine model
described in 4.1. This demonstrates the benefits as well as the limitations of
additional variabilities. Proceeding from an almost closed wastegate and open VTG
system at standard timings, specific fuel consumption can be reduced by approx.
5.4% (green rectangles) by increasing Miller timing. However, closing the VTG by
any more takes the improvement in consumption back to 4.3%. At this operating
point, consumption can only be improved further by reducing turbine size. The
white rectangle represents a simulated engine operating point with a turbine
downsized by 25%. Further investigation would, however, be needed to ascertain
whether it is possible to achieve the effective power output of 96 kW with this
turbine.

50
Figure 7: Fuel savings measured and computed while varying
scavenging-pressure differential and timing at operating
point n=1,500 rpm and full load

6 SUMMARY AND OUTLOOK

It was possible to show that changes in the engine’s boundary conditions have a
considerable impact on engine behavior. By way of example and representative of
future technologies, Miller timing was applied to a 1.4-l spark-ignition engine as the
basis for examining the behavior of different single-stage boosting systems.
Comprising a variable trim compressor (VTC) and a variable geometry turbine
(VTG) with wastegate, the “single-stage advanced+” configuration was shown to be
highly promising in this regard. This configuration is capable of achieving the basic
full-load curve. Mean effective pressure can be increased by 10% as well. On
account of high throughput spread, the higher demand for boost pressure can also
be met for engine operating points between intake full load and full load. The
potential is also given for satisfying the other future boundary conditions stated
above, such as the lean combustion process.

Measurements on a single-cylinder engine were able to confirm the fuel-saving


trends identified in simulation.

7 REFERENCES

[1] Grigoriadis, P., Binder, E., Böttcher, L., Benz, A. et al., Advanced Turbocharger
Model for 1D ICE Simulation - Part I, 2013, SAE Technical Paper 2013-01-
0581, doi:10.4271/2013-01-0581
[2] Otobe, T., Grigoriadis, P., Sens, M., Berndt, R., 2010, Method of performance
measurement for low turbocharger speeds, 9th International Conference on
Turbochargers and Turbocharging, IMechE, London

51
[3] Grigoriadis, P., 2008, Experimental data acquisition and modeling of unsteady
flow phenomena of vehicle engine turbocharger compressors, Ph.D. Thesis,
Technical University Berlin, Germany
[4] Hart, M. et al., 2013, Der neue Hochleistungsvierzylindermotor mit
Turboaufladung von AMG, 34th International Vienna Motor Symposium, Vienna
[5] Riess, M., Benz, A., Wöbke, M., Sens, M., 2013, Intake Valve Lift Strategies
for Turbulence Generation, Article in MTZ worldwide Edition: 2013-07
[6] Kneifel, A., Pape, J., Sens, M., 2008, Investigations on Supercharging
Stratified Part Load in a Spray-Guided DI SI Engine, 2008, SAE Technical
Paper 2008-01-0143, doi: 10.4271/2008-01-0143
[7] Lee, S., Gu, Y., Kim, T., Hahn, J., 2011, The New Hyundai-Kia 1.0 l Three-
cylinder Gasoline Engine, Article in MTZ worldwide Edition: 2011-08
[8] Ernst, R., Friedfeldt, R., Lamb, S., et al., 2011, The New 3 Cylinder 1.0L
Gasoline Direct Injection Turbo Engine from Ford, 20th Aachen Colloquium
Automobile and Engine Technology, Aachen
[9] Cuniberti, M., Micelli, D., Stroppiana, A., Venezia, C., 2011, Charging system
for a small bi-cylinder engine: the TwinAir experience, 16th Supercharging
Conference, Dresden
[10] Kuhlbach, K., Werner, J., Kiener, T., Becker, M., 2013, Innovative Two-Stage
Turbocharging System with Cooled Regulating Valve for Gasoline Engines, 22nd
Aachen Colloquium Automobile and Engine Technology, Aachen
[11] Grigoriadis, P., Sens, M., Müller, S., 2012, Variable Trim Compressor – A New
Approach to Variable Compressor Geometry, 10th International Conference on
Turbochargers and Turbocharging, IMechE, London
[12] Gabriel, H., Lingenauber, R., Ramb, T., 2006, Der Turbolader mit variabler
Turbinengeometrie (VTG) für den neuen Porsche 11 Turbo – Ein Meilenstein in
der Ottomotorenaufladung, 11th Supercharging Conference, Dresden

52
Design and testing of a high flow
coefficient mixed flow impeller
H R Hazby
PCA Engineers Ltd, UK
M V Casey
PCA Engineers Ltd, UK
Institute of Thermal Turbomachinery (ITSM), University of Stuttgart, Germany
R Numakura
Turbo Machinery and Engine Technology Department, IHI Corporation, Japan
H Tamaki
Corporate Research and Development, IHI Corporation, Japan

ABSTRACT

The design of a mixed flow compressor stage with an extremely high flow
coefficient ( ) of 0.25 and a high pressure rise coefficient ( ) of 0.56 is described.
The objective of the work was to explore the performance potential in this highly
unconventional area of the design space and to assess the capability of design
methods. The paper describes the aero-mechanical design approach for the
preliminary design and discusses the challenges involved in developing such highly
loaded compact stages. The test data obtained on a prototype stage is also
presented. The results show that acceptable performance levels can be achieved at
these extreme design conditions and further exploration of the design space is
worthwhile.

NOMENCLATURE

Cm meridional velocity (m/s)


Cp isobaric specific heat capacity (J/kg K)
diameter (m)
total enthalpy (J/kg)
absolute Mach number (-)
= / tip speed Mach number (-)
mass flow rate (kg/s)
SM Surge Margin (-)
temperature (K)
blade speed (m/s)
= / volume flow rate based on inlet total density(m3/s)
relative velocity (m/s)
absolute flow angle from radial direction (º)
relative flow angle from radial direction(º)
total-to-total efficiency (-)
pressure ratio (-)
density (kg/m3)

___________________________________________
© The author(s) and/or their employer(s), 2014
55
Subscripts
0 stagnation properties
1 impeller inlet
2 impeller outlet
m mean diameter
h hub diameter
s isentropic flow process
t tip diameter

1 INTRODUCTION

In many turbocharger applications, such as diesel engines for standby power, there
is a strong requirement for compact designs to reduce the size and cost of the
installation. The requirement for compact centrifugal compressors with high
swallowing capacity leads to the application of small mixed flow impellers with
extremely high flow coefficients (1). The small size and high speed increases the
Mach numbers in the flow channels and this may conflict with the achievement of
good efficiency and wide operating range. Mixed flow stages also have lower radius
change across the impeller than stages with lower flow coefficients so that a high
pressure rise is more difficult to achieve than in conventional stages as there is a
smaller centrifugal effect.

Some guidelines for the expected performance and the preliminary design of radial
stages with a high swallowing capacity have been published by Rusch and Casey
(2). However, the potential performance levels and operating range that can be
achieved by high flow coefficient mixed flow impellers have not been thoroughly
addressed in the literature. The use of modern design methods is expected to
alleviate the performance deficit arising from the extreme duty, but to what extent
this is possible in the relatively uncharted territory of mixed flow designs is not
known. Furthermore, clear design guidelines for this type of turbomachine have not
been published.

This paper attempts to contribute in this area by discussing the issues faced in the
design of mixed flow impellers for extreme duties and the achievable performance
levels, based on measurements on a prototype compressor stage. The target
design point for the stage was set at a flow coefficient ( = / ) of 0.25 and an
isentropic pressure rise coefficient of ( = ∆ / ) of 0.56, which is far outside of
the conventional design space for centrifugal compressors.

In the next sections the general conceptual design issues regarding the design of
very high flow coefficient mixed flow impellers are discussed and the impacts of the
mechanical integrity requirements on the aerodynamic design of the compressor
are addressed. Then, the specifications of the final compressor design accompanied
with the performance measurements are presented. Detailed design of the stage
and further analysis of the performance can be found in Hazby et al. (3)

2 CONCEPTUAL DESIGN CONSIDERATIONS

The current impeller was designed to deliver a pressure ratio of 2.65 at an inlet
volume flow of 1.02 m3/s and a rotational speed of 77525 rpm. The impeller outlet
diameter was fixed at 100mm. As a first indication of the difficulty of the design,
the guidelines of Casey et al. (1) can be used to place the design point in the
recommended design space for centrifugal compressors, as a function of the design
flow coefficient and the required isentropic pressure rise coefficient, defined below,

56
= (1)


= = (2)

Guidelines for the achievable pressure rise coefficient at a particular design flow
coefficient are given in Figure 1, which is a modified form of the Cordier diagrams
given in (1). A similar diagram can be found in Dixon and Hall (4). The dashed lines
represent the range of the current experience. The general trend is a continuous
reduction in the pressure rise capacity of the impeller as the flow capacity is
increased. For the same compressor duty in terms of pressure ratio and mass flow
rate, compact centrifugal compressor stages can be produced by reducing the
impeller outlet diameter. If the rotational speed is maintained constant, the
reduction in outlet diameter can drive the impeller flow coefficient and the pressure
rise coefficient to extremely high values, facing the designer with a number of
challenges regarding the aerodynamic performance and mechanical integrity of the
wheel.

The pressure rise coefficient for the current impeller at the design operating point is
0.56, which is well outside of the range of the conventional centrifugal compressor
design shown in Figure 1. For comparison, the pressure rise coefficients
corresponding to impeller designs with outlet diameters of 110 and 120 millimeters
are also shown in the figure. For a constant rotational speed and pressure ratio the
pressure rise coefficient decreases as the outlet diameter is increased whereby the
design with a 120mm tip diameter is placed within the recommended design range.
The intention of this study was however, to push the limits of conventional designs
as far as possible into the mixed flow region and assess the achievable performance
levels in this design space. Hence, the impeller was designed at a mean outlet
diameter of 100mm. The aero-mechanical design considerations involved in the
development of such stages is discussed in more detail in the next sections.

0.7

0.6
Pressure rise Coefficient

100 mm
0.5

110 mm
0.4
120 mm
0.3

0.2

0.1

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Flow Coefficient
Figure 1 Comparison of the effect of the impeller diameter on the
location of the design point in a modified Cordier diagram.

57
3 PRELIMINARY DESIGN CONSIDERATIONS

The conceptual considerations given above identify the level of difficulty in terms of
the global experience with centrifugal compressors from other sources. In this
section some more detail is provided of the geometry of possible competing stages
for this application and the relevant aerodynamic design issues are discussed.

= 120 = 110 = 100 = 100


= 0.15 = 0.19 = 0.25 = 0.25
= 0.39 = 0.47 = 0.56 = 0.56
= 54° = 40° = 18° = 18°
= 55° = 55° = 55° = 55°
= 0.72 = 0.80 = 0.89 = 0.89
/ = 0.69 / = 0.59 / = 0.52 / = 0.61
/ = 1.22 / = 1.05 / = 0.93 / = 0.80
/ = 1.28 / = 1.17 / = 1.07 / = 1.16

Figure 2 Effect of outlet diameter on the impeller design at the same duty
of pressure ratio and mass flow rate.

The amount of pressure rise that can be achieved at a particular mass flow rate is
primarily a function of the wheel diameter, relative diffusion and the backsweep at
the outlet of the impeller. It is the variation of these parameters that influences the
design constraints and performance of the impeller as its size is reduced. These
effects are discussed with the help of Figure 2, which demonstrates four impellers,
designed for the same duty but with different outlet diameters. The impellers were
designed for the compressor duty as mentioned above and assuming a constant
impeller total-to-total efficiency of 0.86 for all cases. To obtain a fair comparison,
the absolute flow angle at the inlet of the diffuser is kept equal to 55˚ for all cases.
The required pressure ratio was achieved by varying the backsweep angle as the
impeller tip diameter was reduced.

The first stage, shown in Figure 2, with an impeller diameter of 120mm has the
lowest flow coefficient of 0.15 which is considered above the value for an optimal
efficiency (which is about 0.09) due to relatively high inducer losses (see Rusch and
Casey (2)). The tip diameter ratio of / = 1.28 suggests a relatively sharp
turning of the flow on the shroud contour of this impeller. A difficulty that arises
from reducing the outlet diameter is that with the inlet tip diameter fixed, a higher
curvature is needed on the shroud to turn the flow from axial to radial direction.
This results in the acceleration of the high speed flow on the shroud and increased
hub-to-shroud secondary flows inside the passage. The flow acceleration on the
shroud can be seen in the results of the throughflow calculations, using the
throughflow code Vista TF as published by Casey and Robinson (3), shown in Figure
3 and Figure 4. It should be noted that the extreme flow acceleration observed
near the outlet of the 0.25 flow coefficient radial impeller, corresponding to an
outlet diameter of 100mm, does not happen in reality as the flow will most
probably separate on the sharp corner. In this impeller, which has a tip diameter
ratio of 1.07, turning of the flow from axial to radial is not practically possible due
to the high curvature near the impeller outlet and therefore a mixed flow design
with the same flow coefficient becomes necessary.

58
Cm (m/s)

= 0.15 = 0.19 = 0.25 = 0.25

Figure 3 Meridional velocity contours from a throughflow calculation of the


impellers shown in Figure 2.

Hub section Tip section


0.80 1.60

Mean relative Mach number


Mean relative Mach number

0.70 120mm radial flow 1.40

0.60 110mm radial flow 1.20

0.50 100mm radial flow 1.00


100mm mixed flow
0.40 0.80 120mm radial flow
0.30 0.60 110mm radial flow
0.20 0.40 100mm radial flow
0.10 0.20 100mm mixed flow
0.00 0.00
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Meridional distance Meridional distance

Figure 4 Variation of the mean relative Mach number on the hub and
shroud contours.

The higher inlet relative velocity at the tip of an impeller blade means that the flow
streamline at the tip section undergoes a stronger diffusion from inlet to outlet than
that on the hub. In a typical centrifugal impeller such as stage 1 in Figure 2, the
ratio of the outlet to inlet relative velocity at the tip section is close to 0.7.
However, the flow on the hub section in fact accelerates due to a high relative
velocity at the impeller outlet compared to that at the inlet. This makes the
diffusion on the shroud a critical constraint for the impeller design with values of
relative velocity ratio of greater than 0.5 usually used to avoid excessive diffusion
at the tip section. The accelerated flow on the hub is generally more tolerant to the
angle variations. However, the deceleration at the hub increases as the outlet
diameter is reduced. This can be observed in the value of / given in Figure 2
where it reduces from 1.22 to 0.93 when the impeller outlet diameter is reduced
from 120mm to 100mm. Moving from a radial design to a mixed flow design with
the same mean outlet diameter and 45˚ meridional exit angle, relieves the diffusion
on the casing but further increases the diffusion on the hub by more than 15% due
to the reduction in the hub diameter at the outlet. This can also be seen in Figure 4,
where the variations of the mean relative Mach number is shown for hub and
shroud contours for all four impellers. In the limiting case of an axial compressor it
is actually the hub section which imposes the critical limit on the loading, due to
low blade speed in the hub section.

The higher overall diffusion level that is needed at smaller outlet diameters is partly
due to the smaller centrifugal pressure rise inside the impeller passage. In a typical
centrifugal compressor, about 50% of the pressure rise is generated by the
centrifugal effects due to the rotation of the fluid particles about the axis. This is an
efficient means of compression as it is independent of diffusion process and flow
separations inside the narrow passages. This effect is reduced as the flow
coefficient is increased (outlet diameter is reduced), which means more diffusion
needs to be done inside the passage to achieve the same level of pressure rise.

59
The lower centrifugal pressure rise due to smaller impeller outlet diameters cannot
be entirely compensated by a reduction in the relative velocity ratio and therefore
smaller backsweep angles are needed to deliver the target pressure ratio. As can be
seen in Figure 2, the required backsweep angle decreases from 54º to 18º as the
impeller outlet diameter is reduced from 120mm to 100mm. A higher backsweep
angle was not used for the 100mm diameter design as this would increase the
relative diffusion at the impeller tip to unacceptably high levels. The required low
backsweep angle results in narrower characteristics with smaller surge margin by
moving the peak efficiency operating point closer to the surge line. This can be
observed in the expected performance maps of the three radial stages described
above with 120mm, 110mm and 100 mm diameters, shown in Figure 5. The
performance maps were predicted using the method presented by Casey and
Robinson (6), modified to be used for very high flow coefficient impellers similar to
that of the current study. A validation of the code for mixed flow stage is provided
in Casey et al. (7). The position of the target design point is depicted by the red
circles in the pictures. The diagrams identify that considerable difficulty will be
expected with a small operating range as the characteristics become less steep as
the backsweep is reduced for the smaller stages.

Figure 5 Expected variable speed performance maps for the three radial
compressors with 120, 110 and 100 mm diameter, designed for the same
aerodynamic duty, as predicted by the method of Casey and Robinson (5).

4 MECHANICAL DESIGN IMPLICATIONS

As discussed earlier, the aerodynamic design of the tip section of a centrifugal


impeller is more critical than the design of the hub section due to its higher
diffusion levels. This is particularly true for high speed impellers where the control
of the supersonic flow in the inducer region has a significant impact on the
performance and operating range of the compressors. Therefore, a greater degree
of freedom is often exercised in the design of the hub section to satisfy the
mechanical integrity criteria. The natural frequency of the impeller vanes decreases
significantly as the outlet diameter is reduced, demanding a greater compromise on
the thickness and camber distribution at the hub section. This, however, opposes
the requirement of higher relative diffusion at the hub of the high flow coefficient
impellers, which could limit the freedom of the design in this section.

60
These mechanical requirements also compromise the aerodynamic design at higher
spanwise positions, where the vane thickness and curvature need to be minimized
to reduce the aerodynamic shock losses. In addition, the choice of the vane
thickness can limit the number of the vanes that can be fitted on the hub section.
This in turn can alter the detailed design of the tip section, optimized for handling
the high speed flow. Furthermore, it requires thinner splitter vanes with their
leading edge placed further downstream in the passage to allow a sufficient flow
capacity at the hub section. The result could be swept splitter vanes with high
stress levels in the vane area as shown in Figure 6. Under these circumstances the
aerodynamic optimization of the splitter vanes is very much limited by the
mechanical design requirements. Further details of the mechanical design are given
by Hazby et al. (3).

High stress
region

Figure 6 Stress distribution in the main and splitter vanes.

5 FINAL STAGE DESIGN AND TEST DATA

In the previous sections, various aspects of the aero-mechanical design of a high


flow coefficient centrifugal impeller were discussed. The final design of the stage is
presented here. This impeller was designed to achieve a pressure ratio of 2.65 at
an inlet volume flow rate of 1.02 m3/s and a rotational speed of 77525 rpm. The
outlet mean diameter was set at 100mm, corresponding to an extremely high flow
coefficient of 0.25.

As discussed in section 3, a mean outlet diameter of 100mm requires a mixed flow


style impeller, where turning of the flow from axial to radial is partly accomplished
in the vaneless diffuser. The constraint was applied only on the outlet mean
diameter and therefore the hub and tip diameters could change as the meridional
position of the blade at the trailing edge was not restricted. The final impeller
consisted of 9 main vanes and 9 splitters and was designed with a back sweep
angle of 28˚ degrees at the outlet mean radius.

The impeller vane profiles were designed at several spanwise locations to optimize
the aerodynamic performance of the impeller by controlling the supersonic flow at
the inlet as well as ensuring the mechanical integrity of the wheel. This meant that
the conventional straight line generators needed for flank milling of the impeller
blades could not be used and therefore the impeller was point milled to allow the
use of curved line generators. At the exit of the impeller, the flow was turned to the
radial direction using a converging vaneless diffuser followed by a parallel walled
diffuser section, which was required as a design constraint. The impeller geometry
and meridional shape of the passage are presented in Figure 7. Further details of
the aerodynamic design of the blade sections are given by Hazby et al. (3).

61
Figure 7 Meridional passage and impeller geometry.

The stage as designed was tested in a dedicated turbocharger test stand. The
measurements were carried out at five rotational speeds ranging from 44350 rpm
to 77100 rpm. The variations of the stage total pressure ratio and total-to-total
efficiency against the inlet volume flow rate are presented in Figure 8. For
comparison, the results of the 1D performance predictions, using the method of
Casey and Robinson (5), are also shown in the figure. At near design speed of
77100 rpm, the stage achieved a peak total-to-total efficiency of above 80% at a
pressure ratio of 2.75. This corresponds to an isentropic pressure rise coefficient of
0.595 which is about 6% higher than the intended design value. As no test data in
this region of the design space was previously available to the authors, the
coefficients in the 1D prediction method have been adapted to match the current
stage performance. Overall, an excellent agreement can be observed between the
measured and the predicted performances across the map.

The surge margin at a nominal pressure ratio can be defined as:

SM = (Design flow rate – Surge flow rate) / (Design flow rate)

Figure 8-a Measured and predicted performance map,


stage pressure ratio versus volume flow.

62
Figure 8-b Measured and predicted performance diagrams,
stage efficiency versus volume flow.

Based on the mass flow rate at peak efficiency condition, a surge margin of just
above 6.1% was achieved by the compressor. Such a narrow flow range is typical
of high speed compressors with highly supersonic flows at the inducer tip section.
Other design features, such as low backsweep angle and possible flow separation in
the curved vaneless diffuser, could also contribute to narrow the flow range at this
extreme design condition. Therefore, the application of the compressor for practical
applications at such high flow coefficients may be possible only through the
application of Map Width Enhancement devices.

6 CONCLUSIONS

The conceptual design and testing of a mixed flow compressor stage in a highly
unconventional region of the design space with an extremely high flow coefficient
and a high pressure rise coefficient is described. The paper shows some of the
issues related to the conceptual and preliminary design of the stage. The
preliminary design methods used for the design identified many of the difficulties
expected from such an unconventional design, in particular the need for a mixed
flow stage, the expected narrow operating range and the difficult mechanical
issues. The final testing of the stage identified the need for some form of
recirculating bleed system to increase the operating range of the stage. The test
data has been used to improve the prediction capability of the Casey Robinson map
prediction method for such stages.

7 REFERENCES

[1] Casey, M.V., Robinson C.J., and Zwyssig, C. (2010), “The Cordier line for
mixed flow compressors”, Proceedings of ASME Turbo Expo 2010, June 14-18,
2010, Glasgow, UK, ASME Paper GT2010-22549
[2] Rusch, D., and Casey, M.V., (2013), “The Design Space Boundaries for High
Flow Capacity Centrifugal Compressors” ASME Journal of Turbomachinery. Vol.
135 (3), 031035
[3] Hazby, H., Casey, M.V., Numakura, R., Tamaki, H. (2014), “A Transonic Mixed
Flow Compressor for an Extreme Duty”, ASME Paper GT2014-25378,

63
submitted to ASME Turbo Expo 2014: June 16 – 20, 2014, Düsseldorf,
Germany
[4] Dixon, S. L., and Hall C.A., (2010), Fluid mechanics and thermodynamics of
turbomachinery, Butterworth-Heinemann, Boston/Oxford, 6th Edition
[5] Casey, M.V., and Robinson, C.J., (2010), “A new streamline curvature
throughflow code for radial turbomachinery”, ASME Journal of turbomachinery,
2010, Vol. 132, April, 2010
[6] Casey, M.V, and Robinson, C.J., (2013) “A Method to Estimate the
Performance Map of a Centrifugal Compressor Stage, ASME Journal of
Turbomachinery. 135 (2), 021034
[7] Casey, M.V., Krähenbuhl, D., Zwyssig, C., (2013), “The design of ultra-high
speed miniature centrifugal compressors”, 10th European Turbomachinery
Conference, ETC10, Lappeenranta, Finland, April 2013

64
Aerodynamic design optimization of a
centrifugal compressor impeller based on
an artificial neural network and genetic
algorithm
S Ibaraki 1, R Van den Braembussche 2, T Verstraete 2, Z Alsalihi 2,
K Sugimoto 1, I Tomita 1
1
Mitsubishi Heavy Industries, Ltd, Japan
2
von Karman Institute for Fluid Dynamics, Belgium

ABSTRACT

Centrifugal compressors are applied to turbochargers, industrial compressors and


turbo shaft gas turbine engines because of their high pressure ratio, relatively wide
operating range and cost benefits. The internal flow in a centrifugal compressor
impeller is however three dimensional and shows very complex flow phenomena,
which makes the understanding of the loss generating mechanisms difficult and
requires a considerable design effort to reach good performance. Especially
turbocharger compressors impose a challenge to the designer when both a very
wide operating range and high efficiency are required. The design effort can
however be reduced by applying an advanced design optimization system as an
alternative to a conventional manual design based on the experience of the
designer.

In this study a centrifugal compressor impeller for an automotive turbocharger was


designed by means of an aerodynamic design optimization system composed of an
artificial neural network (ANN) and a genetic algorithm (GA). This resulted in two
newly designed centrifugal compressor impellers which were further studied both
numerically and experimentally. One has higher efficiency with slightly wider
operating range compared to the baseline impeller. The other one has a twice as
wide operating range compared to the baseline impeller with a minor decrease in
efficiency.

1 INTRODUCTION

Today turbocharging has become a fundamental technology to realize engine


downsizing, which is an attractive strategy for low carbon emissions. High efficiency
and wide operating range are strongly required for the automotive turbochargers(1).
Especially centrifugal compressors for automotive turbochargers should operate
with high efficiency from the surge limit to the choke limit. To increase the
operating range of centrifugal compressors, recirculation bypass of the inducer is
employed for some applications(2). Also some devices such as variable inlet guide
vanes(3) and low solidity diffusers(4) have been studied. Furthermore two stage
turbocharging(5) has been applied to give a considerable wider operating range.

_______________________________________
© The author(s) and/or their employer(s), 2014
65
However, all above countermeasures need extra cost and complexity of the
turbocharging system.

Modern, high performance centrifugal compressor impellers demonstrate complex


3D blade geometries with a highly complex 3D flow(6)-(9). The efficiency of
centrifugal compressors has been improved over several decades by the progress of
computational fluid dynamics and experimental fluid dynamics, which makes it
increasingly more difficult to improve the efficiency further.with reasonable stable
operating range. Therefore, centrifugal compressor design can benefit largely from
modern optimization techniques. Inverse design method have been applied to
centrifugal compressors and other turbomachinery as optimization methods(10). This
method requires to specify an optimal blade loading to generate the blade profile.
In general, designers have difficulties to specify the optimal blade loading to
achieve a performance improvement. The loss generating mechanisms in a
centrifugal compressor are very complicated and have not been revealed completely.
The blade loading is one of the key parameters affecting the compressor
performance. Therefore, it is very difficult to specify the optimum blade loading.
Alternatively, design optimization methods composed by an ANN and a GA requires
no prescribed aerodynamic parameters.

In this study a centrifugal compressor impeller for automotive turbochargers was


designed by means of an aerodynamic design optimization system composed of an
ANN and a GA. This resulted in two new centrifugal compressor impellers which
were further examined. One has a higher efficiency with a slightly wider operating
range compared to the baseline impeller. The other one has a twice as wide
operating range compared to the baseline impeller with a minor decreased
efficiency.

This paper describes the concepts and the procedure of the design optimization
system based on an ANN and a GA. Experimental and numerical investigations of
the newly designed impellers are presented and discussed.

2 OPTIMIZATION STRATEGY
Start
Geometry
2.1 Optimization procedure
Figure 1 shows schematically the design Requirement
procedure. The optimisation is driven by a GA
GA in which the performance of each Geometry Blade
Generation
geometry is analysed by means of an ANN
ANN
trained on the information contained in a
3D NS
database. Once the GA has found an Analysis
Performance Prediction
optimum it is verified by a Navier Stokes
solver(11) and is added to the database,
Performance Data Learning
resulting in a more accurate ANN. This Base
procedure is repeated for a given number
of iterations. A more complete description Stop
of the optimisation method can be found Figure 1: Optimisation procedure
in published papers (12),(13).

2.2 Boundary conditions


The compressor is intended to work for atmospheric conditions and should deliver a
total to static pressure ratio of 1.53. At a result, the atmospheric pressure is
imposed as total inlet pressure, and the total inlet temperature is fixed at 293.15K.
The inlet flow is axially oriented. At the outlet a static pressure is imposed
corresponding to the required pressure ratio.

66
2.3 Geometry definition
The impeller is backswept with splitter blades. The blade number is allowed to be
changed in between minimum 4+4 and maximum 6+6. The minimum and
maximum axial length are fixed. The maximum lean angle at the leading edge and
trailing edge, which are not directly controlled by the used parameterisation, are
given as design constraints.

The geometry is defined by the meridional contour and the camberline blade angles
of the full and splitter blades at hub and shroud. The meridional contour definition is
schematically shown in Figure 2, where subscript 0 denotes the leading edge of the
blade and 3 denotes the trailing edge. The inlet contour at the hub and shroud is
defined by a second-order Bézier curve. The hub inlet and leading edge diameters
are fixed. The hub axial position and the shroud leading edge diameter are design
variables. The leading edge axial position of the full blade at the shroud is set equal
to the leading edge at hub. The points denoted by in (Xin,Rin), A and 0 (X0,R0) are
Bézier control points and constitute the Bézier polygon. The leading edge position of
the splitter blade is defined by the design variables Ule at hub and shroud. Ule is the
percentage of meridional length where the leading edge starts.

Figure 2: Meridional contour definition Figure 3: Blade angle definition

The blade meridional contour at hub and shroud are defined by third-order Bézier
curves with 4 control points, denoted by 0, 1, 2 and 3. At hub both axial and radial
coordinates of points 1 (X1,R1) and 2 (X2,R2) are design variables. The axial
position of point 1 and both coordinates of point 2 are design variables. The trailing
edge diameter is set equal at hub and shroud. The trailing edge diameter is not a
design variable but it is adjusted to compensate the variation in the blade trailing
edge metal angle, which is a design variable. The diffuser exit diameter is fixed
equal at the hub and at the shroud. The diffuser exit width is also fixed. The diffuser
meridional contour is defined by second-order Bézier curves and the intermediate
Bézier control point B is computed to have a parallel diffuser downstream of that
point and smooth trailing edge curve.

The camberline blade angle definition is achieved using third-order Bézier curves in
the form of Bernstein polynomials as shown in Figure 3.

 (u )   0 (1  u )3  1u (1  u ) 2   2u 2 (1  u )   3u 3 (1)

Here u is a parameter between 0 at the leading edge and 1 at the trailing edge. The
parameters β0 and β3 are the metal angles at the leading and trailing edges, β1 and
β2 are intermediate parameters in eq. 1 and have no physical meaning. These

67
parameters are the 8 design variables for the full blade camberline at hub and
shroud. The splitter blades are restricted to have the same trailing edge metal angle
as the full blades, hence this results in 6 additional design variables for the splitter
blades. Same pitch at the hub and shroud is imposed at the trailing edge. The
description of the blade is completed with the pre-specified thickness distribution. In
total 27 design variables define the geometry.

2.4 Objective functions


The GA is based on the evaluation of each individual’s fitness, which is the inverse
of the total penalty called the objective function. The penalties are associated with
the Navier-Stokes analysis of the best individual proposed by the GA at the end of
each optimization cycle. The ANN penalty is associated to each individual’s
performance predicted by the ANN during the optimization cycle. In both cases the
penalty is formulated identically. The same performance vector is either predicted
by the ANN or post-processed from the Navier-Stokes solution.

The vector of the 27 shape parameters, normalized by their limit values (the ANN
requires input and output values between 0 and 1), constitutes the ANN input. The
performance vector (as well the ANN output) from which the objective is computed,
consists of the two mass flows m1 and m2 (see ∆mass penalty below), the total-to-
total efficiency, distortion and skew of the diffuser exit radial velocity profile and full
blade and splitter blade Mach numbers. The Mach number is parametrized at 20
points each, along respectively the pressure and suction sides of the blades and
splitters. The total performance vector has 5+8*20=165 elements. The total
penalty is the sum of the following items computed using the elements of the
performance vector:

Mass penalty
The difference between the required mass flow and the actual mass flow is
penalized.

max[m req  (m1  m 2 )]  mref / 300,0.0 (2)


Pm  wm
mref

There is a tolerance value of mref/300.0 with mref=mref for which there is no penalty.
Here wm is the weight of the mass penalty with respecto to the other penalties.
Initially it was 300 and adjusted during the iterations. The mass flow m1 and m2 are
the values for splitter blade pressure side and suction side respectively.

Δmass penalty
The difference between the mass flows of each flow channel (divided by the splitter
blade) is penalized. Here wm2 is the weight of the penalty.
2
 m  m2  (3)
Pm 2  wm 2  1 
 m1  m 2 

Efficiency penalty
The difference between the required efficiency (always 100%) and the actual
efficiency is penalized. The efficiency used in the penalty function is the total-to-
total adiabatic efficiency. Here weff is the weight of the penalty.

Peff  weff |  tt   req | (4)

Distortion-skew penalty
The diffuser exit flow is desired to be as uniform as possible. Any distortion or skew
of the radial velocity profile is penalized.

68
Pdis  skew  wdis | d  1 | wskew | s | (5)

Here the distortion d and skew s are defined as

2 Vmid (6)
d
VR  VL

2 (V R  V L ) (7)
s
VR  VL
The radial velocity profile at the diffuser exit is considered. The velocity at the
diffuser exit mid-channel is Vmid, near the hub is VR and near the shroud is VL.

Loading penalty
The loading difference between the splitter and full blade is penalized.

2 ( AB  AS ) (8)
PL  wL
AB  AS
AB and AS are the full blade and splitter blade loading. The loading is calculated as
the area between the suction and pressure side isentropic Mach numbers along the
normalized blade length.

Negative loading penalty


It is desired to have a positive loading everywhere. The negative loading penalty
penalizes the reversed loading due to higher isentropic Mach numbers at the
pressure side of the blade. The negative loading is the area where the suction side
isentropic Mach number is below the pressure side isentropic Mach numbers.

PNL  wNL A (9)

Mach number peak penalty


It is desired to have a smooth acceleration near the leading edge. The Mach number
peak near the leading edge is penalized.

2 ( M P  M min ) (10)
Pmach peak  wmach peak
M P  M min
The peak Mach number Mp is searched near the leading edge. The minimum Mach
number Mmin is also confined to that interval.

Mach number acceleration penalty


It is desirable to have a smooth deceleration in compressors. Any reacceleration
after the peak is penalized.

2( M i  M i 1 )
Pmach acc  wmach acc max( ,0) (11)
xi  xi 1
The summation in eq. 11 is the first derivative of the Mach number and it is
performed after the location of the peak Mach number up to approximately the
middle of the blade, avoiding high Mach numbers near the trailing edge.

Mach number deceleration penalty


It is desirable to have a smooth deceleration. Any change in deceleration gradient
after the peak is penalized.

69
  ( M i  M i 1 ) ( M i 1  M i )  
Pmach dec  wmach dec max    ,0 (12)
  xi 1  xi xi  xi 1  

The summation in eq. 12 is performed downstream the location of the peak Mach
number and it is confined to approximately half the blade.

3. OPTIMIZATION RESULTS

In this study two optimized impellers were designed by the same optimization
procedure. The only difference between both sduties is the blade number limitation.
In the first optimization (OPT1) the minimum blade number is 5+5. In the second
optimization (OPT2) the minimum blade number is 4+4. The maximum blade
number is 6+6 in both cases. Table 1 shows the specifications of the optimum
design found in both optimization studies compared to the baseline impeller. Figure
4 show the optimum design impellers and the baseline impeller. Figure 5 show the
meridional contour of the optimum design impellers with the baseline impeller.

Table 1: Specifications of optimum


design impellers

Baseline
OPT1
OPT2

OPT1 OPT2 Baseline


Figure 4: Optimum design impellers Figure 5: Meridional contour
and baseline impeller

As described later, OPT1 has 0.5-1.5% higher efficiency with slightly wider
operating range compared to the baseline impeller. On the other hand, OPT2 has a
twice as wide operating range with very suitable characteristic operating curve with
a sufficient negative pressure gradient, at a 1% lower efficiency compared to the
baseline impeller.

3.1 Optimised impeller 1 (OPT1)


Figure 6 shows the convergence history and penalty breakdown of the design
iterations. OPT1 was obtained at iteration 17. This impeller is one of the best ten
designs ordered by the NS total to total efficiency.

Figure 7 shows the comparison of the shroud Mach number distribution of iteration
17 obtained by ANN and NS calculation. A general good agreement is observed.
Moreover, the Mach number distribution is very smooth. A good split of blade
loading between the full blade and the splitter blade is also obtained. The best
impeller is not chosen only for its high efficiency. The Mach number distribution, off-
design performance and mechanical considerations play as well an important role in
the selection process. Consequently the authors selected the iteration 17 as the
best impeller(OPT1) because it has a good Mach number distribution and equivalent

70
blade loading between the full blade and the splitter blade with a preferable blade
geometry for manufacturing, almost radial leading edge of the splitter blade. The
blade number of OPT1 is 5+5. OPT1 has larger inlet blade height and smaller exit
blade width compared to the baseline impeller.

150 1.5
NS Efficiency penalty NS Full blade
NS Mass penalty NS Splitter blade
NS Loading penalty ANN Full blade
ANN Splitter blade
100 1.0

Mach number
Penalty

50 0.5

0 0.0
0 10 20 30 40 50 60 70 0.0 0.2 0.4 0.6 0.8 1.0
Iterations Nodimensional meridional length
Figure 6: NS and ANN FigureFigure
7: Shroud Mach number
7: Shroud Mach
penalty breakdown distribution
numberof OPT1
distribution
f OPT1
3.2 Optimised impeller 2 (OPT2)
In Figure 8 the convergence history of total to total efficiency is shown. Similar to
the selection strategy of OPT1, iteration 46 was selected as best impeller (OPT2) in
this optimization run. Iteration 46 is one of the best designs and has the third
highest total to total efficiency shown in Figure 8. Figure 9 shows the shroud Mach
number distribution of iteration 46 obtained by ANN and NS calculation. Similar to
OPT1, iteration 46 was selected as best because it has not only high efficiency but
also a good Mach number distribution and an equal loading split between full blades
and splitter blades as shown in Figure 9. The number of blades is 4+4. The inlet
blade height is slightly smaller. Axial length is smaller than that of baseline impeller.
A remarkable feature of OPT2 is the forward inclined leading edge of the splitter
blade.

1.5
NS Full blade

1%
NS Splitter blade
ANN Full blade
Total to total efficiency

ANN Splitter blade


1.0
Mach number

0.5

0.0
0 10 20 30 40 50 60 70 0.0 0.2 0.4 0.6 0.8 1.0
Iterations Nondimensional meridional length
Figure 8: Efficiency history Figure
Figure 9: Shroud
9: Shroud MachMach
number
number
distribution ofdistribution
OPT2

71
4. PERFRORMANCE TEST RESULTS AND DISCUSSIONS

The optimal shapes resulting from both optimization processes have been further
analyzed both experimentally and numerically. Performance tests have been
conducted using the compressor inside an automotive turbochager. Both optimal
impellers are each powered by a turbine driven by heated air or exhaust gases,
representing engine like conditions. Each compressor volute has the same geometry,
as well as same exit diameter of the vaneless diffuser. Each diffuser has a different
width but the ratio of diffuser width and impeller exit width is fixed as same value.
The compressor characteristics have been measured and compared with the
baseline impeller.

4.1 Performance test results


Figure 10 shows the comparison of compressor characteristics of OPT1, OPT2 and
the baseline impeller. The compressor total to total efficiency shown in Figure 10 is
normalized by the peak efficiency of the baseline impeller. The flow rate is also
normalized by the reference flow rate. OPT1 has a slightly higher efficiency and
operating range compared to the baseline impeller. On the other hand OPT2 has a
very wide operating range with a 1% decrease of compressor efficiency.

Baseline OPT1 OPT2


Figure 10: Comparison of compressor characteristics

Figure 11 shows the comparison of the peak efficiency for each rotational speed
shown in Figure 10. Compared to the baseline OPT1 has a 0.5% higher peak
efficiency around the total to total pressure ratio of 1.8 and more than 1.0% higher
efficiency at pressure ratios above 2.2. On the other hand, OPT2 has about 1.0%
lower efficiency compared to the baseline impeller.

In Figure 12 the operating range of the optimum designed impellers and the
baseline impeller are compared. The operating range is defined as the flow range
between maximum flow rate and surge limit at each rotational speed and
normalized with the surge limit. In this study the maximum flow late is defined as
the flow rate at which the compressor efficiency drops below 65%. The horizontal
axis of Figure 12 means the pressure ratio at surge limit at each rotational speed.
The operating range of OPT1 is slightly smaller up to pressure ratio 1.9 compared to
baseline impeller. However, the baseline impeller decreases its operating range
rapidly at pressure ratios above 1.9. OPT1 does not have this rapid decrease of its
operating range and remains a wide operating range up to a pressure ratio 2.5.

72
OPT2 has almost equivalent operating range of the baseline impeller at the pressure
ratio below 1.9 but a very wide operating range at pressure ratios beyond 1.9.
OPT2’s operating range has more than doubled at the pressure ratio above 2.2
compared to the baseline impeller. OPT2 has a very gradual decrease of its
operating range with pressure ratio increase. As a result, OPT2 has a suitable
compressor characteristic for turbocharger applications which require a wide and
stable operating range.

FigureFigure 11: Comparison


11: Comparison of
of total Figure 12: Comparison of
total to total
to total efficiency operating range
ffi i

Even though OPT1 and OPT2 have been selected as best impellers which show
higher efficiency compared to the baseline impeller during the optimization
procedure, OPT1 and OPT2 have completely different characteristics as
demonstrated by the performance test. OPT1 has achieved higher efficiency
compare to the baseline impeller as expected. On the other hand OPT2 has lower
efficiency but has a significant increase of its operating range. The difference of the
compressor characteristics of OPT1 and OPT2 shows the diversity and complexity of
the aerodynamic design of a centrifugal compressor impeller. There is a possibility
to get the variety of geometry and characteristic of optimum design such as OPT1
and OPT2 with same design procedure. The discrepancy of OPT2’s efficiency
between the estimation and the test result remains to be solved.

It is obvious that the optimization procedure composed of a GA and an ANN in this


study is beneficial to find high performance impellers in a short amount of time.

4.2 CFD results


The flow phenomena in the optimum designs have been investigated by more
detailed steady NS simulation. The commercial code CFX ver.12 was used for this
study. The whole domain of a compressor stage including the impeller, diffuser and
volute has been calculated. The number of grid cells is about 2,470,000 for the
impeller, 410,000 for the diffuser and 250,000 for the volute. A frozen rotor
interface between the impeller outlet and the diffuser inlet has been applied. The k-
ε model was used as the turbulence model. CFD was conducted at the normalized
rotational speed of 0.89 (maximum speed: 1.0) and the flow rate near the peak
efficiency condition.

In Figure 13 the limiting stream lines on the blade surface, stream lines (only in
right figures of Figure 13) and the entropy distribution and are compared. Also
shown in Figure 13 are the identified vortex cores which are coloured with
normalized helicity. Normalized helicity is defined as cosine of the angle made
between the vortex vector and the velocity vector, and the domain where its
absolute value is 1 indicates a strong rolling-up of a streamwise vortex. Regarding

73
the entropy distribution the regions which have a relatively high entropy are
mapped in streamwise sections in Figure 13.

Low momentum fluids Low momentum fluids


by LE tip leakage vortex by secondary flow and
Accumulation of low momentum fluids and tip leakage flow tip leakage flow
Hn [-] by LE tip leakage vortex
Tip leakage flow Tip leakage flow

Secondary flow
LE tip leakage vortex
Secondary flow
Accumulation of low momentum fluids
by secondary flow
(a) Baseline
Hn [-] Tip leakage flow Tip leakage flow

LE tip leakage vortex

(b) OPT1
Hn [-] Tip leakage flow Tip leakage flow

Secondary flow
LE tip leakage vortex
Accumulation of low momentum fluids
by secondary flow
(c) OPT2

Figure 13: Comparison of limiting stream lines, vortex structure


and entropy distribution

From Figure 13(a), it is clear that the baseline impeller has a strong secondary flow
rolling up from hub to shroud on the suction surface at the inducer. Because of this
strong secondary flow, low momentum fluids start to accumulate on the suction
surface of the full blade near the end of the inducer. This low momentum fluids
combined with the tip leakage flow more downstream the full blade accumulate and
compose the high entropy region near the tip corner of the splitter blade pressure

74
surface at the impeller exit. As shown in Figure 13(a) the LE(leading edge) tip
leakage vortex is generated at just downstream the full blade leading edge. This LE
tip leakage vortex moves into the passage of the splitter blade suction surface and
composes a low flow region near the shroud. This LE tip leakage vortex travels
further downstream while accumulating low momentum fluid and composes the
high loss region near the shroud of the splitter blade suction surface at the impeller
exit. At the tip corner of the full blade pressure surface at the impeller exit mostly
tip leakage flow downstream the inducer accumulates and highest loss region is
composed. As demonstrated above the strong secondary flow motion and tip
leakage flow are the main causes of loss generation in the baseline impeller.

In contrast to the baseline impeller OPT1 does not have a remarkable secondary
flow at the inducer as shown in Figure 13(b). Owing to this, the accumulation of low
momentum fluid is suppressed near the tip corner of the splitter blade pressure
surface at the impeller exit compared to the baseline impeller. The LE tip leakage
vortex is remarkable different between the baseline impeller and the OPT1 impeller,
as it only develops more downstream the main blade and postpones its movement
into the passage of splitter blade pressure surface. This is due to a wider blade pitch
distance as a result of the smaller blade number, even though LE tip leakage vortex
is much stronger due to the increased blade loading resulting from the smaller
blade count. Because this stronger LE tip leakage vortex travels into the passage of
the splitter blade pressure surface, the maximum loss of this passage is higher than
that of the baseline impeller. However the area of highest loss region is much
smaller than that of the baseline impeller because there is almost no accumulation
of low momentum fluid by the secondary flow. Regarding the passage of the splitter
blade suction surface, it is obvious that the high loss region is suppressed clearly
compared to the baseline impeller because the LE tip leakage vortex of the full
blade leading edge does not come in and the related loss is diminished. OPT1 has a
very smooth flow without pronounced secondary flow at the inducer and thus
achieves a higher efficiency compared to the baseline impeller. It seems that the
spanwise and streamwise blade loading distribution between the full blade and
splitter blade are optimized to succeed suppressing the secondary flow by the
aerodynamic design optimization system composed of an ANN and a GA in this
study.

The internal flow phenomena and loss generation mechanisms of OPT2 are almost
identical to those of OPT1. In contrast to OPT1, OPT2 has a secondary flow rolling
up from mid-span to shroud on the suction surface at the inducer (similar to the
baseline) as shown in Figure 13(c). Because of this secondary flow, the
accumulation of the low momentum fluid on the full blade suction surface near the
end of the inducer is observed. As a result, OPT2 has a relatively larger area of low
momentum fluid at the exit of the passage of the splitter blade pressure surface and
has lower efficiency compared to OPT1. On the other hand, compared to the
baseline impeller, the secondary flow and the accumulation of the low momentum
fluid are not as severe. It means that the design optimization system has succeeded
to improve the internal flow similar to the design of OPT1. According to the CFD
result, OPT2 has higher efficiency compared to the baseline impeller in spite of
lower efficiency confirmed by the performance test. This discrepancy between CFD
and test results remains to be solved. Significant increase of the operating range
has been confirmed by the performance test of OPT2. The investigations to find the
mechanisms of map width enhancement of OPT2 are beyond the scope of this paper.
The authors have found out the breakdown of the tip leakage vortex plays an
important role in stabilizing the unsteadiness of the flow near the stall condition in a
separate study(14).

75
5. CONCLUSIONS

In this study an advanced optimized design system composed of an ANN and a GA


alternative to a conventional design system has been proposed. The authors have
applied this optimized design systems to the aerodynamic design of centrifugal
compressor impellers for automotive turbochargers. Two impellers have been
designed and their performance has been further studied experimentally and
numerically. The following conclusions are obtained.

(1) Two optimum design impellers, OPT1 and OPT2, were designed by an advanced
optimization system. OPT1 has a higher efficiency with slightly wider operating
range compared to the baseline impeller. OPT2 has a twice as wide operating
range compared to the baseline impeller with a minor decrease in efficiency.
This successful result clearly demonstrates the benefits of advanced optimization
systems composed of an ANN and a GA, and shows that innovative
aerodynamic design can be found while speeding up the design time.

(2) OPT1 has a 0.5% higher peak efficiency at pressure ratio 1.8 and more than
1.0% higher efficiency above pressure ratio 2.2 compared to the baseline
impeller. The operating range above the pressure ratio 1.9 is wider than that of
the baseline impeller. There is almost no secondary flow on the suction surface
of the full blade. It seems the spanwise and streamwise blade loading, and
loading split between the full blade and splitter blade are optimized to suppress
the secondary flow and achieve a higher efficiency.

(3) OPT2’s operating range has increased more than double at the pressure ratio
above 2.2 with almost equivalent operating range at low pressure ratio and 1%
lower efficiency compared to the baseline impeller. The internal flow phenomena
and loss generation mechanisms of OPT2 are almost identical to those of OPT1.
Owing to a slightly stronger secondary flow compared to OPT1, OPT2 has lower
efficiency. Significant increase of the operating range was investigated in a
separate study. The breakdown of the tip leakage vortex plays an important role
in this.

6. REFERENCES

(1) Osako, K., Jinnai, Y. Samata, A., Suzuki, H., Ibaraki, S., Hayashi, N., 2006,
“Development of the High Performance and High Reliability VG Turbocharger
for Automotive Applications”, MHI Technical Review, Vol. 43, No. 3.
(2) Fisher, F. B., 1988, “Application of Map Width Enhancement Devices to
Turbocharger Compressor Stages”, SAE Paper 880794.
(3) Tomita, I., An, B. And Nanbu, T., 2014, “A New Operationg Range
Enhancement Device Combined with a Casing Treatment and Inlet Guide
Vanes for Centrifugal Compressor”, IMechE, 11th International Conference on
turbochargers and turbocharging.
(4) Ibaraki, S., Ogita, H. and Yamada, T., 2007, “Development of a Wide
Operating Range Turbocharger Compressor with a Low Solidity Vaned Diffuser”,
CIMAC No. 166.
(5) An, B., Shiraishi, T., 2010, “Development of Variable Two-stage Turbocharger
for Passenger Car Diesel Engines”, MHI Technical Review, Vol. 47, No. 4.
(6) Ibaraki, S., Higashimori, H. and Mikogami, T., 1998, “Flow Investigation of a
Centrifugal Compressor for Automotive Turbochargers”, SAE Paper 98-P94.
(7) Ibaraki, S., Higashimori, H. and Matsuo, T., 2001, “Flow Investigation of a
Transonic Centrifugal Compressor for Turbocharger”, 23rd CIMAC.
(8) Ibaraki, S., Matsuo, T., Kuma, H., Sumida, K and Suita, T., 2003,
“Aerodynamics of a Transonic Centrifugal Compressor Impeller”, ASME Journal
of Turbomachinery, Vol.125, No.2,pp. 346-351.

76
(9) Ibaraki, S., Furukawa, M., Iwakiri, K. and Takahashi, K., 2001, “Vortical Flow
Structure and Loss Generation Process in a Transonic Centrifugal Compressor
Impeller”, ASME Paper No. GT2007-27791.
(10) Zangeneh, M., Roduner, D. V. C., 2002, “Improving a Vaned Diffuser for a
Given Centrifugal Impeller”, ASME Paper No. GT-2002-30621.
(11) Arnone A. and Pacciani R., 1995, “Rotor-Stator Interaction Analysis Using the
Navier-Stokes Equations and a Multigrid Method”, ASME Paper 95-GT-177.
(12) Pierret S. and Van den Braembussche R.A., 1998, “Turbomachinery blade
design using a Navier-Stokes solver and Artificial Neural Network”, ASME
Trans. Journal of Turbomachinery, Vol. 121, No. 9, (pp. 326-332).
(13) Alsalihi Z. and Van den Braembussche R.A., 2002, “Evaluation of a Design
Method for Radial Impellers Based on Artificial Neural Network and Genetic
Algorithm”, ASME ESDA 2002/ATF-069, Istanbul.
(14) Tomita, I., Ibaraki, S., Furukawa, M., Yamada, K., 2012, “The Effect of Tip
Leakage Vortex for Operating Range Enhancement of Centrifugal Compressor”,
ASME Paper, GT2012-68947.

77
Inlet recirculation in automotive
turbocharger centrifugal compressors
P X L Harley, S W T Spence, J Early
School of Mechanical & Aerospace Engineering, Queen’s University Belfast, UK
D Filsinger, M Dietrich
IHI Charging Systems International GmbH, Germany

ABSTRACT

As the designers of modern automotive turbochargers strive to increase map width


and lower the mass flow rate at which compressor surge occurs, the recirculating
flows at the impeller inlet are becoming a much more relevant aerodynamic
feature. Compressors with relatively large map widths tend to have very large
recirculating regions at the inlet when operating close to surge; these regions
greatly affect the expected performance of the compressor.

This study analyses the inlet recirculation region numerically using several modern
automotive turbocharger centrifugal compressors. Using 3D Computational Fluid
Dynamics (CFD) and a single passage model, the point at which the recirculating
flow begins to develop and the rate at which it grows are investigated. All numerical
modelling has been validated using measurements taken from hot gas stand tests
for all compressor stages. The paper improves upon an existing correlation between
the rate of development of the recirculating region and the compressor stage,
which is supported by results from the numerical analysis.

NOMENCLATURE

‫ܣ‬ Area Subscripts


‫ܤ‬ Blockage (-) ܾ Blade
‫ܥ‬ Absolute velocity (m/s) crit Critical
‫ܦ‬ Diameter (m) ݉ Meridional
DF Disk Friction max Maximum
݉ሶ Mass flow rate (kg/s) 0 Stage inlet
OP Operating Point 1 Impeller inlet
PR Pressure Ratio (-) ߠ Tangential
ܷ Blade velocity (m/s)
ܸሶ Volumetric flow rate (m3/2)
VL Volute Loss
ߚ Relative angle to meridional (deg)
ߟ Efficiency Total-Total (-)
ߩ Density (kg/m3)
߶଴ଵ Stage flow coefficient (-)

_______________________________________
© The author(s) and/or their employer(s), 2014
89
1. INTRODUCTION

A historic problem with turbocharged passenger cars was turbo lag. Modern
automotive turbocharger centrifugal compressors are designed to provide boost as
quickly as possible when power is demanded to improve driveability at low engine
speeds. It would be ideal to have instant boost and hence instant torque, but due
to inertial effects this is not possible. Reducing the inertia of the rotating
components is a continual design target; however the spin up rate of the
compressor can be limited by the surge margin of the stage. Typical centrifugal
compressor performance dictates that the surge mass flow rate increases with
rotational speed; during the spin up the compressor must not operate below the
surge mass flow rate, thus avoiding the unstable surge region of the map. New
design methods have reduced the mass flow rate at which compressor surge occurs
while maintaining the flow range from surge to choke. Some ways of achieving this
are by applying backsweep to the impeller trailing edge and increasing the inlet-to-
outlet radius ratio, also known as ‘trim’. The result is an improved surge margin,
although the drawback is significant recirculation at the inlet to the impeller driven
by adverse pressure gradients along the shroud, inlet recirculation is addressed in
this paper at steady state operating conditions.

Recirculation has a very notable effect on the compressor stage performance, most
notably in the pressure ratio map. The recirculating flow causes an aerodynamic
blockage which affects the incoming flow to the compressor. The effect on the
pressure ratio map is evident in Figure 1 where a deviation from the trend line is
seen at the mass flow rate that inlet blockage begins. Inlet recirculation blockage in
this case is defined as the percentage of the inlet area that does not contribute to
the stage mass flow rate as shown by Figure 2. The blockage caused by the
recirculating flow continues to grow in size toward the surge region of the map.

Figure 1 CFD compressor map C-1 showing contours of impeller


inlet blockage (%) as a result in recirculation

90
Figure 2 Typical meridional streamlines showing impeller inlet
recirculation

Typically the recirculating body of fluid fills a portion of the meridional passage
similar to that shown in Figure 2. The recirculation zone is almost self-sustaining
with regard to mass flow. The stage mass flow is very similar to the mass flow rate
passing through the unblocked active flow entering the impeller, normally within a
couple of percent based on recent numerical studies.

This study investigates the impeller inlet recirculation for three automotive
turbocharger centrifugal compressors, henceforth referred to as C-1, C-2 and C-3.
The flow feature is analysed qualitatively and shows how the size of the inlet
recirculation varies predictably with respect to stage geometry.

2. MODELLING

Due to the difficulty in measuring the size of the recirculating region via
experiment, CFD (Computational Fluid Dynamics) was used to analyse the flow
feature numerically. The geometry and mesh used in the simulation of the three
stages have been developed so as to ensure comparability. A single blade passage
was simulated using the ANSYS CFX RANS solver and the geometry prepared using
the dimensions shown in Figure 3.

The inlet domain was made long enough so as to ensure that even at the most
extreme operating condition the inlet recirculation did not cross the inlet boundary.
The inlet and vaneless diffuser domains were stationary and impeller domain was
rotating. Frozen rotor interfaces were used between the stationary and rotating
domains.

The CFX SST (Shear Stress Transport) turbulence model was used which required a
maximum y+ (dimensionless wall distance) of less than 5; the majority of the
passage had a y+ of less than 2 at all operating conditions. The stationary inlet and
vaneless diffuser domains were meshed in ANSYS ICEM and had approximately
200k and 100k cells respectively, and the rotating domain was meshed using
Turbogrid and contained approximately 1.2million cells. Total cell count for all three
stages was approximately 1.5million which had been reached through a mesh
independence study.

Convergence was defined when the RMS residuals fell below 1e-4, the imbalances
of mass, energy, and momentum in all domains fell below 0.01%, and the total-
total isentropic efficiency was fluctuating less than ±0.05%.

91
S2
4D

0.5D R1

S1

1mm D z -ve 0

Figure 3 CFD compressor single passage model dimensions

2.1 1D loss modelling


In order to ensure comparability of the single passage CFD results with test data
(presented later in Section 2.2), certain losses had to be applied to the CFD results.
The main differences between the CFD and the test data is the presence of a
scroll/collector. In the turbocharger compressors used in this study a scroll volute is
the norm and for that reason a volute model had to be applied. The meanline
volute modelling method of Weber & Koronowski [1] was used to model the effect a
volute would have on stage performance.

Another difference between the single passage model and the test data was the
lack of an impeller back disk in the CFD model. The impeller back disk creates
friction with the housing adjacent to it due to viscous shear of fluid in a small
space. To compensate for this, the disk friction (DF) loss of Aungier [2] was used as
it is the most representative of the classic Daily and Nece [3] study.

2.2 CFD validation


The CFD results with and without the 1D loss models applied are plotted in Figures
4, 5, and 6 against hot gas stand data. All hot gas stand data was collected in
accordance with SAE J1826 [4]. Surge is defined on the test rig using pressure
fluctuations in the compressor discharge line to ensure repeatability. All compressor
maps have been non-dimensionalised using the maximum pressure ratio and mass
flow rate for each complete compressor data set of CFD and test results.

Good pressure ratio and efficiency correlations between the test data and CFD
simulations are shown for C-1 and C-3 in Figures 4 and 6 respectively. Some error
exists in the simulated pressure ratio for C-2 (Figure 5); this is potentially related
to the interaction of the simulated stage with the scroll volute.

92
1.0
20%
0.9

0.8

η/ηmax [-] 0.7

0.6

0.5

1.0
0.4
0.0 0.1 0.2 Test
C-1 0.3
Data0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1

0.9 CFD without VL & DF


CFD with VL & DF
0.8
43%
0.7
PR/PR max [-]

0.6

0.5

0.4

0.3

0.2
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
ṁ/ṁmax [-]

Figure 4 C-1 compressor map comparison

The test data shows a drop in efficiency at low tip speeds which is not supported by
the CFD. The reduction in efficiency is the result of the inescapable heat transfer
inherent with hot gas stand testing as heat is exchanged between the hot turbine
and relatively cool compressor through the bearing housing. The same effect is
present in all three compressor maps (Figures 4, 5, and 6). As expected the largest
of the three compressors, C-3 has the smallest efficiency decrement of only
approximately 7 non-dimensionalised percentage points, whereas C-1 and C-2 have
a similar drop in peak efficiency at the lowest tip speed of almost 20 points.

93
1.0
20%
0.9

0.8

η/ηmax [-] 0.7

0.6

0.5

1.1
0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 C-20.8
Test 0.9
Data 1.0 1.1
CFD without VL & DF
1.0
CFD with VL & DF

0.9
PR/PR max [-]

65%
0.8

0.7

0.6

0.5

0.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
ṁ/ṁmax [-]

Figure 5 C-2 compressor map comparison

The surge margin in the CFD was defined as the mass flow rate that the solver
would no longer converge to the criterion described previously. For C-1, surge at
low and high tip speeds was predicted very well but the error increases from low to
mid tip speeds. The maximum error in the CFD surge mass flow rate prediction for
C-1 is ~43% as shown in Figure 4. For C-2 in Figure 5 considerably more error is
present in the surge margin prediction, this time a maximum of ~65%; again this is
potentially related to the simulation not containing the asymmetric volute found on
the tested compressor. The surge mass flow rate CFD prediction for C-3 in Figure 6
is much better with a maximum error of ~16%. In the C-1 and C-3 simulations the
position of the typical ‘kink’ in the surge margin is captured very well.

94
1.0

0.9
7%

0.8

η/ηmax [-] 0.7

0.6

0.5

1.0
0.4
0.0 0.1 C-3
0.2 Test Data0.4
0.3 0.5 0.6 0.7 0.8 0.9 1.0 1.1
CFD without VL & DF
0.9
CFD with VL & DF

0.8
16%

OP6
PR/PR max [-]

0.7 OP5 OP4 OP3 OP1


OP2

0.6

0.5

0.4

0.3
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1
ṁ/ṁmax [-]

Figure 6 C-3 compressor map comparison

2.3 CFD post-processing


The ‘Turbo Mode’ within ANSYS CFX Post was used to develop meridional plots
which were used to analyse the growth of the recirculation flow paths using
streamwise velocity plots. Also, the ANSYS ‘Turbo Line’ feature was used to extract
relevant data very close to the leading edge of the impeller. All Turbo Line data was
mass flow averaged circumferentially.

To find the extent to which the blockage fills the impeller inlet the axial velocity and
density was extracted at 200 sample points from hub to shroud. The sample points
were distributed so as to allow equal area sample regions. The sample points were
then used to calculate an average flow property for the area contained between the
sample points and a mass flow rate calculated. The mass flow rate recirculating
close to the shroud is considered to be negative. Starting from the shroud, the
mass flow rate through each circular segment was cumulatively summed until zero
mass flow was achieved. The recirculating limit shown in Figure 7 defines the size
of the inlet blockage.

95
1.13 1.14 1.15 1.16 1.17

Out
Cm1 [m/s] (bottom axis)
Recirculation Limit
Density [kg/m^3] (top axis)
Recirculation
span [-]

In

Active Flow

-40 -20 0 20 40
Figure 7 Typical low tip speed low mass flow rate impeller inlet flow
conditions

The velocity profile shown in Figure 7 is similar to that proposed by Qiu et al. [5]
and is therefore supported by the CFD simulations. The variation of spanwise
density in the recirculation zone shows an increase toward the shroud driven by the
radial pressure gradient of the swirling flow. The density toward the hub is
maintained by the incoming active flow.

3. RESULTS AND DISCUSSION

As mentioned earlier the ‘Turbo Line’ used was close to the leading edge of the
impellers as represented in Figure 7. It was realised that the recirculation actually
starts within the impeller, and that using the Turbo Line method as described will
miss the onset of recirculation due to the current definition (Figure 7). To ensure
that the method was valid a series of meridional plots were used to judge the
starting position of the recirculation and compare this to the blockage seen at the
Turbo Line.

Based on the knowledge developed the Qiu et al. [5] method is improved upon to
provide a more accurate estimation of blockage for an automotive turbocharger
style compressor.

3.1 Recirculation initiation


In order to demonstrate the growth of the recirculation zone a series of meridional
plots are shown for compressor C-3. Figure 8 shows meridional plots of streamwise
velocity at six operating points (OP) with the contours limited to show positive
(white) and negative (black). The six operating points in question are plotted in
Figure 6 for reference.

As is clear from Figure 8, the recirculation starts along the shroud surface within
the blade passage. The recirculation seen in OP2 (operating point 2) is driven by
the adverse meridional pressure gradient at the shroud, but it is encouraged by the

96
tip leakage flows. The dotted line in Figure 8 represents the position of the Turbo
Line placed in ANSYS CFX. At OP2 the recirculation zone does not actually cross the
Turbo Line and therefore the post processing method outlined earlier will not
register this as an operating condition with blockage. As expected, with further
reductions in mass flow rate the recirculation zone grows, and hence the blockage
grows. The recirculating zone extends not only in the spanwise direction, but also in
the axial direction (although this is not investigated here) allowing the Turbo Line to
detect the recirculation.

OP1 OP2 OP3

OP4 OP5 OP6

Figure 8 Meridional plots showing regions of negative streamwise flow


(see Figure 6 for corresponding operating points)

As is outlined by Qiu et al. [5] there is also recirculation at the trailing edge of the
impeller. This recirculation zone is not analysed here, although the same process
could be used to measure and quantify the outlet recirculation and blockage.

3.2 Blockage correlation


In a recent paper, Qiu et al. [5] produced a blockage correlation based on a critical
area ratio. It is based on the ‘Two Element In Series’ model of Japikse [6]. It
assumes that the maximum diffusion area ratio that can be maintained stably is
1.5; hence the recirculation fills a portion of the passage and maintains the ratio of
1.5 in the active flow region. Blockage is defined in this model by Equation 1 [5].

=1− − 1 [1]
( − ) cos ( )
where

cos
=
cos

In order to analyse this model the CFD results were used to calculate the critical
area ratio required to provide the inlet blockage seen in the results. In order to
compare the critical area ratio for all operating conditions and all compressors, the
inlet stage flow coefficient was used as given by Equation 2.

97
= [2]

The results in Figure 9 show that a constant value of critical area ratio is not
supported by the CFD simulations. Instead a trend is shown to increase in a
parabolic shape toward lower flow coefficients. All compressors appear to fall on to
this common trend line. The recirculation starts at an area ratio of approximately
1.3 as opposed to the 1.5 value Qiu et al. [5] suggests.

The data in Figure 9 was used to further develop the Qiu et al. [5] model by
correlating the critical area ratio against the flow coefficient; the result is Equation
3.

= 160 − 25 + 2.2 [3]

>0

The current correlation is plotted on Figure 9 and is only valid when the calculated
inlet blockage is greater than zero. The correlation can then be used to compare
the CFD blockage against the blockage now predicted using the Qiu et al. [5] model
with a varying critical area ratio.

With a varying critical area ratio, the improved Qiu et al. [5] model is shown in
Figure 10 to have very good agreement with the CFD simulation. The original Qiu et
al. [5] model is shown in their paper to have good general agreement with a range
of impeller types, although the target here was a model that worked well with
automotive turbocharger style centrifugal compressors.

2.4

2.2 C-1
C-2
C-3
2.0 Qiu
Current
ARcrit [-]

1.8

1.6

1.4

1.2

1.0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

ϕ01 [-]
Figure 9 Critical area ratio required to provide blockage from CFD using
Qiu et al. [5] inlet recirculation method

98
0.6

C-1
0.5 C-1 (Qiu)
C-1 (Current)
C-2
C-2 (Qiu)
0.4
C-2 (Current)
C-3
C-3 (Qiu)
B1 [-]

C-3 (Current)
0.3

0.2

0.1

0.0
0.00 0.05 0.10 0.15 0.20 0.25

ϕ01 [-]
Figure 10 Comparison of CFD inlet blockage with Qiu et al. [5] model and
current

4. CONCLUSIONS

Recirculation is a flow feature that has a significant impact on compressor


performance. Recirculation can exist at the inlet and exit of the impeller although
this study focuses on the inlet recirculation.

Pressure ratio trends in compressor test data can show a deviation from the
expected trend when operating at relatively lower mass flow rates. CFD results are
used to show that this coincides with impeller inlet recirculation which in turn
causes an aerodynamic blockage. The result is two distinct regions of flow; an
active flow region and a recirculating zone.

The recirculating flow region is almost self sustaining with regard to mass flow rate,
and the mass flow rate passing through the active flow region is very similar to the
stage flow rate. For this reason it could be assumed that the flow passing through
the active flow region wholly contributes to the flow conditions at the impeller
outlet.

CFD simulations are carried out for three automotive turbocharger centrifugal
compressors and the data used to analyse the rate at which the inlet blockage
grows. The results are compared with a recent blockage prediction model by Qiu et
al. [5]; the CFD does not support this model. The weakness of the Qiu model was
identified to be the assumption of a constant critical area ratio.

99
The critical area ratio was analysed and correlated against a flow coefficient. The
Qiu et al. [5] method was still used but with a varying critical area ratio. The new
correlation is shown to be supported by the CFD results.

ACKNOWLEDGMENTS

The authors would like to thank IHI Charging Systems International for their
technical support and provision of the required compressor geometry and test data.
Thanks are also given to ANSYS Inc for the use of their software for numerical
modelling in this research programme.

REFERENCES

[1] Weber, C.R., & Koronowski, M.E., 1986, “Meanline Performance Prediction
of Volutes in Centrifugal Compressors”, ASME Conference Proceedings, New
York, USA, 86-GT-216.
[2] Aungier, R.H., 2000, “Centrifugal Compressors: A Strategy for Aerodynamic
Design and Analysis”, ASME Press, ISBN 0-7918-0093-8, New York, USA.
[3] Daily, J.W., & Nece, R.E., 1960, “Chamber dimension effects on induced
flow and frictional resistance of enclosed rotating disks”, ASME, Journal of
Basic Engineering, Volume 82, New York, USA.
[4] SAE J1826 March 1995 “Turbocharger Gas Stand Test Code”, Pennsylvania,
USA.
[5] Qiu, X., Japikse, D., & Anderson, M., 2008, “A Meanline Model for Impeller
Flow Recirculation”, Proceedings of ASME Turbo Expo 2008: Power for Land,
Sea and Air, Berlin, Germany, GT2008-51349.
[6] Japikse, D., 1996, “Centrifugal Compressor Design and Performance”,
Concepts NREC, ISBN 0-933283-03-2, Vermont, USA.

100
Modelling of turbocharger heat transfer
under stationary and transient engine
operating conditions
R D Burke1, P Olmeda2, F J Arnau2, M Reyes-Belmonte2
1
Department of Mechanical Engineering, University of Bath, UK
2
CMT – Motores Térmicos, Universitat Politècnica de València, Spain

ABSTRACT

A lumped capacity heat transfer model has been developed and compared to
measurements from a turbocharger operating on a 2.2L Diesel engine under steady
and transient conditions ranging from 1000-3000rpm and 2-17bar BMEP. The
model parameters have been estimated based on similar devices and this study
quantifies the errors associated with this approach. Turbine outlet gas temperature
prediction was improved with RMSE reduced from 29.5oC to 13oC. A sensitivity
study showed the parameters of the heat transfer model influence gas
temperatures by only ±4oC but housing temperatures by up to 80oC. Transient
simulations showed how errors in the thermal capacitance also lead to errors. This
study shows the importance of undertaking a full thermal characterisation and the
need for accurate adiabatic maps in turbocharger simulations.

1 INTRODUCTION

The accuracy of turbocharger models needs to increase to allow more use of


simulation tools in engine development. Typically heat transfer is ignored in
turbocharger models and these are map based, derived from steady state
measurements taken on steady flow facilities. A number of studies into the effect of
heat transfer have been published over the past 10 years (1-5). The major findings
from these investigations are that the heat transfer has a limited effect on the flow
behaviour of the turbocharger turbine and compressor, but has a significant
influence on the outlet temperatures of the device. On the turbine side heat
transfer can represent 20-90% of the total enthalpy drop. Most of this heat is then
lost to ambient with the remainder transferred along the bearing housing, primarily
to oil and cooling water if present. On the compressor side, the heat transfer is
more complex as unless the turbocharger is insulated, heat transfer from the
turbine to the compressor is small (6). On the gas stand, as temperature are used
to measure work, this gives poor estimates of the aerodynamic performance. When
these maps are used on the engine this leads to poor estimation of operating
speed, gas temperatures and air flows.

Various models have been proposed to account for heat transfer each employing
the assumption that work and heat transfer occur independently in the compressor
and turbine. The flow process is therefore split into one or two heat transfers at
constant pressure and an adiabatic compression/expansion. Shaaban (3) and
Romagnoli and Martinez-Botas (5) provide an analytical solution to heat transfer in
the bearing housing simplifying the geometry to a series of cylinders and fit results
for an individual device for other model unknowns. Other authors (7, 8) use the

_______________________________________
© The author(s) and/or their employer(s), 2014
103
lumped capacitance method, splitting the turbocharger housing into a number of
thermal nodes linked through convection, conduction and radiation. In this
approach, the difficulty lies in determining the thermal resistances between the
different nodes. Serrano et al (9) propose a method to calculate conductivities and
capacitances using the turbocharger housing as a 1D heat flux probe. They then
compare adiabatic, non-adiabatic/externally insulated and non-adiabatic/non-
insulated maps to derive the internal convection and external heat losses. In both
cases, significant experimental effort on dedicated test facilities is required to
parameterise the models. If correctly parameterised, these models offer the
potential to reduce exhaust gas temperature prediction errors from +40oC to ±10oC
(6). However, in this work the model is applied to a turbocharger operating on an
engine system where detailed model parameters were not available.

2 MODEL DESCRIPTION

2.1 Model Overview


An overview of energy flows in turbochargers is shown in figure 1 (a). In addition to
work, heat is also transferred between turbine and compressor. This heat can flow
to ambient or along the bearing housing towards the cooling oil and further ambient
losses. Depending on the operating conditions, heat may continue to be conducted
to the compressor housing or, under high compression ratios, this heat flow can
reverse.

Turbine Compressor

Wf

W Wc
T

QG/T QT/B QC/A


QB/C

Q
B/Oil
QB/Amb
QC/Amb
QT/Amb

(a) (b)
Figure 1: (a) Work and heat flows in the turbochargers and
(b) lumped capacity thermal model

Several studies have demonstrated that radial temperature distribution in a cross


sectional area is negligible compared to the axial temperature distribution (10).
That evidence allows the simplification of heat transfer problem inside the
turbocharger considering it as a one-dimensional problem (9) instead of the three-
dimensional case. Figure 1 (b) shows the proposed 1-D lumped model to account
for heat transfer effects inside an automotive turbocharger. It includes five metal
nodes that are representative for the geometry of the turbocharger. These are T
(turbine case), C (compressor case) and three nodes for the turbocharger bearing
housing named H1 (placed near turbine case), H2 (placed in the central part, where
oil and water comes to the turbocharger) and H3 (placed near compressor case).
That division is justified by the high temperature gradient across the housing.
Working fluids in turbocharger model have been named from here on as Gas
(exhaust gases moving the turbine), Air (compressed air by the compressor), Oil
(lubricating oil) and Ambient. In order to simplify the complex heat transfer

104
phenomena in the turbine side and taking into account the exposed areas, it has
been assumed that the heat transfer phenomena occurs before the expansion
process as it is shown in the left h/s diagram of figure 1 (b). The opposite is true
on compressor side and it has been assumed that heat transfer phenomena occurs
after the compression work.

Metal nodes are connected by means of conductive conductances (Ki,j) and the heat
flow between two nodes is calculated using Fourier's Law (equation 1). Conductive
conductances are represented with thermal resistors drawn in black in figure 1 (b).
Conductive conductances between adjacent metal nodes in the turbocharger are
constant for any operative condition since that property depends only on the
internal geometry and cases material.

Q icond
,j  Ki , j ·Ti  T j  1

Metal nodes are connected to fluid nodes by means of convective conductances


(hAl,i) and heat transfer is calculated using Newton's cooling law (equation 2).
Connection between metal and fluid nodes is represented by bold grey thermal
resistors in figure 1 (b). An additional term for radiation heat transfer is included for
the heat transfer between nodes and ambient taking into account the view factors
between nodes from simplified geometries (11) (light grey resistors in figure 1 (b)).

Q lconv
,i  hAl ,i · Tl  Ti  2

Metal nodes can store energy during transient processes as their temperature
increases or decreases due to their associated mass. That phenomenon has been
taken into account in the network model by means of capacitors (represented in
light grey). Heat flows in the model are summarised by the following process:

 A proportion of exhaust gases energy is transmitted to the turbine case


(node T) before the turbine stator and rotor. This energy (QGas/T) reduces
the temperature governing the expansion process.
 From node T, heat is conducted to node H1 (QT/H1) or transferred to
ambient (QT/AMB), or via radiation to nodes H1, H2, H3 and C.
 At node H1, heat flux can be transmitted following these possible paths:
conduction to node H2 (QH1/H2), forced convection to oil (QH1/Oil); to
ambient (QH1/AMB); or radiation with nodes T and C.
 Node H2 is similar to node H1 for non-water-cooled devices and similar
heat flow paths are defined (QH2/H3, QH2/Oil, QH2/AMB and radiation).
 Housing node H3 in connected by conduction to node C (QH3/C), to ambient
(QH3/AMB) and nodes T and C by radiation.
 Compressor node (C) is connected to ambient (QC/Amb), radiation with
nodes T, H1, H2 and H3, and by forced convection with a node placed at
compressor diffuser (QC/Air).

It is worth noting that for some operative conditions, the temperature after
adiabatic compression (node Air) is lower than node C (low loads): in this case the
compressed air received heat (QC/Air). At higher loads, the temperature after
adiabatic compression can exceed the temperature of node C and this heat flow
reverses. Depending on the relative temperatures and thermal resistances
connecting nodes C, H3 and Ambient, heat flow can also reverse in the bearing
housing.

The oil temperature rises both as a result of heat transfer and frictional losses. It is
assumed that heat is first exchanged by forced convection between nodes H1 and
oil (QH1/Oil); secondly, the oil temperature rises due to frictional losses (Wm); Finally,
further heat is transferred by forced convection with node H2 (QH2/Oil).

105
2.2 Model parameterisation
The turbocharger thermal model contains a number of parameters that are not
easily determined by inspection of the geometry without considerable simplification
(5 capacitances, 4 conductances and 7 convective correlations). Although an
experimental procedure has been determined to characterise these parameters for
an individual turbocharger, the aim of this work is to assess the predictive
capability of the model for a new turbocharger, based on the parameters
established previously for a similar device. A range of turbocharger sizes has
previously been measured and simple correlations have been observed between
external geometries and thermal parameters: these have been used to calculate
the parameters in this study, this replicating an industrial scenario.

3 APPROACH Tcentre
T15mm
3.1 Experimental Setup and Test Points
A variable geometry turbocharger has been
instrumented to capture fluid and structure
T10mm
temperatures at each of the locations described
by the nodes of the model. For all gas a) Gas temperature
measurements in ducts
temperatures, multiple thermocouples were used
to capture distributions across the cross section
of the ducts (figure 2a). For the compressor and
Tneu
turbine housings, temperatures were measured
at 3 azimuths and 2 depths to provide E
information about temperature distributions n
Tamb
g
(figure 2b). The turbocharger was installed on a
i
2.2L Diesel engine for which the usual application Teng n
is a light commercial vehicle. The engine was e
operated on an AC transient dynamometer. Mass
flow was measured at engine intake using an
ABB Sensyflow hot wire flow meter (accuracy
b) Compressor/Turbine
<1%); Pressures were measured using Kistler Scroll
Piezo-Resistive sensors (accuracy 0.5%); gas
and metal temperatures were measured using Figure 2: Thermal
1.5mm k-type thermocouples and turbocharger instrumentation of
speed was measured using an eddy current blade turbocharger
count device from Micro-Epsillon.

20
BMEP (bar)

20 Engine
Transient 10
BMEP (bar)

15 Steady
4000 0
Speed (r/min)

10

5 2000

0 0
1000 2000 3000 4000 0 2000 4000 6000
Brake Speed (rpm) Time (s)

(a) (b)
Figure 3: Engine speed and torque operating points for stably and
transient experiments (BMEP: Brake Mean Effective Pressure)

The engine was operated over under thermally stable and transient conditions.
These are shown with respect to the engine speed/torque map in figure 3 (a). For
steady state tests, the engine was allowed to stabilise for 7 minutes before
recording and measurements represent average behaviour over 60 seconds. The

106
transient experiment consisted of a series of 32 step changes as shown in figure 3
(b) with a hold time of 3 minutes between steps; this tests is superimposed onto
figure 3 (a) in grey.

3.2 Model Control


One-dimensional simulation code OpenWAMTM (12, 13) has been used to create the
turbocharger model with the connecting ducts as shown in figure 4. This highlights
the compressor and turbine models (that are based on the manufacturer’s maps)
connected via a mechanical shaft and the lumped thermal model of the housing.
The turbine map was determined from compressor enthalpy rise and is therefore
adiabatic with respect to turbine performance, however this includes mechanical
losses and was modified by applying the mechanical losses models developed by
Serrano et al. (6, 14). The turbine is represented by a series of two nozzles with an
intermediate reservoir to account for acoustic effects (15, 16).

Figure 4: Turbocharger model boundary conditions for steady


and transient simulations

To avoid inaccuracies of a full engine model, measured conditions have been


imposed at the compressor and turbine inlets (pulsating pressure for turbine). A
back-pressure valve has been introduced at compressor outlet to represent the flow
restriction of the engine; the opening was tuned to give measured compressor
mass flow at measured compressor speed. This approach was preferred to imposing
compressor outlet pressure as it allows simulation errors to occur in pressure and
mass flow. On turbine side outlet pressure was imposed and VGT position was
tuned to match measured mass flow at measured turbine speed. For steady state
the simulations were conducted in two phases:

1) The turbocharger was simulated at fixed speed whilst the flow coefficient of
the orifice at compressor outlet and the turbine VGT positions were
determined to match compressor and turbine flows respectively.
2) The VGT position and compressor outlet orifice were fixed based on the results
from 1; the turbocharger speed was now calculated from the compressor and
turbine power balance.

For the transient experiments, the PID approach is no longer possible as the
dynamics of the controller would interfere with the physical dynamics. To avoid this
problem look up tables were derived from the steady state data and interrogated
via interpolation during transient simulations.

4 RESULTS

4.1 Steady State Conditions


The model accuracy with and without heat transfer simulation are compared in
table 1. In terms of the turbocharger operating speed, pressure ratio and mass flow
predictions there is no significant change in the accuracy of the simulation by

107
adding the thermal model. Turbine outlet temperature prediction is improved
somewhat with the RMSE reducing from 29oC to 13oC whilst compressor outlet
temperature prediction remains similar. The thermal model allows a number of
additional temperatures to be estimated: notably the oil temperature rise in the
turbocharger and housing temperatures to within 10-17oC. The oil temperature rise
Includes both friction and forced convective heat transfers. For the housings,
previous studies by Romagnoli and Martinez-Botas (5) have shown that the
temperature can vary locally by more than the model accuracy when operated on-
engine; therefore discrepancies may be a result thermocouple installation.

Table 1: Root Mean Square Error (RMSE) for steady and transient
simulations, with and without heat transfer (HT) model

Steady State Transient State


Factor
HT OFF HT ON HT OFF HT ON
Turbocharger Speed 4.0% 5.0% 2.7% 2.3%
Compressor pressure Ratio 3.1% 4.5% 5.1% 4.7%
Compressor Mass Flow 2.4% 3.1% 6.1% 6.0%
Compressor Outlet Temperature 7oC 8.4oC 11oC 10oC
Compressor Housing Temperature N/A 17oC N/A 12oC
Turbine Pressure Ratio Imposed
Turbine Mass Flow 2.3% 2.3% 4.2% 4.3%
Turbine Outlet Temperature 29oC 13oC 35oC 14oC
o
Turbine Housing Temperature N/A 13 C N/A 15oC
Oil Outlet temperature N/A 9oC N/A 7oC
Bearing Housing Temperature N/A 11oC N/A 16oC

2500 900 250


Q (T to AMB) Q (BH to Oil) Q (C to Amb)
800 200
2000 Q (T to H1) Q (BH to AMB) Q (C to A)
700
Q (BH to C) 150
600
Heat (W)

Heat (W)

Heat (W)

1500 100
500
400 50
1000
300
0
500 200
100 -50
0 0 -100
1000rpm/50Nm
1500rpm/50Nm
2000rpm/50Nm
2500rpm/50Nm
3000rpm/50Nm

1000rpm/100Nm
2000rpm/100Nm
1000rpm/150Nm
3000rpm/150Nm
1500rpm/200Nm
2500rpm/200Nm
1500rpm/300Nm
2000rpm/300Nm
2500rpm/300Nm
1000rpm/100Nm
2000rpm/100Nm
1000rpm/150Nm
3000rpm/150Nm
1500rpm/200Nm
2500rpm/200Nm
1500rpm/300Nm
2000rpm/300Nm
2500rpm/300Nm

1000rpm/50Nm
1500rpm/50Nm
2000rpm/50Nm
2500rpm/50Nm
3000rpm/50Nm

1000rpm/50Nm
1500rpm/50Nm
2000rpm/50Nm
2500rpm/50Nm
3000rpm/50Nm
1000rpm/100Nm
2000rpm/100Nm
1000rpm/150Nm
3000rpm/150Nm
1500rpm/200Nm
2500rpm/200Nm
1500rpm/300Nm
2000rpm/300Nm
2500rpm/300Nm

(a) Node T (b) Nodes H1/H2/H3 (c) Node C

Figure 5: Predicted heat flows in selected turbocharger nodes

A breakdown of thermal energy flows between selected model nodes is given in


figure 5; heat flows are presented for each node against engine operating point,
ranked as increasing engine power. Considering first node T (figure 5a), the total
heat transfer from exhaust gases increases with torque and power, with more than
half this heat transferring to ambient. The central bearing housing node (figure 5b)
shows that higher heat flow from the turbine housing is primarily absorbed into the
oil. At higher engine powers, the heat flow to the compressor housing reduces
which is also highlighted by the reversal of heat flow between air and compressor
housing shown in figure 5c.

During the simulation, small errors in the operating maps will be compounded as
they will shift simulated speed, mass flow and pressure ratio. A number of possible
sources for these errors are: engine installation effects, operation under pulsating
flows and manufacturing tolerances. Figure 6 shows the proportions of temperature

108
change attributable to work and heat transfer: it is therefore unrealistic to assume
that all simulation errors could be corrected by the heat transfer model.

675 200
625 Heat 180 Heat
575 Expansion 160 Compression
525 140

Tgas (oC)
Tgas (oC)
475 120
425 100
375 80
325 60
275 40
225 20

100rpm/1000Nm
100rpm/2000Nm
150rpm/1000Nm
150rpm/3000Nm
200rpm/1500Nm
200rpm/2500Nm
300rpm/1500Nm
300rpm/2000Nm
300rpm/2500Nm
50rpm/1000Nm
50rpm/1500Nm
50rpm/2000Nm
50rpm/2500Nm
50rpm/3000Nm
50rpm/1000Nm
50rpm/1500Nm
50rpm/2000Nm
50rpm/2500Nm
50rpm/3000Nm
100rpm/1000Nm
100rpm/2000Nm
150rpm/1000Nm
150rpm/3000Nm
200rpm/1500Nm
200rpm/2500Nm
300rpm/1500Nm
300rpm/2000Nm
300rpm/2500Nm
(a) Turbine (b) Compressor

Figure 6: Temperature changes due to compression/expansion


and heat transfer for (a) turbine and (b) compressor

A sensitivity study was conducted to quantify the influence of each parameter in


Table 2. Based on the results in figure 7, the following observations are made:

 Turbine Internal convection has a direct effect on the heat flux between the
gas and the housing (20-30%). This is not directly proportional to heat transfer
coefficient because node T varies ±20-80oC whilst outlet gas temperature varied
up to ±8oC. Heat flow through nodes H1, H2 and H3 is affected causing changes
in node C of ±2oC, but less than 1oC change in compressor outlet temperature.

Table 2: Ranges of sensitivity tested

Factor Variable Variation


Turbine Internal convection hAGAS/T (eq 2) -50% +50%
Compressor internal convection hAC/AIR (eq 2) -50% +50%
External Heat transfer hAX/AMB (eq 2) 0 +100%
Internal Conduction KX/Y (eq 1) -50% +50%

 Compressor Internal Convection caused variation in compressor outlet gas


temperature and compressor housing of up to ±4oC. No significant effect on
turbine temperatures was observed.
 External Heat Transfer had a significant influence on all four temperatures with
gas temperatures and metal temperature varying ±2-14oC and ±5-80oC
respectively.
 Internal Conduction controls the amount of heat conducting between the
turbine and compressor housing. The sensitivity showed a ±19oC and ±7oC
variation in turbine and compressor housing respectively and ±4oC and ±3oC in
compressor and turbine gas outlet temperature respectively.

The study shows that the turbine behaviour is independent of compressor heat
transfer, but that the opposite is not true. Metal temperatures are more sensitive to
model parameters than gas temperatures, especially at high engine powers (with
high compression/expansion ratios). The crucial influence of internal conduction for
estimating metal temperatures demonstrates the need for a rigorous identification
whilst the significant effect of external heat transfer highlights the need for
accurate control and modelling of this. Nevertheless, the use here of the heat
transfer model using parameters with high uncertainty has shown improved
prediction of turbine outlet gas temperature.

109
550
170
500

Tgas comp out (oC)


Tgas turb out (oC) 150
450
130
400
350 110

300 90

250 70
200 50
1 2 3 4 5 1 2 3 4 5
(a) (b)
620 180

520 160

T node C (oC)
T node T (oC)

140
420
120
320
100
220 80

120 60
1 2 3 4 5 1 2 3 4 5
(c) (d)
1 – 50Nm/1000rpm 2 – 50Nm/2500rpm 3 – 100Nm/2000rpm
4 – 300Nm/2500rpm 5 – 300Nm/1500rpm

Figure 7: Model sensitivity to model parameters for (a) turbine gas outlet,
(b) compressor gas outlet, (c) Node T and (d) Node C (Axis refers to
different operating conditions, HT: Heat Transfer)

4.2 Transient conditions


The prediction performance of the “best estimate” model over transient events is
given in table 1 and compared graphically to measured results in figure 8. Overall
the prediction is of similar accuracy to steady state performance, however under
these conditions in addition to the parameters discussed above, the thermal
capacity of each node is also important. A sensitivity analysis has also been carried
out where thermal capacity was varied ±50%: figure 8 shows the results over a
particular step in engine operating point from 280Nm/2000rpm to 60Nm/2300rpm.
As the engine power is reduced, there is a rapid drop in compressor and turbine
outlet temperature associated with the expansion and compression processes.
Subsequently a much slower temperature reduction is observed, associated with
heat transfer and thermal inertia. On the turbine side, ±50% in housing thermal
capacity resulted in ±20-40oC and ±8-11oC in metal and gas temperature
prediction over the transient; on the compressor theses were ±7oC and ±4oC
respectively.

110
(a) (b)
Figure 8: Model sensitivity to thermal capacitance for (a) turbine
and (b) compressor

5 CONCLUSIONS

A study of the predictive performance of a turbocharger heat transfer model has


been undertaken under steady and transient conditions. This showed that turbine
outlet temperature prediction RMSE could be reduced from 29 to 13oC for steady
state conditions whilst other predictions were of similar accuracy. The model
allowed metal temperature to be estimated to within 17oC and 15oC for compressor
and turbine housings respectively. A sensitivity study showed that uncertainty in
the model parameters could affect gas temperatures up to ±4oC and metal
temperatures up to 80oC. Under transient events, the uncertainty associated with
thermal capacity of the nodes increases the errors but this has limited influence on
gas temperatures. The results show that at this stage it is possible to improve the
prediction of turbine outlet temperature on-engine by using the heat transfer model
and inferring properties from a similar device. However the sensitivity study has
suggested that a further accuracy improvement would be expected if a full thermal
characterisation was undertaken.

REFERENCES

[1] M. Cormerais, J. F. Hetet, P. Chesse, and A. Maiboom, Heat Transfer Analysis


in a Turbocharger Compressor: Modeling and Experiments, SAE Paper Number
2006-01-0023, 2006.
[2] J. R. Serrano, C. Guardiola, V. Dolz, A. Tiseira, and C. Cervelló, Experimental
Study of the Turbine Inlet Gas Temperature Influence on Turbocharger
Performance, SAE Paper Number 2007-01-1559, 2007.
[3] S. Shaaban, Experimental investigation and extended simulation of
turbocharger non-adiabatic performance, PhD, Fachbereich Maschinenbau,
Universität Hannover, 2004.
[4] N. Baines, K. D. Wygant, and A. Dris, The analysis of heat transfer in
automotive turbochargers, Journal of Engineering for Gas Turbines and Power,
vol. 132(4), 2010.
[5] A. Romagnoli and R. Martinez-Botas, Heat transfer analysis in a turbocharger
turbine: An experimental and computational evaluation, Applied Thermal
Engineering, vol. 38, pp. 58-77, 2012.
[6] J. R. Serrano, P. Olmeda, A. Tiseira, L. M. Garcia-Cuevas, and A. Lefebvre,
Theoretical and experimental study of mechanical losses inautomotive
turbochargers, Energy, vol. 55, pp. 888-898, 2013.

111
[7] M. Cormerais, P. Chesse, and J.-F. Hetet, Turbocharger heat transfer
modeling under steady and transient conditions, International Journal of
Thermodynamics, vol. 12, pp. 193-202, 2009.
[8] P. Olmeda, V. Dolz, F. J. Arnau, and M. A. Reyes-Belmonte, Determination of
heat flows inside turbochargers by means of a one dimensional lumped
model, Mathematical and Computer Modelling, vol. 57, pp. 1847-1852, 2013.
[9] J. R. Serrano, P. Olmeda, A. Paez, and F. Vidal, An experimental procedure to
determine heat transfer properties of turbochargers, Measurement Science &
Technology, vol. 21, Mar 2010.
[10] S. Shaaban, J. R. Seume, R. Berndt, H. Pucher, and H. J. Linnhoff, Part Load
performance prediction of turbocharged engines, presented at the 8th
International Conference on Turbochargers and Turbocharging, May 17, 2006
- May 18, 2006, London, United Kingdom, 2006.
[11] F. Payri, P. Olmeda, F. J. Arnau, A. Dombrovsky, and L. Smith, External heat
losses in small turbocharger: model and experiments, Submitted to Energy,
2013.
[12] OpenWAM, 2012. Available: www.Openwam.org
[13] J. Galindo, J. R. Serrano, F. J. Arnau, and P. Piqueras, Description of a Semi-
Independent Time Discretization Methodology for a One-Dimensional Gas
Dynamics Model, Journal of Engineering for Gas Turbines and Power-
Transactions of the Asme, vol. 131, May 2009.
[14] J. R. Serrano, P. Olmeda, A. Tiseira, L. M. García-Cuevas, and A. Lefebvre,
Importance of Mechanical Losses Modeling in the Performance Prediction of
Radial Turbochargers under Pulsating Flow Conditions, SAE Int. J. Engines,
vol. 6, pp. 729-738, 2013.
[15] J. R. Serrano, F. J. Arnau, V. Dolz, A. Tiseira, and C. Cervelló, A model of
turbocharger radial turbines appropriate to be used in zero- and one-
dimensional gas dynamics codes for internal combustion engines modelling,
Energy Conversion and Management, vol. 49, pp. 3729-3745, 2008.
[16] M. A. Reyes-Belmonte, Contribution to the Experimental Characterization and
1-D Modelling of Turbochargers for IC Engines, PhD, Departamento de
Máquinas y Motores Térmicos, Universitat Politècnica de València, València,
2013.

112
A 1-D analytical code for the design and
multi-objective optimisation of high-
pressure compressors within two-stage
turbochargers for marine applications
O F Okhuahesogie1, J Stewart1, M J W Riley1, F Heyes2, P Roach2
1
School of Engineering, University of Lincoln, UK
2
Napier Turbochargers Ltd Lincoln, UK

ABSTRACT

This paper describes a novel method of designing the HP compressor of a two-stage


turbocharger using a 1-D code coupled with an optimisation algorithm. The 1-D
code solves relevant thermodynamic and fluid dynamics equations given geometric
variables, the target pressure ratio and the operating speed as inputs. A multi-
objective optimiser is used to return potential designs that will maximise
compressor efficiency and minimise deviation from the target mass flow. These
designs are then taken forward for detailed 3D-CFD analysis. An example based on
a HP compressor impeller and vaneless diffuser is presented to demonstrate the
potential of this 1-D methodology.

NOTATION

Pressure (bar) Local speed of sound (m/s)


Temperature (K) Blade height (m)
Area ( ) Velocity correction coeff.
Efficiency Offset constant
Density ( ) ζ Entropy loss coeff.
Number of cells in grid
Ratio of specific heats
Gas incidence angle(deg)
Specific gas constant/Radius
Inducer throat area
(as used)
Pi
Mass flow rate (kg/s)
Specific heat at const. Subscripts
pressure (J/kgK) 1 Inducer inlet location
Shaft speed (rev/min) 2 Impeller exit location
Mach Number 3 Diffuser exit location
Number of main blades Blade
Total number of blades Flow
Blade velocity (m/s) ℎ Thickness
Gas absolute velocity (m/s) Slip
Gas relative velocity (m/s) − Total to static
Gas absolute angle to axial − Total to total
direction ℎ Hub
Gas relative angle (deg) Shroud
Blade angle (deg) Choke
Slip factor Average
Work input coefficient Relative

_______________________________________
© The author(s) and/or their employer(s), 2014
125
1 INTRODUCTION

International Maritime Organisation (IMO) legislation demands a reduction in NOx


emissions to 20% of the pre 2011 levels by 2016. The development of new, high
pressure-ratio turbochargers is the key enabling technology that will allow marine
engines to meet this target. High pressure-ratio turbocharging can be achieved
using two-stage turbochargers. It is difficult to achieve high pressure ratios in
single stage compression due to limits on material strength (e.g. aluminium) and
the prohibitive cost of more durable materials (e.g. titanium) (1). The compression
ratio achievable over the past 40 years using aluminium for single stage
compression has reached a limit of about 6:1 (2) which is not high enough to meet
IMO III requirements as pressure ratios above 8:1 are more suitable. More so,
when a single stage compressor is designed for pressure ratios of up to 9 (e.g.
when used in gas turbines), the efficiency and map width are significantly lowered
and not sufficient for turbocharger applications.

The development of new turbochargers is challenging in a highly competitive


market place, particularly when the design process relies heavily on the use of
computationally- and time-expensive CFD models. However, the preliminary design
phase of centrifugal compressors often applies a one-dimensional approach to solve
the thermodynamics and fluid equations that describe the internal flow physics.
This computationally inexpensive method gives results that guide the engineer
towards the optimum design space for further exploration with 3-D CFD. Yuri Biba
and Peter Menegay (3) identified two main methods used in the preliminary design
process for centrifugal compressors. In the direct method, performance is
calculated from a given compressor geometry. For the inverse design approach, the
geometry is calculated to meet a required performance. The inverse design
algorithm implemented by Yuri Baba et al (3) calculates the required one-
dimensional geometry and relies on a trusted database of loss models. M. S.
Kamaleshaiah et al (4) implemented a centrifugal compressor performance
algorithm based on a direct prediction method with empiricism for the loss models
and boundary layer growth within the blade channel. This method computes
performance parameters such as stage pressure ratio and efficiency for a given
geometry and inlet conditions.

2 CURRENT STUDY

In this study, a 1-D code is used to calculate centrifugal compressor performance


based on a known parametric geometry at a desired operating point using empirical
correlations without looking into the details of the flow within the components. A
multi-objective optimisation algorithm calls this 1-D code and seeks the most
promising candidates by modifying the parameters that specify the geometry. The
current work describes the configuration of the 1-D code, how the optimisation task
has been phrased for a Differential Evolution for Multi-Objective Optimisation
(DEMO), and presents the results using a test case of the preliminary design phase
of a high pressure centrifugal compressor within a two-stage turbocharger.

2.1 Design optimisation process


Most turbomachinery design processes start from simple 1-D calculations of a
desired operating point and compressor characteristics before a more detailed
design 3D analysis is carried out. The design process discussed in this paper
focuses on 1-D aerodynamic optimisation of the design point of a centrifugal
compressor. Differential Evolution (5) has attained much popularity in the
optimisation community. The variant used here, DEMO, was found to have
comparable performance to other state-of-the-art optimisation algorithms (6). For
this work, the two potentially conflicting objectives are to maximise the efficiency

126
and minimise the deviation from the desired mass flow at the target design speed
line and pressure ratio as shown in figure 1.

2.2 Description of computer program and algorithm


The 1-D code used in this paper was written in C++ for the Linux operating system.
It solves the thermodynamic and fluid dynamics equations in order to describe the
flow physics in a radial compressor impeller and vaneless channel diffuser. The
speed of execution of the code (1 second) on an Intel i5-2520M 2.5 GHz, makes it
feasible for use with multi-objective optimisation algorithms on a regular PC
configuration. The design objective of the 1-D algorithm is to find the geometry of a
compressor that will deliver a required pressure ratio at the design speed line. The
program then calculates the mass flow and efficiency based on an initial geometry.
The user can later modify the geometry manually or automatically using
optimisation algorithm described here until a suitable mass flow and efficiency is
achieved. The algorithm uses a number of iteration loops until error is within 1% as
shown in figure 2.

Figure 1: Design Optimisation Flow Chart Figure 2: Algorithm Flow Chart

2.3 Compressor mathematical modelling

2.3.1 Inducer calculations


At the inducer, the geometric and thermodynamic properties are calculated using
the method described by Whitfield A et al (7). Gas and blade velocity components
are defined at the root-mean-square location. Choke mass flow corrected with a
blockage factor (7) is calculated using equation 1.

2 −1
= 1+ (1)
+1 2

127
At this point, an initial estimate of mass flow (which must be less than choke mass
flow) and static density is used to start the iterative calculation in the inducer. The
other thermodynamic and velocity parameters are then computed. Inducer
blockage is then updated using blockage correlations as described in section 2.3.4.
The estimated density is corrected using the calculated value, the process continues
until the error is less than 1%.

2.3.2 Discharge calculations


A typical velocity triangle at the exit of an impeller with backward swept blade
angle (backsweep angle) is shown in figure 3.

Figure 3: Impeller discharge velocity triangle

The dashed lines (blue and green) represent gas absolute velocity and gas relative
velocity. The black and red lines represent blade and slip velocity respectively.
These dash lines are as a result of slip (relative flow at impeller exit not tangential
to the blade profile) (8). At the discharge, the gas absolute ( ) and relative
( ) flow angles are not known, hence the calculation is started by using an initial
estimate of impeller exit density ( ). The velocity components and thermodynamic
properties at the discharge are then calculated using the equations as described by
Whitfield et al (7). Slip factor , which estimates the deviation of the gas angle from
the blade angle, can be calculated using established correlations, for example those
described by Stodola A. (1927), Stanitz J.D (1952a) and Wiesner F. J (1967) in
(7,9). The 1-D code presented here allows the user/DEMO to select any of the three
slip correlations or a no-slip option. Slip factor is related to slip velocity and blade
speed as shown in equation 2.

=1− (2)

Efficiency is calculated using an equation for peak total-to-total efficiency via


equation 3 where is the entropy or energy loss coefficient (10).

=1− (3)
2 ( − )

The peak efficiency is corrected to obtain the total-to-total efficiency using


correlations described in section 2.3.4 and the density loop is iterated until the
error is less than 1%.

2.3.3 Diffuser calculations


A mathematical model for one-dimensional flow analysis for vaneless diffuser which
accounts for heat transfer effects and wall friction effect in both the linear and
angular momentum equations (11) is implemented in the 1-D code using equations
4 to 8.

1
: + − + =0 (4)

128
: + + =0 (5)

1 1 1 1
: + + + =0 (6)

−1 (7, 8)
∶ = : = +
2

Equations 4 to 8 can be re-arranged to obtain the forms below.

1 1 1 1
− − + − + = 0 = (9, 10)

Equations 4 to 10 were modelled using an upwind finite difference scheme of first


order accuracy (12) on a grid of 40 steps. The velocity and thermodynamic
parameters at the exit of the diffuser are determined at the end of the calculation.
The target total-to-total efficiency can then be calculated using target values and
compared with the calculated total-to-total efficiency described in section 2.3.2.
The calculated value is corrected until the error is less than 1% to end the
efficiency loop. The calculated static pressure at diffuser exit is compared with the
target static pressure and mass flow is varied until the error is less than 0.25%.

2.3.4 Loss models correlations

2.3.4.1 Friction loss


Friction loss in the diffuser is implemented via a friction factor in the radial and
tangential momentum equations describing diffuser flow physics. Friction factor can
be expressed in terms of Reynolds number (7).

2.3.4.2 Impeller blockage


Blockage correlations developed and implemented in the code are based on the
concepts (angle offset, axial length – discharge diameter ratio and velocity
coefficient) as described below.

2.3.4.2.1 Angle offset:


This is based on the idea that the larger the difference between inducer and
discharge blade angles, the more the flow will turn from inducer to discharge.
Larger turning increases flow separation and losses due to blockage. The correlation
is expressed in equation 11.

=1− × − (11)
ℎ 0 ≤ ≤1

The limitation of this idea is that the impeller is a 3-D object and blade angles vary
from inducer to discharge. This variation affects the location where separation may
occur in the impeller. However, validation calculations carried out using angle offset
correlations shows reasonable correlation with CFD results.

2.3.4.2.2 Axial Length – Discharge Diameter Ratio:


The impeller length affects the length of the compressor, the shaft dynamics;
impeller bore stresses and aerodynamic performance (13). It is therefore necessary
to have a balance of axial length and discharge diameter. Birdi (14) suggested a
correlation of the ratio of axial length to discharge diameter for a range of inlet
Mach number according to equation 12.

129
, − (12)
= , + 1−

where = 0.28 and = 0.8. According to Birdi, the equation above yields
values of 0.32 – 0.37 for inlet Mach numbers of 0.9 – 1.2 for typical ranges of
diameter ratios. In this 1-D code, ≤ ≤ , where

= 0.32 = 0.37. The effect of on impeller blockage was


2 2
estimated using equation 13

≤ ≤ ∶ = 1

< _ ∶ = (13)
_ .

.
_
> _ ∶ = .

2.3.4.2.3 Velocity coefficient:


According to Blasius theory for flat plate incompressible flow (15), boundary layer
grows from inlet to outlet and is inversely proportional to Reynolds number or
velocity over distance downstream of the boundary layer as shown in equations 14
and 15.

4.91 0.382
≈ ≈ (14, 15)

where is the distance downstream of the start of the boundary layer and is the
boundary layer thickness. In this 1-D code, Reynolds number is directly
proportional to the gas relative velocity through the impeller blades; hence a model
to account for the effect of gas relative velocity on impeller blockage is modelled
with equation 16 where is a velocity correlation coefficient.

= ,
(16)

The above equation is only valid when , < 1 , otherwise is set to 1 in


the code and choking will cause blockage when , > 1. Reasonable values
are in the range 0.35 ≤ ≤ 0.85. The overall impeller blockage factor is
computed as a product of all the blockage factors.

2.3.4.3 Opening Loss


An opening loss model is implemented in the diffuser to predict pressure drops due
to rapid increase in diffuser area from inlet to exit. The concept is borrowed from
the opening angle correlation of a volute exit cone where pressure loss factor varies
from a value of 0.15 for an opening angle of 100 to 1 for an opening angle of 600
(8). The pressure loss during each step is formulated with the equation of a straight
line using equation 17.

∆ = (0.017 × 2 − 0.02)/ (17)

The opening angle 2 is calculated by first estimating equivalent radius for diffuser
inlet and outlet assuming the diffuser was a cone using equations 18 and 19.

130

= = = tan (18, 19)

The ratio of diffuser outlet to inlet height is included so it has a significant influence
on the opening angle (7).

2.3.4.4 Efficiency Correlations


Peak efficiency calculated in section 2.3.2 is corrected using the impeller exit gas
angle, the rate of diffusion in the impeller using limits suggested in (8) and the gas
incidence angle at the inducer. Loss models were then developed according to the
equations 20 to 22.

= 1 ∶ 65 ≤ ∝ ≤ 80
(20)
.
= ∝ : ∝ < 65 ∝ > 80
.
= ∶ ≥ 0.95
(21)
= 1 ∶ < 0.95

.
= ∶ > 0
(22)
=1∶ < 0

The actual total-to-total efficiency is then calculated by multiplying the peak


efficiency with the corrections.

2.4 Validation of the 1-D code


The 1-D code was originally written to calculate a design point on a compressor
map; however, due to its functionality, it can be used to produce an approximate
compressor map by running the code at a different outlet pressure at the same
speed. As a result, the design point geometry is analysed using the 1-D code for
different slip correlations and compared with ANSYS CFX CFD results (on a mesh of
about 600,000 cells) and experimental data of the same geometry as shown in
table 1. Surge point in the 1-D code was determined to be when the iteration does
not converge or when the gas angle at impeller or diffuser exit exceeds 80o
depending on which situation occurs first (8).

Table 1: Design points comparison

Mass Flow Pressure Total-static


Corrected Ratio efficiency
1-D Code (No Slip) 7.212 2 0.609
1-D Code (Stanitz) 8.961 2 0.728
1-D Code (Wiesner) 9.105 2 0.702
1-D Code (Stodola) 8.993 2 0.725
ANSYS CFD 8.500 2 0.827
Experiment 9.900 2 0.790

131
2.40

Total-to-static efficienccy
0.80
Pressure Ratio
2.20 D0_NoSlip
D0_Stodola
D0_Wiesner
2.00 D0_Stanitz
D0_NoSlip
D0_Stodola 0.60 CFD_old
1.80 D0_Wiesner Exp
D0_Stanitz
1.60 CFD_Old
Exp
1.40 0.40
4 6 8 10 4 5 6 7 8 9 10 11
Corrected Mass Flow Corrected Mass Flow

Figure 4 (a, b): Validation Performance Curves

Table 1 shows that the 1-D code (except no slip correlation) under-estimation of
mass flow is within 10% of CFD and experimental values at design point. There is
more deviation in between the 1-D code and CFD/experiment in estimated
efficiency. Overall, the result shows that the correlations developed here are helpful
in capturing the basic physics. The performance curve in figure 4 shows that the
1-D code using Stanitz and Stodola (7) slip correlations predicts efficiency to within
4% of CFD/experimental values towards the surge region and to a maximum of
around 20% towards choke. The 1-D model was able to predict the width of the
performance curve to within 20% deviation from CFD/experiment measurements.
The most obvious deviation in cases presented is the no-slip correlation model and
as such will not be an option in the optimisation calculations. Also, the 1-D code
was able to predict choke more accurately compared to surge. Any differences
between the 1-D and CFD results are due to extensive use of correlations to
capture certain 3D physics in the 1-D model.

3 RESULTS

DEMO (16) was set up with 75 candidates in a generation and run nominally for
1000 generations. However, by the 600th generation it was clear that the
progression of the Pareto front shown in figure 5 had stagnated as the objective
functions ceased to improve. The optimiser was set to reject candidate geometries
with mass flows not within 20% of target while finding a design as close as possible
to the target mass flow. For this case, the target mass flow was 11.4kgs-1. During
the optimisation, the following criteria were set and penalised if violated:
 Impeller exit blade speed must be less than 580 m/s
 Relative Mach number at inducer shroud must be less than 1.25
 Impeller exit gas angle must be between 65 and 80 degrees.
 Impeller diffusion ratio ( ) must be between 0.25 and 0.95
 Diffuser Opening Angle must be less than 15 degrees
 Minimise deviation of calculated mass flow from target mass
flow(objective)
 Maximise compressor efficiency (objective)

3.1 Optimised design(s)


The current design of the HP compressor is set to operate at a pressure ratio of 2:1
and a target mass flow of about 11.4kgs-1 at 66.31 rads-1 and has an inlet
stagnation pressure and temperature of 350000 Pa and 333K respectively. Table 2
shows the current design and 5 (2 from the 600th generation and 3 from the 300th
generation) new possible improved designs using the 1-D code. In table 2, D0
refers to original design. New designs were selected based on efficiency and
closeness to target mass flow.

132
0.9

Total-Static Efficiency
Gen_10
0.8
Gen_100
Gen_200
Gen_300
0.8 Gen_400
Gen_500
Gen_600
0.7
0 5 10 15
Mass Flow Violation %

Figure 5: Pareto front of generations

Table 2: 1-D results of original and new designs

Figures 6a and 6b compare the 1-D performance curve and speed line of the new
and existing design with CFD and experimental data. All of the new designs have
fairly similar performance, but design D2 gives a reasonable operating range and
higher pressure (about 5% more) near surge compared to the other new designs.
More so, there is a reasonable distance between the peak efficiency mass flow and
choke mass flow in design D2 compared to the other designs. Based on figures 6a
and 6b, design D2 was chosen as the best compromise design.

1.00
2.3
Total-to-static Efficiency

0.80
Pressure Ratio

2.1
D0_Stodola
D0_Stodola D0_Wiesner
D0_Wiesner 0.60 D0_Stanitz
1.9 D0_Stanitz CFD
CFD
Exp
Exp
D1 D1
1.7 0.40
D2 D2
D3 D3
D4 D4
D5 D5
1.5 0.20
3 5 7 9 11 3 5 7 9 11
Corrected Mass Flow Corrected Mass Flow

Figure 6 (a, b): Performance curve of new and old designs

133
1 1.00
2.25
0.8
D0

Pressure Ratio
D2 0.80
0.6 2.00

Pressure Ratio
Radius

0.4
D2_1D 0.60
1.75 D2_1D
0.2 D2_CFD D2_CFD
CFD_Original CFD_Original
0 1.50 0.40
0 0.5 1 4 6 8 10 4 6 8 10
Axial Length Corrected Mass Flow Corrected Mass Flow

Figure 7 (a, b, c): Meridional view and performance curves of new


and old designs

Figure 7 shows good correlation in mass flow calculations at different outlet


pressures. Also, the new design gives slightly higher efficiencies at lower pressures.
Any difference in efficiency estimation is due to the effect of blade curvature on
efficiency which is not captured in the 1-D model.

4 CONCLUSIONS

A computer code has been developed to predict the performance of a centrifugal


compressor at the design point. The code can also be used to predict the
performance at off-design point by changing the outlet pressure at a specific speed
for a given geometry. The code is shown to be accurate enough to be relied upon
during the preliminary design phase of a centrifugal compressor. The code has been
linked with a multi-objective optimiser and ran hundreds of potential geometries to
find the compressor geometry with optimum efficiency at a target mass flow,
pressure ratio and shaft speed with constraints. Validation of the 1D code with a 3D
CFD solver and experimental data revealed that reasonable approximations for the
pressure ratio and efficiency were being returned with a greatly reduced
computational cost. Although the equations used to model the 1-D system do not
capture the full physics of the compressor flow, this work confirms that such a
method can be used to calculate an initial feasible design in the early stages of the
design process of a new compressor. This would ultimately speed up the design
process if less computationally expensive methods described here are used in the
early phase of the design process by reducing the design space to explore during
the more computationally expensive CFD phase.

5 ACKNOWLEDGEMENT

This partnership received financial support from the Knowledge Transfer


Partnerships programme (KTP). KTP aims to help businesses to improve their
competitiveness and productivity through the better use of knowledge, technology
and skills that reside within the UK Knowledge Base. KTP is funded by the
Technology Strategy Board along with the other government funded organisations.

6 REFERENCE LIST

[1] Okhuahesogie O.F, Stewart J, Heyes F. J. G, Roach P.E, Design Optimization of


a two-stage Compressor Impeller, KTP Associate Conference, Brighton, 2012
[2] Codan E., Mathery C. Emissions – A new challenge for turbocharging. Paper
No. 245, CIMAC, Vienna, 2007

134
[3] Yuri B, P Menegay Inverse Design of Centrifugal Compressor Stages Using a
Meanline Approach, International Journal of Rotating Machinery, 10:75-84,
2004
[4] M.S. Kamaleshaiah, N. Venkatrayulu, S. Ramamurthy An Improved Method of
Centrifugal Compressor Performance Prediction, Indian Institute of
Technology, Madras
[5] R. Storn, K. Price Differential Evolution - A simple and efficient adaptive
scheme for global optimization over continuous spaces, International
Computer Science Institute, 1947 Center Street, Berkeley, CA 94704
[6] T. Tusar, B. Filipic Differential Evolution Versus Genetic Algorithms in Multi-
objective Optimization, Department of Intelligent Systems, Jozef Stefan
Institutez, Jamova 39, SI-1000 Ljubljana, Slovenia
[7] Whitfiled A., Baines N.C. Design of Radial Turbomachines, John Wiley and
Sons, 1990
[8] R.A. van den Braembussche. Centrifugal Compressors Analysis and Design,
von Karman Institute for Fluid Dynamics, Course Note 192, February 2012
[9] Wiesner, F.J. A review of slip factor for centrifugal impellers. Journal of
Engineering for Power, Vol. 89, No4, pp 558-572, October 1967
[10] Denton J.D., Loss Mechanisms in Turbomachines, 93-GT-435, ASME, May,
1993
[11] Stanitz J. D., One-dimensional compressible flow in vaneless diffusers of radial
or mixed-flow centrifugal compressors, including effects of friction, heat
transfer and area change. NACA TN 2610, 1952
[12] K. A. Hoffmann, S.T. Chiang Computational Fluid Dynamics, Vol. 1, 4th Edition,
Engineering Education System, 2000
[13] Came P. M., Robinson C J, Centrifugal Compressor Design, Proc. Instn. Mech.
Engrs. Vol. 213, Part C, 1999
[14] Birdi K, Mixed flow compressors, Cranfield University Short Course on
Centrifugal Compressors, 1992
[15] H. Blasius, Boundary layers in fluids with little friction, NACA technical
memorandum 1256, 1950
[16] T. Robic, B. Filipic DEMO: Differential Evolution for Multiobjective Optimization,
Department of Intelligent Systems, Jozef Stefan Institute, Jamova 39, SI-
1000 Ljubljana, Slovenia

135
Improvement of a turbocharger by-pass
valve and impact on performance,
controllability, noise and durability
L Toussaint, M Marques, N Morand, P Davies, C Groves
Honeywell, France
F Tomanec, M Zatko, D Vlachy, R Mrazek
Honeywell, Czech Republic

ABSTRACT

Over the past years, the evolution of emission standards as well as the advent of
Fuel economy targets has driven the need for higher controllability of boost
pressure. This is a strong trend, which will continue to be true for both gasoline and
diesel engine applications as we prepare for further evolutions of standards and
introduction of fuel economy regulations. While engine boundary conditions have
become harsher, fine, accurate and sustainable control of boost is challenging the
way mechatronics components are engineered. This requires us to apply a System
Engineering approach and improve our understanding of the behaviour of the entire
control chain so that we can ultimately influence the design and drive performance
at the System level.

This paper describes key elements of the kinematic tools developed by Honeywell
Turbocharger Technologies in order to optimize turbocharger control solutions. It
includes details on CAE applications such as fluid dynamics, flexible multi-body
dynamics and thermo-mechanical simulations and how they can be linked together
to analyse a complete system. It highlights the importance of understanding the
behaviour of the whole control chain from the actuator command to the
turbocharger and engine response.

The paper presents results from one main examples on a new waste-gate concept
for Fixed geometry turbochargers primarily for gasoline applications.

This examples show how kinematic tools can be applied to achieve controllability
and durability targets while accelerating development cycle time.

1 INTRODUCTION

Effective in 2014, the new emission regulation EU6 drives toward very low emission
levels. At the same time, fleet average C02 emission have to meet 130g/km.
Significant upgrade of chassis and powertrain are necessary to meet these
challenging targets. Turbocharging boosting system plays a major role in
powertrain optimization, for both emission control, allowing higher EGR rate, and
fuel economy.

1.1 Impact on Gasoline engine turbocharger control strategy


Explanations are given below as to why modern gasoline turbocharging largely rely
on fixed geometry / by-pass control of boost:

_______________________________________
© The author(s) and/or their employer(s), 2014
137
- With the stoichiometric burn, gasoline engines typically do not require any boost
pressure to operate in part load, in particular on the low load steady state points
of fuel economy driving cycles. As such, a by-pass system facilitates the
regulation of exhaust gas energy available at the turbine wheel and, thus allows
the turbocharger to be controlled and not to produce too much boost pressure.
- Adopting a by-pass strategy also allows sizing the turbocharger in order to
maximise boost at low engine speed, improving as well low end transient
performance. This is particularly key as a quick and high boost response is
needed in case of high driver torque demand in order to have the downsized
engine still behave like a bigger naturally aspirated engine.
- The sizing of the by-pass also allows the boost to be reduced in certain operating
conditions, even in the case of a small turbocharger. This directly impacts the
capability to control the driver’s torque demand without having to use the intake
throttle and avoids generating an undesirable engine back pressure, thus
increasing engine efficiency at key fuel economy in the point driving cycle.
- Note finally that, despite the small turbocharger matching generally used to get
good vehicle driveability, the relatively low amount of air required by gasoline
engines in typical rated power conditions still allows the engine to reach very high
specific power outputs in the range of 90 to 120 kW/l.

1.2 Impact on Diesel engine turbocharger control strategy


By-pass controlled turbochargers also exist on passenger vehicle diesel applications
but are becoming less and less used in developed countries as they are becoming
non competitive in terms of specific power (<50 kW/l).

Variable geometry turbochargers are already main stream for Diesel passenger
vehicles and are becoming increasingly popular in Euro 6 applications for the
following reasons:

- The high air to fuel ratio in Diesel engines requires a significantly higher boost
demand in most engine operating conditions. There is no need to operate at low
or zero boost as is the case with gasoline engines.
- Heavy EGR rates required to pass emission standards even increases the
compressor power needed in part load operating points.
- Allows reaching higher and higher specific torque and power while providing world
class fuel economy: Vehicles reaching up to 200N.m/l & 75kW/l for a CO2
emission in the range of 85 to 120g/km in ECE fuel economy cycle.

As a consequence, most of the typical control strategies for passenger vehicle


diesels rely on variable geometry.

2 NEW WASTE-GATE CONCEPTS FOR FIXED GEOMETRY TURBOCHARGERS


PRIMARILY FOR GASOLINE APPLICATIONS

In the following section, an example of development of a new by-pass system is


given as an illustration of kinematic simulation tools that have been developed in
order to address arising concerns.

First of all, a qualitative analysis of key by-pass functionalities is provided, linking


OEM requirements like performance, noise or durability with component level
targets (closed valve sealing, valve response time...).

In a later portion, orders of magnitudes are given for the key environmental
conditions (temperature, pressure, vibrations) in which the valve has to operate.

138
Example of the typical simulations methodologies used to study the interactions
between valve concept, design parameters and the different constraints and targets
are given.

These examples illustrate the interest of a newly developed valve solution taking
the form of a monoblock spheroid arm and valve (1 piece in opposition to widely
used 2 piece arm and valve construction).

2.1 Expected by-pass system functionalities for modern gasoline engines


The primary functions that a modern gasoline engine by-pass system has to full-fill
are the following:

- Maintain sealing in closed valve operation in order to maximise gas energy


directed to turbine wheel. This condition is typically required to either achieve low
end torque at low engine speed (< 1300 RPM) or provide quick boost response in
transient (load step-in at low engine speed). The closed valve sealing is also
important to entertain high instantaneous exhaust pressure peaks which are key
to recover max possible energy form exhaust gases into turbocharger power. The
key customer impacted criteria is therefore here low end transient and steady
state performance.

- Provide a smooth and progressive opening of the valve in order to avoid abrupt
air – and therefore torque – fluctuations that may be perceived by the driver as
the by-pass system needs to start opening to control boost pressure (typically
during the second phase of low end full load transients). The key customer
impacted criteria is therefore here vehicle driveability (comfort).

- Allow an easily controllable quasi steady state part opening of the valve to enable
the turbocharger waste-gate valve to partially control the boost pressure level in
part load. The goal here is to enable so called “de-throttling” strategies which in
turn reduce the pumping losses of the engine, thus improving slightly its
efficiency and fuel economy. The key customer characteristic impacted here is
therefore vehicle fuel economy.

- Ensure full control of over all environmental and operating ranges of the critical
physical parameters that must be maintained in order to protect
vehicle/engine/turbocharger integrity (Engine maximum mean effective pressure,
engine inlet pressure and temperature, engine exhaust pressure and
temperature, turbocharger speed...). This requirement basically results in the
sizing of the waste-gate port (in terms of mass flow Vs pressure delta
characteristic). The sizing allows the proper control of system critical parameters
but also depends on the matching of the turbocharger.

- Promote the minimum heat transfer to the surroundings and hence the maximum
exhaust gas heat / temperature transport directly to catalyst in the case of cold
start as measured during typical emission cycle. This requirement is not only
about the mass flow by-passed directly to the catalyst, but also about the velocity
and direction of the waste-gated flow and the likelihood to initiate fast catalyst
light off.

Following complementary functions also have to be taken into account while


designing the by-pass system:

- Enable quick valve response on transient event requiring large valve travel from
either large open to close or close to open. Events like stop and start followed by
a quick driver torque demand, or take-off after idling, or sudden drop of boost
demand at high engine speeds need a fast response of the waste-gate valve to
provide a prompt feed-back to the driver while keeping emissions under control.

139
- Prevent noise. As a consequence of the control strategies OEMs are putting in
place to meet fuel economy and fast response requirements, many low engine
speed / low engine load operating points are typically operated with a part open
waste-gate. With the high pressure fluctuations and engine vibration encountered
on modern and downsized engines, any kinematic joint operating with significant
clearance is likely to generate intermittent contact and shocks likely to emit air-
borne noises or to trigger structure borne noises. As engine operating conditions
typically happen while vehicle is moving at very low speed or even standing still,
any noise from the kinematic chain has to be prevented.

- Resist harsh engine boundary conditions. Emission and fuel economy standards
are also driving a clear trend on modern gasoline engine which is mostly known
as downsizing. The industry trend is clearly to reduce displacement and to have a
lower cylinder count while increasing engine output in terms of specific torque and
power. This results in a significant increase of engine exhaust pressure peaks
which have more than doubled in the last 5 to 6 years. This drives a trend of
increasing engine vibration as well as maximum exhaust gas temperatures.
Finally, using a downsized engine in a comparable vehicle results in simply having
a higher average operating load, thus increasing the time spent at higher
pressure, temperature and vibration.

- Promote usage of right sized actuators to limit their impact on current


consumption, self heating and its impact on sustainable maximum under hood
temperatures, packaging, mass and cost.

- Prevent turbocharger wheel outlet flow perturbation. As by-pass system typically


has to connect turbine inlet and outlet, design of the waste-gate port and nature
of the flow in partial or full opening of the valve usually results in wheel outlet
perturbations. This phenomenon, which is heavily influenced by turbine housing
and packaging design constraints, can adversely affect the boundary conditions at
the wheel outlet, therefore impact energy recovered by the wheel.

- Limit turbocharger outlet temperatures on highly loaded point once the catalyst
has reached steady state temperature. This requirement, which is a priori
contradictory with the necessity to promote quick catalyst light-off, aims to limit
the stress put into the catalyst under sustained durability. The by-pass can be
seen as a way to vary the exhaust gas flow to the catalyst so that it is favourable
to light off in typical cold start conditions and in different way in highly loaded
conditions to rather lower the direct exposure of the catalyst to the waste-gated
flow.

By pass function

Figure 1: Key by-pass functionalities

140
2.2 Typical boundary conditions for valve operation
In order to design turbocharger by-pass valves for gasoline engine application, The
key environmental boundaries resides in the description of pressures, temperatures
and vibrations. As explained earlier, fuel economy regulations are promoting the
use of downsized engines. These engines feature less cylinder count and less
displacement. As such many new 3 cylinders 1 to 1.5l engines are for example
being developed and released by OEMS.

They target the driveability of naturally aspirated superior displacement and


cylinder count engines like 4 cylinders 1.8l to 2.0l in the case of the above
example.

In order to keep the same low end full load torque, these engines are mechanically
resulting in higher maximum cylinder mean effective pressures (above 20b BMEP).
The lower cylinder count also results mechanically in higher exhaust peak pressures
to reach the same mean effective pressure.

Finally, those engines are typically used in vehicle masses which were previously
served with bigger engines. This results in driving a higher proportion of maximum
engine output to run typical steady state points. Note this effect is actually desired
since it allows using higher efficiency operating points. The undesired effect is that
the overall duty cycle of the engine and the turbocharger get increased to operate
more and more time in medium to high loaded points.

As a final effect on valve pressure boundaries, exhaust gas pressures have evolved
from typical 200 kPa average and 350 kPa peaks to 380 kPa average and 500 or
even 580 kPa peaks. The peaks are also changing in frequency because of the
cylinder count and the gearbox matching strategies.

Another factor affecting valve operation directly linked to engine superior loading
and lower cylinder count is the vibration. The acceleration levels are typically
increasing and covering lower firing frequencies. Note the trends are less generic
on vibrations than they are on exhaust pulsations as engine architecture (balancing
shafts or flywheel) is largely affecting them.

A final key environmental factor for valve design is the exhaust peak temperature
as well as the maximum positive and negative gradient the engine may generate.
There is no general trend again here other than the trade-off between system cost
and high power fuel efficiency OEMs are subject to. While exhaust temperatures in
the 980°C remain main stream, cases applying 1020°C or even higher
temperatures are getting released as well. Note this maximum exhaust
temperature selection is also linked to high power combustion strategies which in
turn can also affect exhaust gas oxidation potential. This oxidation potential has to
be taken into account with regard to the adverse effect it may have together with
mass flow and temperature on by-pass mechanism functionality over time.

2.3 Using computer fluid dynamic simulations to understand valve


loading and inherent flow evolution
Allowing an easily controllable by-pass valve in near closed position under the
boundary conditions described in the previous chapter remains a challenge with the
traditional flat poppet design. OEMs have been evolving from basic pressure
actuated valves to vacuum and even electrically actuated valves in order to have
more authority and control on such valves, but still facing difficulties and
instabilities.

In order to understand more closely the behaviour of typical waste-gate valves in


these conditions, computer fluid dynamic simulations have been developed. They

141
allow formulating the nature of the waste-gated flow as a function of time, valve
opening and operating conditions.

The simulation also allows better understanding what governs the valve loading
behaviour, which is another key factor to look at to assess valve controllability.

Figure 2 below illustrates the static pressure contour of a flat poppet as a function
of opening. As can be seen a flat valve results in two consequences that are
detrimental to controllability. First of all, the valve flow increases very aggressively
with the valve angle. Secondly, the valve loading also sees an abrupt increase when
the valve starts to open, and then reaches a maximum before decreasing again.
These flow and load evolutions near the fully closed positions definitely promote
valve instabilities and generate extreme difficulties to control the turbocharger.

Figure 2: Flat poppet flow & loading as a function of valve angle


(top 2 graphs) and valve static pressure contours (CFD contour plots)

Based on this understanding, a new concept of valve has been engineered in order
to provide a smoother evolution of flow and load as a function of valve opening,
especially at low openings. The key feature of this new valve resides in a different
sealing interface between the valve and the seat.

While a traditional poppet relies on a flat surface contact and a significant clearance
between the poppet and the arm to accommodate for tolerances and thermal
relative displacements, the new proposed valve features a conical seat and a
toroidal valve shape interface.

This interface allows reducing the area evolution which is subject to turbine inlet
pressure when the valve opens, as suggested in the illustrations of Figure 3.

142
Another interest of this interface is that it naturally results in a smoother geometric
area increase while the valve is opening and provides a “self centering” ability
which no longer requires clearance between the poppet and the arm. Finally the
valve features a spheroid protrusion (Patent pending) inside the waste-gate
channel that contributes to splitting the flow around 360 degrees.

Figure 3: Valve sealing interface and impact on load evolution

Figure 4 presents the results in terms of flow and load that are achieved with this
new concept. As can be seen, not only is the valve angle required to achieve the
same flow higher, but the load required to keep the valve in this position is also
lower. The overall effect is to decrease the load by approximately 30% compared to
the original valve. This is a clear advantage for valve and kinematic chain
durability. It also decreases drastically the impact pressure pulsations may have on
the valve position or load fluctuation. Finally, a nice monotonic load profile is
achieved, which is definitely helping the ability to maintain the valve in near closed
position whether it is with vacuum or electric actuator.

Note that alternative protrusion shapes have also been investigated (conical as well
as a complex non axi-symetrical 3D design). These shapes still provide the nice
features of monotonic load evolution as well as a better flow progressivity than the
baseline flat poppet. However, the best achieved characteristic is definitely
achieved with the spheroid protrusion version.

The resulting proposed conical seat to toroidal poppet monoblock arm and valves
have also been tested on a cold flow bench, confirming the results anticipated by
the CFD. As a consequence the spheroid protrusion has been selected to proceed to
further on engine and vehicle controllability assessment.

143
Figure 4: Flat poppet Vs Monoblock valve flow and load evolution (CFD)

Figure 5 presents the experimental results that have been collected on the cold flow
bench. Note that this test has been performed using the target pneumatic vacuum
actuator in order to directly understand the relationship that exists between the
commanded vacuum and the resulting waste-gate effective area. This metric
basically combines the effects of flow and load progressivity, since vacuum actually
sets a force on the actuator and not a position.

The combined effect of the monoblock arm and valve result in a smoother effective
area progressivity.

On the other side of the graph towards the full open position, the monoblock arm
and valve finally catches up the full open effective area on the last 20 mb. This is
due to the fact that the valve spheroid protrusion finally gets totally out of the
channel in the last degrees of opening of the valve.

Figure 5: Experimental results on waste-gate effective area (cold bench)

144
2.4 Using combined full stage finite element analysis and multi-body
simulation to ensure valve functionality in critical thermal transient
conditions
As explained earlier, in order to ensure full functionality, the proposed monoblock
toroidal portion of the arm and valve has to automatically centre in the turbine
housing conical seat.

Complex 3 dimensional stack-up calculations have been performed in the cold state
to ensure clearances in the kinematic chain allow the valve to self centre without
any hyper static conflict.

As the valve actually operates rather in a hot temperature environment, effects of


thermal distortions and friction coefficients while the valve approaches the seat also
had to be taken into account. To that end, finite element analysis of the turbine
stage (including arm & valve) has been performed.

The goal of the finite element study is to understand what are the thermal transient
events and critical time through those events that impact the deformation of critical
geometries entering into valve self centering stack-up. Those key geometries
include seat and valve deformation in the region of sealed contact, but also the
relative displacement between the arm and valve guiding and the seat.

The study then consists in retrieving a worst case deformation of all the parts and
their relative position at the critical time step of the most severe transient thermal
event.

This critical hot deformed shape is then imported into a multi-body simulation tool
in order to assess if the valve does naturally self centre in the seat while it is being
closed through actuator force. Note this multi-body simulation also uses as an input
the valve loading computed through CFD and explained previously.

The figure 6 explains the simulation flow logic that is used to assess the self
centering of the valve in hot conditions.

Figure 6: Combining CFD, FEA and multi-body simulation

145
As an outcome of these simulations, the monoblock arm and valve design has been
refined in order to achieve an optimum design in terms of performance, durability,
noise and controllability.

With breakthrough results, this design is about to enter production on a 3 cylinder


gas application.

It has demonstrated a remarkable robustness, allowing reducing significantly


kinematic wear, especially at the valve arm to bushing contact. This results in an
actuator stroke drift (difference between the actuator position in closed valve
position at new and after test) which is limited to less than 0,2 mm over more than
750h of engine test on dynamometer bench where a traditional arm and valve used
to result in up to 2 mm stroke drift due to kinematic wear.

Having less actuator stroke drift over engine life not only allows to keep the same
actuator position to valve position characteristic, but also ensures a constant
available force to keep the valve closed.

Indeed, regular valve closed turbine stage permeability measurements have been
performed during vehicle endurance tests. They have demonstrated very low and
stable valve leakage behaviour over more than 1000h on multiple units, resulting in
no noticeable low end torque drift on the vehicles.

This kinematic wear robustness and stroke drift improvement as well as the stable
and low valve closed leakage allows to keep the control system closer to the new
state over the whole life of the vehicle (same command needed at the actuator for
the same engine operating point). This results in having a stable low end torque
performance as well as a stable dynamic response and driving comfort.

Since it removes clearance between the valve and the arm, the monoblock
construction introduction also solved noise issues linked with valve rattling against
arm in mid open positions.

Finally the controllability enhancement provided by the smoother flow progressivity


and the monotonic evolution of valve loading as a function of valve opening, did
enable to achieve a stable control strategy with part open waste gate in part load,
therefore enabling the customer to meet its driveability and fuel economy targets.

3 SUMMARY AND OUTLOOK

In this paper, the methodology used to develop the new monoblock spheroid valve
design has been presented, from performance target setting to kinematics design.

Overall performance benefit versus previous generation flat poppet has been
exposed. The new valve brings significant controllability, noise, durability and
performance benefits.

This translates in vehicle in an improved driving experience, less sensitivity to noise


while enabling world class fuel consumption and emissions meetings EU6 targets.

Looking at the next evolution of emission regulation, challenges for powertrain


designers seem even more emphasized. Indeed, for 2017, EU6.2 regulation is
expected to introduce a new emission cycle (WLTC) with much higher dynamics and
load than NEDC cycle, in addition to the need to respect emission during real drive
test (RDE, Real Drive Emissions). This higher dynamics, and the need to control
emissions at higher load with high accuracy will require further enhancement on
turbocharging control chains, both on performance and controllability.

146
Experimental and analytical investigation
of implementing a ball bearing
turbocharger on a production diesel
engine
Q Zhang, T Duda, R Burke, S Akehurst, C Brace
University of Bath, UK
G Capon, P Dowell
Ford Motor Company, UK
P Davies
Honeywell Turbo Technologies, France

1. ABSTRACT

Ball bearing turbocharger technology has started to be adopted for mass-production


engines due to the potential benefit in transient performance and fuel consumption.
Compared to the conventional journal bearing, the low friction of the ball bearing
allows the turbocharger to accelerate faster so that the engine can be supplied with
boost pressure more quickly following a transient torque request and under steady
state offers reduced engine back pressure, which can reduce engine fuel
consumption. In this study, the benefits of using a ball bearing turbocharger
compared to a conventional journal bearing turbocharger were identified first in
simulation and then validated in a back to back comparison of two otherwise
identical turbochargers through extensive experimental analysis.

The cold start engine performance was significantly improved as the ball bearing
turbocharger was able to boost the engine to the full load level within a few engine
cycles. The hot engine transient response was also improved, but the full potential
of the ball bearing turbocharger in terms of the transient performance can be
further exploited by recalibrating the engine. The fuel consumption of the engine
was greatly reduced by the ball bearing turbocharger. However, closer scrutiny
reveals that the insufficient EGR rate of the ball bearing turbocharger equipped
engine was the main cause of the reduced BSFC, and that further engine calibration
is a must before any fair evaluation of the BSFC benefit can be done.

NOMENCLATURE

BB ball bearing
BMEP brake mean effective pressure
BSFC brake specific fuel consumption
, specific heat capacity at compressor condition
, specific heat capacity at turbine condition
EGR exhaust gas recirculation
axial thrust force
JB journal bearing
journal bearing correction factor
thrust bearing correction factor

_______________________________________
© The author(s) and/or their employer(s), 2014
149
compressor mass flow rate
turbocharger oil mass flow
turbine mass flow rate
NEDC new European drive cycle
turbocharger speed
, , , turbocharger inlet/outlet pressures
turbine inlet pressure
turbine outlet pressure
compressor inlet temperature
compressor outlet temperature
turbine inlet temperature
VGT variable geometry turbine
WLTC worldwide harmonized light vehicles test cycle
compressor power
power of turbocharger mechanical loss
turbine power
heat capacity ratio
ΔH compressor enthalpy change
ΔH , isentropic turbine enthalpy change
, turbine efficiency on gas stand map
turbocharger mechanical efficiency
, turbine efficiency w/o mechanical loss
oil viscosity

2. INTRODUCTION

As a cost effective method to increase the power density of internal combustion


engines, the turbocharging technology is becoming the standard component on
most Diesel engines and many gasoline engines. The high driver expectation and
strict emission legislation have never ceased to push this technology even further.
Many new technologies are becoming available and affordable because of the
intensive research input and the surging fuel price. One of such technologies is the
implementation of the ball bearing in the turbocharging system.

The bearing system of a turbocharger is a crucial component as it represents an


energy loss in the transmission of power from the turbine to the compressor. At the
same time, it must be tolerant of high thrust loading, oil contaminants, oil supply
delay and hot shutdown. The conventional journal bearing solution has been able to
fulfil these requirements and is the prevalent solution nowadays. On the other
hand, the ball bearing system has significantly lower friction and offers better fuel
economy and faster transient response.
However, the high precision
requirement and the high cost has
limited its application to niche products,
large commercial vehicles and racing
cars. After years of consistent
development, the technology has
become mature enough to be supplied
to the mainstream market; and the
high fuel price can well justify the cost
for the technologies providing better
fuel efficiency. A typical turbocharger Figure 1 Turbocharger Ball Bearing
ball bearing cartridge structure is Cartridge w/o outer ring
shown in Figure 1. (Source: Schaeffler AG (1))

150
Compared to the conventional journal bearing cartridge, the turbocharger shaft is
supported by the two ball bearing assemblies in place of floating metal bushes. This
ensures much lower friction performance in a wide speed and temperature range.
The ball bearing can also take the thrust load, eliminating the need for a separate
thrust bearing which can further reduce the friction. The transient response,
especially of the cold engine, is expected to be greatly improved. Fuel consumption
should also see a moderate reduction due to the reduced engine exhaust manifold
pressure (2).

The aim of this study was to provide a back to back comparison of journal and ball
bearing technology both experimentally and in simulation. Two aerodynamically
identical turbochargers were implemented on a 2.2L Diesel engine in a Ricardo
Wave simulation environment and on a transient dynamometer. A series of
experiments were then conducted covering part load, full load and transient
conditions.

3. BALL BEARING MODEL

The characteristic efficiency map of the turbocharger turbine is commonly


determined on a gas stand facility based on the measured enthalpy rise in the
compressor (equation 1). It is determined in this way to avoid the effects of heat
transfer in the turbine and is based on the assumption of adiabatic operation of the
compressor which is justified at high operating speeds. However, as a consequence
the turbine map includes the mechanical efficiency of the bearing system.

ΔH , −
, = =
ΔH , 1
, 1−






(See nomenclature for more)

For the purpose of simulating a particular turbocharger on a particular engine, this


is sufficient as the requirements are to estimate the overall transfer of power from
the turbine to the compressor. However, in the context of this work which is
focused on the benefits from a particular bearing system, the turbine map needs to
be corrected to represent only the aerodynamic performance; this is achieved
through equation 2

,
, = 2

The mechanical efficiency of the bearing system is determined by the ratio of


compressor work to turbine work

= = 3
+

Where is defined by the numerator of equation 1 and ℎ is determined


through the model developed by Serrano et al.(3) which depends on the internal
geometries of the journal and thrust bearings and the oil viscosity. This model is

151
summarised in equation 4, but for full details the reader is directed to the original
publication. The model has been applied to the manufacturer supplied map using
the conditions of oil temperature and viscosity rating under which the map was
originally measured on gas stand. In this way a turbine efficiency map separated
from bearing friction has been calculated.

, , ,
= + 4

:
:
: ℎ
: ℎ

Previous studies of ball bearing friction (4, 5) suggest friction reductions of around
50% compared to journal bearings. Therefore, due to the lack of measured bearing
friction data for the specific turbochargers used in this study, the friction model
introduced here will be applied to the 1-D gas dynamic model. It is expected that
the error introduced by the friction model will only account for a small fraction of
the errors by a 1-D model and that the simulation should, to some extent, allow us
to look at different control logics that are either difficult to check experimentally, or
are to be examined in the next phase of the study.

4. EXPERIMENTS

A 2.2L Diesel engine was chosen as the test engine and the default turbocharger
and an updated device with ball bearing cartridge were used in turn on the engine.
The specifications of the selected engine and turbochargers are summarized in
Table 1. In both cases, the only difference between builds was the turbocharger
bearing and the engine calibration remained unchanged. This is important in that
the engine controller has two key set points of inlet manifold pressure and intake
air mass flow rate. The controller will adjust the VGT rack position and the EGR vale
in order to meet these two targets.

Table 1 Test Engine and Turbocharger Specifications

Engine Configuration L4, Turbocharged, Intercooled


Displacement [L] 2.2
Bore [mm] 86
Stroke [mm] 94.6
Connecting Rod Length [mm] 155
Compression Ratio 15.5:1
Max Speed [rpm] 4900
Max Torque [Nm] 385
Max Cylinder Pressure [bar] 160
Max Turbocharger Speed [krpm] 213
Max Pre-Turbine Temperature [°C] 830
Max. Compressor Outlet Temperature [°C] 180

152
The whole air-engine-gas path is monitored by paired temperature and pressure
sensors at 40~80Hz. High frequency pressure measurements were taken at crucial
locations such as the exhaust ports, post/pre turbocharger so that hot air pressure
rise after transients and exhaust gas pulsation details can be recorded and analysed.
The table 2 summarises a list of types of sensors used in this study.

Table 2 List of Types of Sensors


Pressure PTX: Druck and RS
Temperature K Type Thermocouple
Mass Flow ABB Sensyflow
Fuel Flow (steady state) Gravimetric Fuel Balance
Turbocharger speed MicroEpsillon Eddy Current
Hot Gas Pressure Kistler 4049
Cold Air Pressure Kistler 4007

The experiments were designed to cover the most pertinent engine operative
conditions, including:

 Limiting torque curve


- The engine was tested for steady state full load torque at hot engine
condition.
 Cold start torque transients
- The engine was started from 15 °C and was controlled to enter into the
full load transient schedule within 20s (time for the fuel beaker to settle).
 Hot engine torque transients
- Same test as the cold start torque transients at hot engine condition.
 Part-load steady state points generalized from combined WLTC/NEDC drive
cycles (Figure 2).
- The engine was tested for steady state part-load at hot engine condition.

Figure 2 Combined WLTC/NEDC Minimap Points

153
The part-load steady state points were selected with a weighting process from the
WLTC/NEDC drive cycle simulations using a representative vehicle.

Due to the limit in the paper length, only representative test results of each of the
four tests are presented in the paper.

5. RESULTS AND DISCUSSION

5.1. Limiting torque curve


At full load, the ball bearing turbocharger does not give an apparent improvement
to the engine torque performance. This is largely because the engine full load
condition was often limited by factors other than the turbocharger performance
(cylinder pressure, exhaust manifold temperature, etc). Therefore, with the same
calibrated boost target, the air mass flow rates are similar and the full load torque
was improved up to 6 Nm at 2000 rpm, mostly from reduced pumping work
because the engine back pressure is reduced by the ball bearing turbocharger. The
BSFC is consequently also only marginally improved (Figure 3).

Figure 3 Limiting Torque Comparison and BSFC Reduction


(Positive percentage -> Improvement)

5.2. Transient response


5.2.1. Cold torque transient
The cold start torque transient tests showed a clear advantage of the ball bearing
turbocharger (Figure 4). The curves illustrate that the ball bearing turbocharger
started from Turbocharger speed 50% higher than the journal bearing turbocharger
and maintains a speed at least 10% faster throughout the transient test. The result
in torque performance was significant: the ball bearing turbocharger achieved 90%
of the full load torque from the first tip-in; whereas the journal bearing achieved
only 75% and only reached 90% in the 6th transient. When looking at the first
transient response only, the ball bearing achieved 90% torque 1.3s faster than the
journal bearing turbocharger, which stabilized at torque of 40 Nm lower. It should
also be noted that the full load torque is the same for both turbochargers at this
speed and the ball bearing turbocharged engine managed to achieve this torque
from the 5th transient. The journal bearing did not achieve its full load torque within
the transient test period as it will only happen when the engine oil is fully warmed
up.

154
Figure 4 Cold Start Torque Transient at 1250 rpm and the Turbospeed
Difference

5.2.2. Hot torque transient


For a fully warmed up engine, the ball bearing turbocharger still provides better
torque transient at low engine speed. However, the difference in shaft friction
becomes smaller and the friction loss counts as a smaller proportion of the total
work done by the turbine, especially at higher speed.

The transient response comparison at 1000 rpm is shown in Figure 5. It is clear


that the ball bearing turbocharged engine responds faster than the journal bearing
turbocharged engine. The time to 90% JB torque is reduced by 1.2s (41%
reduction); and the ball bearing stabilized transiently at torque 14 Nm higher.

Figure 5 Hot Engine Torque Transient at 1000 rpm and the Exhaust
Manifold Pressure

In both cases, the VGT vanes are at the fully closed position before the transient.
This is a calibration setting for the engine to generate high EGR gas at low engine
load and/or to prepare a fast spinning turbocharger for the transients. At such
conditions, the ball bearing turbocharger generates a lower engine back pressure.
Therefore, the EGR valve in the ball bearing turbocharged engine was more open to
ensure that similar EGR rate can be achieved with the lower back pressure (EGR
rate 44% compared to 42% for JB). Although with lower back pressure, the ball

155
bearing turbocharger has a higher speed prior to the transient (35.8 krpm
compared to 26 krpm): this translates to faster torque rise.

The transient response comparison at 1500 rpm is shown in Figure 6. Unlike at


1000rpm, the torque rises are very similar for ball bearing and journal bearing
(even slightly faster with the journal bearing); and both reached the same
stabilised torque (BMEP 21.5 bar).

Figure 6 Hot Engine Torque Transient at 1500 rpm and Relevant


Parameters

156
Before the transient, the ball bearing turbocharger was able to generate sufficient
pressure in the inlet manifold to meet the calibration target (VGT 58% shut
compared to 78% for JB). The EGR valve is more opened in the BB case such as to
meet the EGR requirements (61% open compared to 42% for JB).

Compared to the 1000 rpm transient test, the exhaust energy is more abundant so
that the more closed VGT vane position leads to a higher turbocharger speed and
back pressure for the JB turbocharger. When the transient happens, the journal
bearing turbocharger has faster exhaust pressure build up to accelerate the
turbocharger and therefore to provide boost. The overshoot in turbocharger
speed/boost also leads to an obvious overshoot in torque in the journal bearing
transient test. In terms of the torque rise, the benefit of higher back pressure with
the journal bearing turbocharger balances the benefit of lower shaft friction and
lower pumping work of the ball bearing turbocharger. However, in terms of exhaust
manifold component durability, fuel consumption and controllability, ball bearing
turbocharger is the better option.

5.3. Part-load points fuel consumption


The part-load quasi steady drive cycle tests can be seen as a crucial indicator of the
engine fuel consumption. However, when the test results reveal that the ball
bearing turbocharger gives a large fuel consumption benefit, it is clear that without
re-calibrating the engine for the ball bearing turbocharger, it is very difficult to
generate a convincing engine fuel consumption performance in the drive cycle part-
load test.

Figure 7 BSFC Comparison of Minimap Points at 1500 rpm


(Positive percentage -> Improvement)

Figure 7 shows the raw fuel consumption benefits for the range of engine loads at
1500rpm resulting from the back to back comparison of the two turbocharger
bearing technologies. These differences are between 3-11% improvement through
the turbocharger and it is clear that this benefit is not solely a result of reduced
turbocharger friction, but due to interactions with other engine systems. Figure 8
shows the impact of the turbocharger bearing on back pressure and EGR flow. The
back pressure is reduced across the load range by 15 - 29 kPa. This reduction in

157
backpressure has a significant effect on EGR rate, notably between 50-250Nm
where the largest fuel consumption gains are made. This reduction in EGR rate of
up to 15percentage points would have a large impact on fuel economy, but also
NOx and soot emissions.

Figure 8 Back Pressure and EGR Rate Difference between JB and BB


Equipped Engine at 1500 rpm

According to the research on the diesel engine back pressure (6) and EGR rate (7),
for every 1 Bar of increased back pressure, the BSFC penalty is around 50 g/kWh;
while for 1% of EGR rate, there will be a BSFC increase of around 0.3%. A crude
calculation of the acquired BSFC benefit through back pressure and insufficient EGR
flow is illustrated in Figure 9.

Figure 9 BSFC Reduction Analysis Compared to Measurement

158
- The calculated BSFC reduction correlates reasonably well with the measured
BSFC reduction (Total and Measured BSFC Difference). This analysis allows a
crude isolation of the fuel economy benefits due to higher efficiency of the
turbocharger and reduction of the EGR flow, with a prediction of 5-15g/kWh
benefit in fuel consumption from the reduction in back pressure.

Figure 10 is a compiled graph of the rest of the part load points at 1000 rpm, 2000
rpm, 2500 rpm and 3000 rpm. This is consistent with the results shown at 1500rpm.

It should be noted that such BSFC reduction is of limited practical use as the engine
will still be required to meet NOx limits which on the current engine will require
similar EGR rates as the JB configuration. To meet such requirements would involve
the complete re-optimisation of the engine controller and notably changing the
targets for boost pressure and EGR flow based on the emissions/fuel economy trade
off. However such a study is beyond the scope of this paper.

Figure 10 BSFC Comparison of Minimap Points at 1000 rpm, 2000 rpm,


2500 rpm and 3000 rpm

Although a fair comparison was intended between the two turbochargers, it was
later realized that a perfect back to back comparison is near to impossible in the
case of the engine-turbocharger system: any parameter changed would have a
chain effect on the parameters in the engine-turbocharger loop. The dilemma
suggested yet an optimal solution: by choosing a slightly smaller turbo-machinery,
the higher back pressure is expected to be able to drive the EGR gas back to the
inlet manifold which is not so highly boosted. The transient response can be further
improved in the meantime.

Ideally the engine should be at least recalibrated for the newly fit ball bearing
turbocharger. However, this process was not permitted in the time span of this
project. Therefore, it was decided that the engine calibration remained unchanged
for both turbochargers. This gives a comparatively conservative demonstration of

159
the benefit of the ball bearing turbocharger as the calibration was optimized only
for the original journal bearing turbocharger and the simulation study was designed
to cover some of the weakness of the experimental study.

6. SIMULATION STUDY

Since the time and calibration effort required to perform a perfect back to back
comparison of the ball bearing and journal bearing device are not allowed in this
first phase of the project, a simulation study was conducted in an engine wave
action model of the 2.2L Diesel engine using Ricardo WAVE software. The journal
bearing turbocharger was simulated by characteristic maps of the compressor and
turbine, supplied by the turbocharger manufacturer; and turbine map was adjusted
using the Valencia model to represent the ball bearing turbocharger. The model was
run in conjunction with Mathworks MATLAB Simulink to allow a high level
implementation of the engine controller software and was calibrated against both
high and part load engine operating conditions running up to 3000 rpm.

6.1. Hot torque transient simulation


The transient simulation involves improving the 1500 rpm ball bearing turbocharger
torque response. It was observed that the ball bearing turbocharger produced
slower torque rise at higher speed compared to the journal bearing due to the fact
that the more efficient ball bearing turbocharger rests in lower turbocharger speed
compared to the journal bearing turbocharger. Therefore, the boost target of the
ball bearing turbocharged engine is adjusted to a higher value, so that the VGT will
be at a more closed position to prepare for the torque tip-in. The simulation result
is illustrated in the Figure 11.

Figure 11 Simulated Hot Engine Torque Transient at 1500 rpm and the
Exhaust Manifold Pressure

As shown in the figure, with a higher boost target, the ball bearing turbocharger
generated a similar level of engine back pressure which implied a similar
turbocharger speed. As a result, the torque generated by the ball bearing
turbocharged engine has the fastest climb compared to the bench mark engine and
the ball bearing turbocharger running the original engine calibration. This suggests
that an engine re-calibration would benefit the ball bearing turbocharged engine in
the transient performance in a wider speed range.

160
6.2. Part-load points simulations
The part-load points simulation focuses on one of the problematic speed that
produced overestimations in fuel consumption benefits. It was pointed out in the
analysis in section 5.3 that when running the ball bearing combined with original
engine calibration, the engine will not achieve the EGR rate required for emission
control. Therefore, the simulation was designed to achieve the EGR rate targets
instead of the mass flow targets and the results are supposed to reveal the true
fuel consumption benefits can be achieved by implementing a ball bearing
turbocharger on the assumption that similar EGR rate level would produce similar
level of NOx emission. To ensure an accurate as possible prediction of this steady
state fuel consumption simulation study, the original journal bearing turbocharged
engine model was especially calibrated manually for the FMEP of each selected
brake torque.

The results are shown as in the Figure 12 below.

Figure 12 Simulated BSFC Comparison of part-load points at 1500 rpm


(Positive percentage -> Improvement)

The simulation results demonstrated that with a similar level of EGR rate, the
average fuel consumption benefit of the ball bearing turbocharger settled to a
reasonable value of 2.5% BSFC reduction (without the 250Nm negative result taken
into account). The results are also in line with the theoretical fuel consumption
benefit from lowered engine back pressure.

7. CONCLUSIONS

In this paper, a novel turbocharger equipped with ball bearing rotor was installed
on a production engine to evaluate the benefit in terms of fuel economy and engine
transient response. An aerodynamically identical journal bearing turbocharger was
also tested as the benchmark.

161
The test and simulation results showed that

1. There is a significant benefit of cold start transient response can be gained


by implementing the ball bearing turbocharger.
2. Large fuel consumption benefits can be seen when running part-load
steady state tests because of the interactions with EGR system.
3. Simulation results showed that with small modifications to the engine
control strategy, fuel consumption benefit of 2.5% could be gained.
4. An engine control system re-optimisation is needed to make a true back-
to-back comparison.

REFERENCE LIST

(1) P. Davies, D. Marsal, R. Michel, P. Barthelet, Ball Bearing goes Mainstream,


IQPC, Wiesbaden, Germany, 2013
(2) Honeywell Turbo Technologies, Ball Bearing, [online] Available at:
http://turbo.honeywell.com/our-technologies/ball-bearing/ (Last accessed:
Oct 21st 2013)
(3) J. R. Serrano, P. Olmeda, A. Tiseira, L. M. Garcia-Cuevas, and A. Lefebvre,
“Theoretical and experimental study of mechanical losses in automotive
turbochargers”, Energy, vol. 55, pp. 888-898, 2013
(4) B. Griffith, R. C., Slaughter, S. E. and Mavrosakis, P. E., “Applying Ball
Bearings to the Series Turbochargers for the Caterpillar® Heavy-Duty On-
Highway Truck Engines”, SAE Paper Number 2007-01-4235, 2007
(5) M.D. Brouwer, F. Sadeghi, C. Lancaster, J. Archer and J. Donaldson, Whirl
and Friction Characteristics of High Speed Floating Ring and Ball Bearing
Turbochargers, Transactions of the ASME: Journal of Tribology, vol
135/041102-1, 2013
(6) P. Hield, The Effect of Back Pressure on the Operation of a Diesel Engine,
Maritime Platforms Division, Defence Science and Technology Organisation,
2011, Australia
(7) M. van Aken, et al., Appliance of high EGR rates with a short and long route
EGR system on a Heavy Duty diesel engine, SAE Paper Number 2007-01-0906,
2007

162
Optimisation of transient response of a
gasoline engine with variable geometry
turbine turbocharger
H Tang, S Akehurst, C J Brace
University of Bath, UK
S Garrett
Cummins Turbo Technologies Limited, UK
L Smith
Jaguar Land Rover Limited, UK

ABSTRACT

Maintaining transient torque response is challenging on turbocharged engines


because of the period of time required to accelerate the turbocharger. Variable
Geometry Turbine (VGT) turbochargers offer a route to improve the transient
response. In order to explore the transient operation without any limitation imposed
by the production control strategy, an on line search was conducted using a series
of open loop actuator trajectories applied to a VGT turbocharger installed on a
gasoline engine. The trade-off between the responses in different stages in the
transient event has been illustrated. The time required to reach 50% of maximum
torque rise (T50) was improved by up to 0.54s (35.5%) whilst the turbocharger
acceleration was maintained. Fully closing the VGT resulted in high exhaust back
pressure and low volumetric efficiency. This suggests that a simple boost pressure
feedback control will likely not deliver optimised performance due to the excessive
exhaust back pressure, reducing the available brake torque during the early part of
the transient. Therefore, a model based control strategy may be required.

1 INTRODUCTION

Downsizing the internal combustion engine can reduce the fuel consumption by
moving the fuel efficient zones closer to road driving conditions. This can be
achieved by turbocharging [1]. A mass air flow ratio of over 80:1 from rated power
to idle is typical for gasoline engines, in contrast to the ratio of approximately 6:1
for passenger car Diesel engines [2]. It highlights the demand for varying the
characteristics of the boosting systems on gasoline engines [3]. In addition, the
transient response of conventional turbocharged gasoline engines is usually slow
compared to naturally aspired engines. This is due to the period of time required to
accelerate the turbocharger, achieve the target boost pressure, and reach the
required engine torque [4, 5].

One of the available boosting systems that have the potential to achieve rapid
transient response and high fuel efficiency over a wide flow range is the variable
geometry turbine (VGT) turbocharger [3]. Compared to fixed geometry turbine, the
VGT can change the gas velocity and flow angle to vary the turbine characteristics
[6]. In addition, the VGT turbocharger can allow all the exhaust gas to pass through

__________________
© Huayin Tang, 2014
163
the turbine. Hence, the specific enthalpy gradient required to drive the compressor
can be reduced [7]. As a result, the VGT can achieve higher turbine efficiency over
a wider operating range [8]. Moreover, both steady state and transient control
strategy can be improved due to the higher control flexibility [9].

However, due to the complexity of the mechanism, the VGT concept is applied
predominantly on the Diesel engine. This is because the exhaust gas temperature
of gasoline engines can reach over 1000°C, over 200°C higher than that of Diesel
engines [10, 11]. As a result, the application of VGT turbocharger on gasoline
engines is more challenging.

Despite the difficulties, several types of VGT are available. The variable geometry
nozzle turbine (VNT) [12], variable turbine housing throat area (VAT) [13], and
variable flow turbine (VFT) [14] can vary the effective flow area radially, while the
sliding wall with variable axial width [15], and twin scroll switching type [16] vary
the effective flow area axially. The VNT type offers higher peak efficiency and larger
flow range.

To illustrate the potential of using VGT on gasoline engine, the transient response of
a gasoline engine fitted with a VNT turbocharger has been optimised in this study.

2 METHODOLOGY

2.1 Test set-up


The experiments were performed on a 2.0L, four-cylinder, direct injection,
turbocharged gasoline engine with variable intake and exhaust-valve timing system.
The maximum power output of the tested engine is 150 kW. The test was carried
out on a dynamic engine test bed in the University of Bath. The original fixed
geometry turbine turbocharger on the engine was replaced with a VNT turbocharger.
In order to study the effect on transient engine operation unconstrained by possible
deficiencies in the feedback controller, a series of open loop VGT actuator
trajectories generated in dSpace environment were provided to an electric actuator
using CAN message to control the VGT position. The timing of the transient
trajectory in dSpace was triggered by the pedal position voltage signal from test
bed control system. The blow-off valve on the compressor housing was deactivated.

In order to eliminate the interaction between the turbocharger and the valve timing,
the intake and exhaust valve timings were kept at the maximum overlap positions
in the transient test. The Lambda target was maintained at 1, and the original
production level spark timing control strategy was used. Before the start of each
transient test, the engine was settled at 2 bar BMEP for five minutes. The throttle
was fully opened at the start of the transient test, and the VGT transient trajectory
was also started at the same time. The engine speed was maintained at 2000 rpm,
as the available boosting capability is relatively high and the transient response of
the engine is crucial at this engine speed. The coolant temperature at engine outlet
and the air temperature at intercooler outlet were maintained at approximately
90˚C and 30˚C, respectively.

Slow measurements for pressures, temperatures, flow, and torque were logged at
80Hz. Fast measurements of cylinder pressures, port pressures and temperatures,
injection pressures, turbocharger speed as well as pressures and temperatures at
turbine and compressor inlet and outlet, etc. were recorded at 0.5 degree crank
angle resolution. The demanded and actual VGT actuator position signals from the
CAN message were recorded in the fast measurement system, and the response
time of the VGT actuator was approximately 0.05 – 0.1 second.

164
2.2 VGT actuator trajectories
To determine the VGT position before tip-in, the steady state fuel consumption and
turbocharger speed at three VGT positions, 16% closed (fully open), 60% closed
and 90% closed (fully closed) were tested at 2 bar BMEP 2000 rpm. The averaged
fuel consumptions at the three operating points were the same considering that the
accuracy of the measurements is ±0.05%. However, the averaged turbocharger
speeds at the three VGT positions were 7.7 krpm, 12.9 krpm and 19.3 krpm,
respectively. Therefore, the VGT position was chosen to be fully closed at low load
operating point before transient trajectory was triggered.

The purpose of this experiment was to analyse the transient behaviour of the
engine in the first 1.5 seconds, which is crucial in the engine transient response.
Due to the need of having more than two stages in the transient to optimise the
dynamic response of the system, the limit of the number of tests, and the VGT
actuator response time, it was chosen that the 1.5 seconds was divided into three
0.5 second stages, each of which had a number of optional VGT positions. In order
to define the boundaries for the VGT positions in the transient operation, a range of
VGT actuator trajectories with single step change, from fully opened to fully closed
(16% fully open, 40% closed, 60% closed, 70% closed, 75% closed, 80% closed,
85% closed and 90% fully closed), have been tested, shown in figure 1.

Figure 1 Torque responses of the single step change VGT actuator


trajectories.

It was found that with the VGT position between 60% and fully closed after tip-in,
1.1 bar boost pressure downstream the compressor can be achieved at 1.5 second
into transient. In addition, although 50% of the maximum torque rise (144 Nm
higher than the low load operating condition, 32 Nm) can not be achieved within
1.5 seconds with the VGT position 80%, 85% and 90% closed, these three VGT
positions were still selected as options because of the potential of accelerating the
turbocharger faster. Therefore, six different VGT positions (from 60% to 90% closed)
were selected as options at each one of the three transient stages after tip-in. Thus,
216 transient tests with different VGT actuator trajectories were carried out.

165
3 RESULTS AND DISCUSSIONS

3.1 Response of engine torque


The engine torque responses of six tests with different VGT positions at the second
stage (0.5 – 1.0 second) are compared in figure 2. The VGT position described in
the figure is defined as follows: the VGT setting before tip-in, VGT position at first
stage (0.0 – 0.5 s), VGT setting at second stage (0.5 – 1.0 s), and VGT setting at
third stage (1.0 – 1.5 s). This is the same for other following figures. The
comparison of torque response at second stage is also shown in figure 3. The VGT
positions in the first stage and third stage in the six tests were the same,
respectively. The torque and turbocharger speed variations at the end of first stage
were within ±0.6% and ±2.1%, respectively. However, the maximum difference in
torque at the second stage was up to 17.8Nm (14.1%).

Figure 2 Torque responses of the six tests with different VGT settings.
The measured torque has been smoothed using 10-points smoothing.

With the VGT 60% closed in the second stage, torque increased immediately after
the restriction in the exhaust system was released due to the opening of the VGT.
The instantaneous turbine inlet pressures are shown in figure 4. With 60% closed
VGT, the turbine inlet pressure was approximately 0.2 bar lower than that with fully
closed VGT. This resulted in lower turbine total-to-static pressure ratio and the
higher engine volumetric efficiency, shown in figure 5. The calculated volumetric
efficiency was approximately 0.2 seconds delayed compared with the turbine
pressure ratio due to the distance between the flow meter and intake manifold.

166
Figure 3 Torque responses at the second stage (0.5 – 1.0 s).
The measured torque has been smoothed using 10-points smoothing.

Figure 4 Instantaneous turbine inlet pressures at the second stage of


three tests.

However, the turbocharger acceleration rate was 8.8% lower than that with fully
closed VGT. In addition, a torque drop can be observed when the VGT was returned
to 75% closed at the third stage, and approximately 0.35 second was required to
recover this torque drop. This was not preferable in terms of drivability. Compared
to other trajectories, the slow torque rise at the third stage was also a result of the
low turbocharger acceleration at the second stage. This delay was approximately
0.2 second at 1.5 second after tip-in.

On the other hand, if the fastest turbocharger acceleration was pursued, the torque
was dropped by approximately 10 Nm when the VGT was closed. Although the
turbocharger acceleration was the fastest, the drivability was also not acceptable.

Therefore, it shows a clear trade-off between the torque rise and turbocharger
acceleration which affects the torque response at the next stages in the transient
event. The optimised trajectory is also constrained by the drivability requirements.

167
Figure 5 Turbine total-to-static pressure ratio and engine volumetric
efficiency.

3.2 Response of turbocharger speed


On turbocharged gasoline engines, the engine torque response is largely affected
by the acceleration of the turbocharger. Therefore, turbocharger acceleration rate
was analysed.

At each of the three stages in the transient tests, for the tests with the same VGT
setting, first order and second order curves have been fitted to the turbocharger
speeds entering the stage and the turbocharger speed at the end of the stage. A
typical fitting at the third stage with the VGT 75% closed is shown in figure 6, the
R2 of the first order fitting and second order fitting are 0.9476 and 0.9478,
respectively. No significant difference between the first order and the second order
fitted curves was observed. The fittings at other stages with different VGT settings
are showing similar trends, and they are not presented.

Figure 6 Fitted responses of the turbocharger speed at 1.5 s and


turbocharger speed at 1.0 s with VGT 75% closed.

The fitted curves at all stages were then used to predict the turbocharger speed at
the end of transient test (1.5 second) using the averaged turbocharger speed
before tip-in as an input. The maximum deviation between the predicted
turbocharger speeds using second order fitting was 2.94%, lower than that using

168
the first order fitting which was 3.03%. However, the difference between the
predictions from the two fittings was below 0.77%. Therefore, the rise in
turbocharger speed at each 0.5 second stage in the first 1.5 seconds relatively
linearly dependents on the turbocharger speed at the time entering the stage and
the VGT position at the stage. In addition, the chosen stage duration, 0.5 second,
was small enough to optimise the transient operation.

It was also found that in all the three stages, the slopes of the fitted curves were
between 0.5 and 1. This illustrates that, although the higher turbocharger speed at
the start of a stage led to higher turbocharger speed at the end of that stage, the
turbocharger speed benefit gained from having a higher entering speed was smaller
than the difference in speed at the start of the stage.

3.3 Optimisation of the transient operation


To optimise the transient operation, two parameters were chosen as the indicators.
The engine torque rise was selected because this is the output of the engine. The
turbocharger acceleration was also selected because the T90 can not be reached
within 1.5 seconds, and the transient response after 1.5 seconds is largely affected
by the turbocharger acceleration. In addition, it has been illustrated that the
turbocharger speed at the end of a stage is strongly correlated to the turbocharger
speed at the start of the stage.

Firstly, the response in the first 0.5 second is analysed. The elliptical 50%
probability regions of torque and turbocharger speed rises in the first 0.5 second
have been plotted in figure 7. Each one of the six groups of data represents the test
result of 36 different VGT actuator trajectories that had the same VGT position in
the first 0.5 second. Despite the overlap between the groups because the VGT were
close to fully closed position, only 4 test points in total are outside the 75%
probability regions. Therefore, the repeatability of the test is acceptable.

Figure 7 Response curve between the turbocharger speed at the end of the
first stage and the torque rise in the first stage. The ellipses shown in the
figure represent the 50% probability region of the distribution of the data
having the same VGT position at the first stage.

A second order polynomial curve has been fitted to the mean values of torque and
turbocharger speed rises of the six data groups. The non-linearity of the VGT
turbocharger has been illustrated. Fully closing the VGT was not beneficial to both
the torque rise and turbocharger speed rise. Therefore, although the VGT
turbocharger is flexible to reduce the effective area and to accelerate the exhaust

169
gas velocity, the engine response in the first 0.5 second can not be improved by
fully closing the VGT. This may be a result of low turbine efficiency at fully closed
VGT position. However, it was difficult to calculate the turbine efficiency in a
transient test due to the response time of the thermocouples and the heat transfer
effect on turbocharger.

It is also illustrated that, because of the non-linearity of the VGT turbocharger,


overshoot in VGT position feedback control may result in both slow turbocharger
acceleration and slow torque response. Thus, to enable the use of VGT turbocharger
on gasoline engines, fine tuning of VGT controller and model-based control strategy
may be required.

On the other hand, opening up the VGT position to below 60% closed only gave
marginal benefit on the torque rise at the first stage, and the turbocharger speed
rise was significantly deteriorated. Therefore, the calibration of the strategy
depends on the requirement of the transient operation, which can be either a fast
torque rise or fast turbocharger acceleration at the first stage.

Secondly, the VGT positions at the second and third stage were optimised. The
turbocharger speed at 1.5 second is plotted against T50, shown in figure 8. As the
response in the first 1.5 seconds was concerned, tests in which 50% of the
maximum torque rise can not be reached within 1.5 seconds are not shown. A
Pareto optimal front was drawn. It was found that the highest Pareto efficiency can
be achieved by having a VGT position between 70% and 85% closed at the first
stage. Therefore, optimum turbocharger speed rise in the first 1.5 second and T50
can not be achieved with VGT position 60% closed at the first stage, although the
fastest torque rise in the first 0.5 second was reached.

Figure 8 Trade-off between the turbocharger speed at 1.5 s and T50.

It is worth noting that the baseline control strategy with the production FGT
turbocharger achieved T50 of 0.51 s and turbocharger speed of 113 krpm at 1.5
seconds, which can not be achieved by the VGT turbocharger with the tested
trajectories. It is likely to be a result of the differences in the turbocharger
matching. The VGT turbocharger was not matched to the tested engine. The
turbocharger characteristic maps are not available, therefore, it can not be
compared with the baseline production turbocharger. However, the methodology of
optimising the turbocharger control strategy is applicable to other applications.

170
The Pareto optimal curve is divided into three regions, shown in figure 9. In region
1 where the turbocharger speed at 1.5 second was relatively lower and the T50 was
shorter. The fastest T50 was achieved by opening the VGT to 60% closed at the
second stage to release the restriction in the exhaust system and to allow the 50%
of maximum torque rise to be reached within 1 second. A significant torque drop
can be observed on the torque responses of the majority trajectories in region 1,
shown in figure 10, due to the closing of the VGT to accelerate the turbocharger
after T50 was reached. The trajectory “90-85-60-70” was acceptable because the
VGT was not closed aggressively. With all the possible VGT settings at the third
stage, the turbocharger speed can not be recovered from the losses due to the
opening of the VGT at the second stage.

In the region 2, the two trajectories closed the VGT to intermediate positions (80%
and 75% closed) at the first stage and opened the VGT to 70% until T50 was
reached. No torque drop is observed.

In the third region of the Pareto optimal curve, the turbocharger acceleration was
the main target. Therefore, the VGT was kept between 80% and 75% closed in
most of the trajectories in region 3. The highest turbocharger speed at 1.5 second
was achieved in this region although the torque rise was slower.

Figure 9 Pareto optimal front of the trade-off between turbocharger


speed at 1.5 s and T50.

Therefore, a clear trade-off between the turbocharger speed at 1.5 seconds and T50
has been illustrated. The transient VGT control strategy can be optimised such that:

a. The minimum T50 was achieved by opening up the VGT and releasing the
restriction in the exhaust system. However, the turbocharger acceleration
was deteriorated and the gap can not be recovered.

b. The VGT was closed to intermediate positions at the first stage to


accelerate the turbocharger, and it was then opened mildly to allow both
acceptable turbocharger acceleration and torque build up.

c. The VGT position was kept at intermediate positions to pursue fastest


turbocharger acceleration.

171
Figure 10 Torque responses of the trajectories in Pareto optimal
curve region 1.

On the other hand, it was found that with the VGT 60% closed at the first stage,
which achieved the fastest torque rise in the first 0.5 second, the highest Pareto
efficiency can not be achieved. As the turbocharger speed at 1.5 second strongly
affects the time required to reach higher torque (for example 90% of maximum
torque rise), it is likely to cause a worse torque response at the later stage in the
transient.

Table 1 presents the optimisation of the VGT actuator trajectories targeting


different requirements of the transient operation. It was found that, despite the
flexibility of fully closing the VGT and reducing the effective area to accelerate the
exhaust gas and increase the kinetic energy, the turbocharger acceleration and
torque response can not be benefitted by fully closing the VGT at all the three
stages. The VGT position in the first 0.5 second was crucial, because the achievable
transient response was largely limited by the VGT setting at the first stage.

Table 1 Comparison of the optimised strategies

Requirements of transient response


Fast torque Fast T50 and Highest
response in High turbocharger turbocharger
Fastest T50
the first 0.5 speed at 1.5 speed at 1.5
second second second
VGT position at 75% - 80%
60% closed 85% closed 75% - 80% closed
first stage closed
VGT position at 60% closed 60% closed
70% closed before 75% - 80%
second and third before T50 before T50
T50 is reached closed
stage is reached is reached
Improvement of 14.4 Nm -2.5 Nm
torque rise in the - -
first 0.5 s (10.6%) (-1.8%)

0.51 s 0.71 s 0.54 s 0.25 s


T50 improvement
(33.9%) (47.1%) (35.5%) (16.5%)
Improvement of -3.8 krpm -1.7 krpm 0.5 krpm
turbocharger -
speed at 1.5 s (-6.4%) (-3.8%) (0.9%)

172
The torque responses of optimised VGT actuator trajectories are compared with a
single-step-change VGT actuator trajectory (fully closed before tip-in and 75%
closed after tip-in), shown in figure 11. The T50 was reduced by up to 47.1%
although the torque rise in the first 0.5 s and the turbocharger speed at 1.5 s were
worsened. If the turbocharger speed at 1.5 s was not to be compromised, the T50
can be improved by 35.5%. The turbocharger speed at 1.5 s can be improved by up
to 0.5 krpm while the T50 can still be improved by 16.5%. By pursuing the fastest
torque build-up in the first 0.5 s, the T50 improvement was limited at 33.9% and
the turbocharger speed at 1.5s was also lowered by 6.4%.

It should be noted that the duration of the three stages in the transient was 0.5
second, and the intervals between the optional VGT positions were 5-10%.
Therefore, larger improvement may be achievable if smaller time step and VGT
position interval were used.

Figure 11 Comparison of the optimised trajectories and the


reference trajectory.

4 CONCLUSION

VGT turbochargers have been used predominantly on Diesel engines rather than
gasoline engines due to the complexity of the mechanism and the lower exhaust
gas temperature limit. A VGT turbocharger has been tested on a 2.0L gasoline
engine. In order to avoid the constraints of using a feedback controller, a range of
open loop VGT actuator trajectories were investigated to optimise the first 1.5
seconds of transient operation during an engine load step change from 2 bar BMEP
to full load.

Compared with fully opening the VGT at low load operation point before tip-in, the
steady state turbocharger speed was increased by 11 krpm (138.4%) by fully
closing the VGT, whilst the fuel consumption was maintained the same considering
the measurement accuracy. However, fully closing the VGT during the transient
resulted in high back pressure and low volumetric efficiency. Thus, both the torque
response and the turbocharger acceleration were worsened.

A trade-off between the torque rise in the first 0.5 second, T50, and the
turbocharger speed at 1.5 second has been illustrated in this study. The optimum
calibration depends on the requirement of the transient operation. Compared to a
single-step-change VGT actuator trajectory, the torque rise in the first 0.5 second

173
can be improved by 14.4 Nm (10.6%) at the expense of lowering the turbocharger
speed at 1.5 second by 3.8 krpm (6.4%), and the T50 can be improved by 0.71 s
(47.1%) at the expense of lowering the turbocharger speed at 1.5 second by 1.7
krpm (3.8%). If no turbocharger acceleration compromise was accepted, the T50
can be improved by 0.54 s (35.5%). The turbocharger speed at 1.5 second can be
increased by up to 0.5 krpm (0.9%), while the T50 can still be improved by 0.25 s
(16.5%).

The VGT position in the first 0.5 second was crucial, because the achievable
transient response was largely limited by the VGT setting at the first stage.
Compared with a conventional boost pressure feedback controller, which may fully
close the VGT during transient, a model based transient control strategy may be
required to improve the transient operation and to avoid overshoot.

5 ACKNOWLEDGEMENT

The data shown in this paper was from a collaborative research project,
TurboCentre, between the University of Bath, Jaguar Land Rover Limited, Ford
Motor Company Ltd and Cummins Turbo Technologies Ltd. The project was funded
by the University of Bath’s EPSRC Knowledge Transfer Account and project partners.
The experimental work was carried out in the Powertrain & Vehicle Research Centre
in the University of Bath.

The authors would like to thank all the organisations in this project for the
permission to publish this paper.

6 REFERENCES

[1] Guzzella, L., Wenger, U. and Martin, G., 2000, IC-Engine Downsizing and
Pressure-Wave Supercharging for Fuel Economy, SAE technical paper 2000-
01-1019
[2] Bauer, K.-H., Balis, C., Donkin, G. and Davies, P., The Next Generation of
Gasoline Turbo Technology, 33rd International Vienna Motor Symposium
2012, Vienna 2012
[3] Wang, L.-S. and Yang, S., 2006, Turbo-Cool Turbocharging System for Spark
Ignition Engines, Proceedings of the Institution of Mechanical Engineers, Part
D: Journal of Automobile Engineering, Vol. 220(8), pp. 1163-1175
[4] Lundstrom, R. R. and Gall, J. M., 1986, A Comparison of Transient Vehicle
performance Using a Fixed Geometry, Wastegated Turbocharger and a
Variable Geometry Turbocharger, SAE technical paper 860104
[5] Singer, D. A., 1985, Comparison of a Supercharger vs. a Turbocharger in a
Small Displacement Gasoline Engine Application, SAE technical paper
850244
[6] Arnold, S., Groskreutz, M., Shahed, S. M. and Slupski, K., 2002, Advanced
Variable Geometry Turbocharger for Diesel Engine Applications, SAE
technical paper 2002-01-0161
[7] Gabriel, H., Jacob, S., Munkel, U., Rodenhauser, H. and Schmalzl, H. P., 2007,
The Turbocharger with Variable Turbine Geometry for Gasoline Engines, MTZ,
Vol. 68(2), pp. 96-103
[8] Capobianco, M. and Gambarotta, A., 1992, Variable Geometry and Waste-
Gated Automotive Turbochargers - Measurements and Comparison of Turbine
Performance, Journal of Engineering for Gas Turbines and Power-
Transactions of the ASME, Vol. 114(3), pp. 553-560

174
[9] Lezhnev, L., Kolmanovsky, I. and Buckland, J., 2002, Boosted Gasoline Direct
Injection Engines: Comparison of Throttle and VGT Controllers for
Homogeneous Charge Operation, SAE technical paper 2002-01-0709
[10] Miller, K., 2008, Turbo: Real World High-Performance Turbocharger Systems:
S-A Design, ed. North Branch: CarTech Inc.
[11] Flardh, O. and Martensson, J., Nonlinear Exhaust Pressure Control of an SI
Engine with VGT Using Partial Model Inversion, 49th IEEE Conference on
Decision and Control, Atlanta, USA, 2010
[12] O'Connor, G. and Smith, M., 1988, Variable Nozzle Turbochargers for
Passenger Car Applications, SAE technical paper 880121
[13] Hirhikawa, A., Okazaki, Y. and Busch, P., 1988, Developments of Variable
Area Radial Turbine for Small Turbochargers, SAE technical paper 880120
[14] Kawaguchi, J., Adachi, K., Kono, S. and Kawakami, T., 1999, Development of
VFT (Variable Flow Turbocharger), SAE technical paper 1999-01-1242
[15] Rogo, C., Hajek, T. and Roelke, R., 1983, Aerodynamic Effects of Movable
Sidewall Nozzle Geometry and Rotor Exit Restriction on the Performance of a
Radial Turbine, SAE technical paper 831517
[16] Umezaki, E., Ogura, M. and Tomita, T., 1989, Study of Variable Scroll Type
Turbocharger (Determination of Shape of Scroll), SAE technical paper
891874

175
The exhaust energy utilization of a
turbocompound engine combined with
divided exhaust period
H Aghaali, H E Ångström
KTH - Royal Institute of Technology, CCGEx, Internal Combustion Engines, Sweden

ABSTRACT

To decrease the influence of the increased exhaust pressure of a turbocompound


engine, a new architecture is developed by combining the turbocompound engine
with divided exhaust period (DEP). The aim of this study is to utilize the earlier
stage (blowdown) of the exhaust stroke in the turbine(s) and let the later stage
(scavenging) of the exhaust stroke bypass the turbine(s). To decouple the
blowdown phase from the scavenging phase, the exhaust flow is divided between
two different exhaust manifolds with different valve timing. A variable valve train
system is assumed to enable optimization at different load points. The fuel-saving
potential of this architecture have been theoretically investigated by examining
different parameters such as turbine flow capacity, blowdown valve timing and
scavenging valve timing. Many combinations of these parameters are considered in
the optimization of the engine for different engine loads and speeds.

This architecture produces less negative pumping work for the same engine load
point due to lower exhaust back pressure; however, the exhaust mass flow into the
turbine(s) is decreased. Therefore, there is a compromise between the turbine
energy recovery and the pumping work. According to this study, this combination
shows fuel-saving potential in low engine speeds and limitations at high engine
speeds. This is mainly due to the choked flow in the exhaust valves because this
approach is using only one of the two exhaust valves at a time. To reveal the full
potential of this approach, increasing the effective flow area of the valves should be
studied.

1 INTRODUCTION

The foreseeable shortage of fuel and the discussion of global warming have
heightened the need for designing more fuel efficient engines. To reduce the fuel
consumption, one of the main developments in internal combustion engines is
waste heat recovery. This can be done by converting exhaust gases heat to
mechanical work in a turbine. This is called turbocompounding. However,
turbocompound turbines create pumping loss in the engines [1,2]. In addition to
the turbocompounding, DEP is another way to reduce fuel consumption that
decreases mainly the pumping loss in engines [3-7].

Turbocompound engines have been widely investigated and it has been shown that
they have fuel-saving potential in the range of 1-5% [1,2]. There have been a few
investigations into the DEP engine that all studies are about a turbocharged engine
equipped with DEP [3-7]. Turbocharged engine with DEP has a limited fuel saving
potential, as well. The main drawback of turbocompound engines is the high

_______________________________________
© The author(s) and/or their employer(s), 2014
179
pumping loss. On the other hand, the main advantage of DEP architecture is the
reduced pumping loss.

Therefore, extending previous research by combining a turbocompound engine with


DEP is the topic for this analysis with the aim of reducing pumping loss while
energy recovery in the turbine(s) is maintained.

This paper gives preliminary results for fuel-saving potential of a turbocompound


engine with DEP. This study was designed to evaluate the role of the DEP
architecture on a turbocompound engine keeping the cylinder head geometry, the
turbine efficiency and the boost pressure unchanged.

The hypothesis is that a turbocompound engine combined with DEP is more fuel
efficient because of the reduced pumping loss. However, the turbine energy
recovery might be lower due to the reduced amount of exhaust flow into the
turbine. In this paper, the hypothesis will be analysed on one engine load point;
then more load points will be examined.

2 SIMULATION

The simulations have been performed in GT-Power [8] on a heavy-duty Diesel


engine for three different architectures, turbocharged engine, turbocompound
engine, and a combination of turbocompound and DEP.

GT-Power is an 1D fluid dynamic tool with engine flow models. It uses maps, lookup
tables and empirical models for some components such as valves, turbochargers
and cylinders. The studied engine is 11.7 liter, in-line six cylinder heavy-duty Diesel
engine. The engine specifications are provided in Table 1. This engine incorporates
a twin scroll turbine and no EGR system, which was chosen to keep the complexity
of the initial study at a minimum. Meanwhile, Kruiswyk [2] has concluded that
elimination of EGR on a turbocompound engine improves the BSFC benefit by 1-
1.5% at high loads. The baseline model for this engine was calibrated in a previous
work [9] and it was used as the base for the simulations. To model the combustion
under different engine speeds and torques, heat release rates based on measured
cylinder pressures were used.

The turbocharged engine was then modified in the simulations to be equipped with
a compound turbine, variable valve actuation (VVA) and exhaust manifolds for DEP.
Firstly; the turbocharger turbine and compressor were disconnected from each
other. The turbine shaft was connected directly to the engine crankshaft by a
continuously variable transmission (CVT). This was done for the compressor shaft,
as well. This gave the possibility to keep the boost pressure constant for each load
point while the turbine swallowing capacity was changed. This is a single-stage
turbocompound engine that has been simulated previously with electric generator
and motor instead of mechanical transmissions [1]. Secondly; the turbocompound
engine was equipped with a VVA for each exhaust valve and two separate exhaust
manifolds, as illustrated in Figure 1. The size of the exhaust manifolds and their
properties have been extracted from a previous study [10]. Figure 1a shows that
one port of each cylinder is connected to the blowdown manifold and the other port
is connected to the scavenging manifold. The blowdown manifold feeds the turbine
and the scavenging manifold bypasses the turbine. Figure 1b shows the timings of
the exhaust and intake valves. The turbocompound engine without DEP uses the
ordinary intake and exhaust valves timings as the turbocharged engine. However,
the turbocompound engine with DEP employs the blowdown and scavenging valves
timings instead. It has to be mentioned that the collision between the scavenging
exhaust valve and piston is not considered in this study.

180
Table 1. Engine specifications.

Engine label SCANIA DC1201


Emission class Euro 3
Max. Power [hp] 380
Max. Torque [Nm] 1900
Displacement [dm3] 11.7
Bore [mm] 127
Stroke [mm] 154
Conn. rod length [mm] 255
Compression Ratio 18
IVO [°ATDC] 346
IVC [°ATDC] 142
EVO [°ATDC] 136
EVC [°ATDC] 359
Turbocharger Twin-entry
EGR System No

(a) (b)
Figure 1. a) The architecture of turbocompound engine combined with
DEP; b) Variable valve timing including blowdown and scavenging
exhaust valves.

The simulation approach was to minimize the break specific fuel consumption
(BSFC) of the engine in different architectures. This has been performed for several
engine loads and speeds, as illustrated in Figure 2.

Figure 2. Chosen load points of the original turbocharged engine.

181
Since the current study aims at comparing the simulated turbocharged engine with
the modified models, any discrepancy from measurements is assumed to equally
affect the models. At each load point, some parameters are kept unchanged and
constant based on the original turbocharged engine for all three architectures such
as:

– Power output of the total system


– Engine speed
– Turbine efficiency
– Turbine speed
– Compressor efficiency
– Compressor speed
– The size of the intake valves
– The size of the exhaust valves
– Intake valve opening time
– Intake valve closing time
– Exhaust (blowdown in DEP architecture) valve opening time
– Exhaust (scavenging in DEP architecture) valve closing time
– Frictional mechanical efficiency of turbocharger = 100%
– Mechanical efficiency of CVT = 100%

Since the turbine efficiency is constant in each load point for the three
architectures, dissimilar pulsating flows will not disturb the results. The research
has focused on the created exhaust back-pressure in different architectures, rather
than on the pulsating flow in the turbine. Different exhaust back pressures will
make altered turbine power and pumping work.

In the simulation of the turbocompound engine, the exhaust back-pressure was


varied by scaling the swallowing capacity of the turbine. This can be done by
varying turbine mass multiplier in GT-Power. The optimized exhaust back pressure
of the turbocompound architecture which gives minimum BSFC has been calculated
for each load point.

In the simulation of the turbocompound engine combined with DEP, in addition to


the turbine swallowing capacity, the blowdown exhaust valve closing time and the
scavenging exhaust valve opening time are varied. To find the optimum
combination of these three parameters which gives the minimum BSFC of the last
architecture, a full factorial design of experiment (DOE) has been run. For each load
point, 245 cases have been considered that contain 5 levels of the turbine mass
multiplier, 7 levels of the blowdown exhaust valve closing time and 7 levels of the
scavenging exhaust valve opening time. Then, a model-based optimization has
been performed using GT-POWER to minimize the BSFC response. The goodness of
the fits has been checked for all load points by calculating the coefficient of
determination (R squared). They were all very close to 1 for all load points. The
modelled parameters for each load point have been extracted and all load points
are run again to be sure that the modelled parameters provide the same responses.

3 RESULTS

3.1 Fuel-saving potential at high load and low speed


Figure 3 shows the BSFC of three different architectures at a high engine load and
low speed (A3 on Figure 2). The result shows that the turbocompound engine has
lower BSFC than the turbocharged engine while the efficiency of the turbine is kept
constant. The combination of turbocompound and DEP has the lowest BSFC at this
load point with the same turbine efficiency, boost pressure and exhaust valves
geometry.

182
Figure 3. Fuel consumption of different engine architectures,
turbocharged, turbocompound and turbocompound with DEP.

To explain the potential of this architecture in fuel-saving, the trade-off between


the surplus turbine power and pumping work shall be examined. The surplus
turbine power is the difference between the turbine power and the compressor
power. On the single-stage turbocompound engine in which the turbine and the
compressor are mechanically connected to the engine crankshaft, the total power of
the system is the summation of net engine power and the surplus turbine power.

Figure 4. Normalized surplus turbine power by engine power vs.


normalized PMEP by BMEP of load point A3 for three different
architectures.

Figure 4 shows the normalized surplus turbine power by the total engine power
against the normalized pumping mean effective pressure (PMEP) by break mean
effective pressure (BMEP) of the load point A3 for the three different architectures.
In this load point the total engine power, consequently BMEP, is kept constant.
Therefore, this figure shows the relation between the surplus turbine power and
PMEP. When the surplus turbine power is zero, it means that the engine is
turbocharged; however, positive and negative signs of the surplus turbine power
indicate turbocompound and supercharged engines, respectively. The turbocharged

183
engine has zero surplus turbine power and almost +2% “PMEP/BMEP”. In a
turbocompound engine, the exhaust back pressure should be increased to create
more turbine power; however, the pumping loss will be increased. The increased
exhaust back pressure has an optimum point where the BSFC of the engine is
minimal. This point is marked along the line of the increased back pressure. For the
turbocompound architecture at load point A3, +6.3% “surplus turbine power/engine
power” is provided while -2.5% “PMEP/BMEP” is a result of the increased back
pressure. However, in total turbocompound engine is clear benefit compared to the
turbocharged engine, because the difference of “PMEP/BMEP” from turbocharged to
turbocompound is -4.5% while “surplus turbine power/engine power” is changed
+6.3%. Besides, the optimum case of the combined turbocompound engine with
the DEP architecture is marked around +4.9% “surplus turbine power/engine
power” and almost -0.8% “PMEP/BMEP”. This means -2.8% difference in
“PMEP/BMEP” compared to turbocharged one. If we compare this point with the
turbocompound one the difference between the “surplus turbine power/engine
power” is almost -1.4% while “PMEP/BMEP” difference is about +1.6% which means
turbocompound engine with DEP consumes less fuel. This figure provides a simple
way to compare different architectures in terms of surplus turbine power and
pumping work. On this figure, it is preferred to move toward more surplus turbine
power and higher positive PMEP.

Although the turbocompound engine with DEP provides less negative PMEP
compared to the turbocompound one, the turbine energy recovery is less. This is
because the exhaust flow from the cylinders to the exhaust system is divided into
two exhaust manifolds and the turbine is just fed by one of them. Therefore, the
exhaust mass flow through the turbine is less. In addition to the less exhaust mass
flow, the optimum exhaust back pressure is lower. As a result, the turbine energy
recovery is less in a turbocompound engine combined with DEP compared to the
turbocompound engine.

The main concerning issue regarding the turbocompound engine with DEP
architecture is the choked flow through the exhaust valves due to the small
effective flow area of the exhaust valves. Figure 5 shows the Mach number of
exhaust flow through the valves for three different architectures at load point A3.

Figure 5. Mach number through the exhaust valves vs. engine crank
angle for three different architectures at load point A3.

184
Apparently, the turbocharged engine is restricted by the choked flow during the
beginning period of the exhaust valve opening. This is shorter for turbocompound
engine due to the higher exhaust back pressure. Thus the pressure ratio of the
cylinder contents to the exhaust system is smaller; therefore, the choked flow is
shorter. In the turbocompound engine with the DEP architecture, firstly the
blowdown valve is opened with half area compared to the original engine. This
means the effective flow area is much smaller while the cylinder contents have
higher time-averaged temperature and higher time-average pressure for the
blowdown period than for the scavenging period. This leads to longer choked flow.
Before the blowdown valve approaches the closing, the scavenging valve starts
opening with an overlap. Due to small effective flow area at the beginning of the
opening, choked flow happens at the scavenging valve as well. These results differ
for another load point and in some points this will be much worse in term of long
choked flow. We should keep in mind that the provided results are at load point A3
where this combination is beneficial. Therefore, reducing the time of choked flow in
the turbocompound engine with the DEP architecture can reveal its fuel-saving
potential more while longer choked flow can happen in some load points that will be
discussed later.

Figure 6. Cylinder pressure - volume diagram of load point A3


for three different architectures; logarithmic scales.

To compare the cylinder pressures in these three architectures, Figure 6 shows the
cylinder pressure versus normalized cylinder volume by the maximum cylinder
volume for load point A3 with logarithmic scales. As it is clear, the intake, the
compression, the combustion and the expansion strokes are almost the same for all
architectures because the boost pressure and compression ratio are kept constant.
The only difference appears during the exhaust stroke due to different exhaust back
pressures. The turbocharged engine in this load point has lower exhaust back
pressure than the turbocompound one. Therefore, the pumping work is positive in
the turbocharged engine while this is negative in the turbocompound engine. The
turbocompound engine with DEP has higher exhaust back pressure remarkably
during the first period of the exhaust stroke (blowdown) than the turbocompound
engine; however, this falls suddenly to lower than the intake pressure and the
exhaust pressure of the turbocharged engine during the rest of the exhaust stroke.

185
This leads to almost zero pumping work while the blowdown pressure is utilized in
the turbine. This is the competence of this combination.

3.2 Fuel-saving potential of all chosen load points


Figure 7 shows the BSFC improvements of the turbocompound engine with and
without the DEP architecture compared to turbocharged engine as a base for
different load points in Figure 2. As shown in Figure 7, the turbocompound
improves the turbocharged engine from 0 to 2.2% in term of BSFC. This
improvement is mainly at high loads. However, the combination of turbocompound
and DEP does not improve BSFC in all load points and in some points it makes even
very much worse condition like 4.5% deterioration compared to turbocharged
engine in A13. The only points that gain from this combination are the cases at low
engine speeds. The BSFC improvements in cases with medium engine speed are
almost zero. The cases with high engine speeds are declined by this combination.

Figure 7. BSFC improvement of chosen engine load points for


different architectures compared to the turbocharged engine;
Constant parameters have been changed for each load point
based on the original turbocharged engine.

As an example, Figure 8 depicts normalized surplus turbine power against


normalized PMEP of one load point which gains no improvement (A7) for three
different architectures. The line on the figure shows the direction of increased
exhaust back pressure for the turbocompound engine. The optimum exhaust back
pressure of the turbocompound engine is marked on the line. The combined
turbocompound and DEP could not lie on the right side of the line where the trade-
off between the surplus turbine power and PMEP is beneficial. Therefore, in spite of
the optimization in the turbocompound and DEP architecture, there is no gain
compared to the turbocharged engine and the turbocompound architecture is
preferred in this load point.

186
Figure 8. Normalized surplus turbine power by engine power vs.
normalized PMEP by BMEP of load point A7 for three different
architectures.

Figure 9 illustrates the summary of this study’s results on the engine map. The
turbocompound engine is beneficial on just high loads and at all engine speeds.
However, the combination of turbocompound and DEP gives improvements just in
low engine speeds. This combination is not beneficial at the medium and high
engine speeds. However, using a turbine with higher efficiency and modifying the
exhaust valves would extend the working range of this combination.

Figure 9. The potential working range of the turbocompound engine and


the combined turbocompound engine with DEP in term of fuel-saving.

187
4 CONCLUSIONS

DEP architecture was introduced to a turbocompound engine to enhance the fuel-


saving potential of the engine. This new approach is beneficial by separating the
exhaust stroke into two periods and reducing the pumping loss. The exhaust flow
feeds the turbine during the blowdown period and it bypasses the turbine during
the scavenging period. As a preliminary result, the combined turbocompound
engine with DEP improves BSFC 0.5 to 3% in low engine speeds if the cylinder head
geometry, turbine efficiency and boost pressure are kept unchanged compared to
the original turbocharged engine. However, this combination deteriorates the fuel
consumption at medium and high engine speeds. Since in this study, several
parameters and geometries are kept unchanged from the original turbocharged
engine, the shown improvement is just due to the different exhaust back pressure
and different exhaust flow through the turbine. This means that if we could change
the geometries, get higher efficiencies of turbine and compressor, and achieve
better turbocharger matching or turbine selection, then the fuel saving potential
should be higher and its working range would extend to higher engine speeds. The
main limitation in this approach is the choked flow through the exhaust valves
which could be an interesting topic for future work.

ACKNOWLEDGEMENTS

The Swedish Energy Agency (Energimyndigheten) and Royal Institute of Technology


(KTH) sponsored this work within the Competence Centre for Gas Exchange
(CCGEx).

REFERENCES

[1] Aghaali, H., and Angstrom, H., Demonstration of Air-Fuel Ratio Role in One-
Stage Turbocompound Diesel Engines, SAE 2013-01-2703.
[2] R. Kruiswyk, "The role of turbocompound in the era of emissions reduction,"
10th International Conference on Turbochargers and Turbocharging, 2012.
[3] Möller, C., Johansson, P., Grandin, B., and Lindström, F., "Divided Exhaust
Period - A Gas Exchange System for Turbocharged SI Engines," SAE Technical
Paper 2005-01-1150, 2005, doi:10.4271/2005-01-1150.
[4] Roth, D., Keller, P., and Sisson, J., "Valve-Event Modulated Boost System,"
SAE Technical Paper 2010-01-1222, 2010, doi:10.4271/2010-01-1222.
[5] Roth, D. and Becker, M., "Valve-Event Modulated Boost System: Fuel
Consumption and Performance with Scavenge-Sourced EGR," SAE Int. J.
Engines 5(2):538-546, 2012, doi:10.4271/2012-01-0705.
[6] Gundmalm, S., Cronhjort, A., and Angstrom, H., Divided Exhaust Period on
Heavy-Duty Diesel Engines, THIESEL 2012 Conference on Thermo- and Fluid
Dynamic Processes in Direct Injection Engines, 2012.
[7] Gundmalm, S., Cronhjort, A., and Angstrom, H., "Divided Exhaust Period:
Effects of Changing the Relation between Intake, Blow-Down and Scavenging
Valve Area," SAE Int. J. Engines 6(2):739-750, 2013, doi:10.4271/2013-01-
0578.
[8] G. T. Inc., GT-SUITE, Flow Theory Manual 7.3, http://www.gtisoft.com.
[9] Winkler, N., Transient simulations of heavy-duty diesel engines with focus on
the turbine, Licentiate Thesis, Royal Institute of Technology, Stockholm,
Sweden, 2008.
[10] Gundmalm, S., Divided Exhaust Period on Heavy-Duty Diesel Engines,
Licentiate Thesis, Royal Institute of Technology, Stockholm, Sweden, 2013.

188
Further development of two-stage
turbocharging systems for large engines
E Codan, C Christen
ABB Turbo Systems Ltd, Switzerland

ABSTRACT

Two-stage turbocharging is a logical development step for large combustion


engines. Several studies have already been published showing the large
improvement potential concerning engine efficiency, emissions and power density.
In this paper some aspects of the design of ABB’s new generation two-stage
turbocharging are presented. The choice of basic design as well as the achieved
performance are discussed. The expected engine results with the new
turbocharging system are presented and discussed. The focus is set on diesel
engines with strong Miller timing and variable valve timing, showing the potential
for steady state and transient operation.

NOMENCLATURE

be Specific fuel consumption (g/kWh) Subscripts


D Diameter (m) ac Start of compression
KJ Specific moment of inertia (kg/m3) rec charge air receiver
n Engine speed (rpm) TI Turbine inlet
p Pressure (Pa, bar) CI Compressor inlet
pmax Firing, maximum pressure (bar)
bmep Brake mean effective pressure (bar) Abbreviations
T, t Temperature (K, °C) C Compressor
V298, ܸሶ Reduced volume flow (m3/s) HP High pressure
 Flow coefficient LP Low pressure
 Efficiency T Turbine
VVT Variable valve timing
C Mass ratio of trapped to
stoichiometric air
aC Specific compressor work: enthalpy
head/peripheral speed squared
 Pressure ratio
 Specific mass (kg/m3)
 Time constant (s)

1 INTRODUCTION

Main drivers in the development of large engines are emissions, operating costs
and first cost.

Compliance with the emission limits is obviously a must for the commercial
application of an engine. The NOx limits will be substantially reduced in the near

__________________________
© ABB Turbo Systems Ltd, 2014
189
future leading to the necessity to adopt after treatment (SCR – Selective Catalytic
Reduction), exhaust gas recirculation (EGR) or gaseous fuel. Other possibilities are
currently investigated on a research level but are not yet validated. The SOx
emission will also be limited, leading to the necessity to use low sulfur fuels or
devices able to reduce the SOx content of the exhaust gas.

Operating costs are currently mainly defined by fuel costs, which have increased at
least by a factor of five in the recent years. Engine efficiency will become ever more
important. Additionally, engine efficiency correlates with CO2 emissions, which are
monitored with the introduction of control parameters such as the Energy Efficiency
Design Index (EEDI). Maintenance costs are a smaller component of the operating
costs, but ease of service is very important for reliability and performance stability,
which helps to keep a high efficiency over the whole life cycle of the power plant.

First cost of a power plant is influenced by many different factors, but for sure one
of them is the power density, given by the product of mean piston speed and mean
effective pressure. On diesel engines, it is not expected that power density will be
considerably increased. On research engines, values of mean effective pressures of
up to 40 bar have been explored. The technical feasibility was proven, but the
development effort and the challenges of keeping efficiency and emissions on good
levels limit the commercial feasibility. The situation on gas engines is different,
because the knock limit is the main limitation for the power density of highly
turbocharged lean-burn gas engines. The Miller process is here a very important
because it allows the knock limit to be shifted, and the power gap with diesel
engines to be closed.

High-pressure turbocharging in combination with the Miller process can offer


significant improvements for all the aspects mentioned above. The ABB A100
turbocharger family has proven its potential for fulfilling today’s emission limits
efficiently without additional measures. Even more potential can be disclosed by
applying two-stage turbocharging and variable valve timing. This potential has been
shown in several studies performed by means of advanced simulation tools and
confirmed by engine tests (1) (2) (3) (4) (5) (6) (7) (8). These studies show that
there is substantial improvement potential for engine performance due to the
reduction of the process temperatures offered by the Miller process. This
temperature reduction has a very positive effect on the engine’s efficiency and its
raw NOx-emissions. At the same time, the high turbocharging efficiency achievable
with two-stage turbocharging allows the efficient delivery of the required air
pressure and additional power due to the improvement in gas exchange work.

1.1 First generation Power2®


The first generation of ABB’s two-stage turbocharging solution, Power2, has been
successfully in use since 2010 (1) (2) (7). Two system frame sizes have been
released for serial production. Until today, about 50 Power2 systems have been
taken into operation. At this point in their evolution, front runner systems could be
operated for more than 10,000 hours. Engine efficiency gains and emissions
reduction clearly beyond the potential of any single-stage system have been
demonstrated. Forecasts were even exceeded.

The first generation was developed for a target range of pressure ratio between 7
and 9 with equivalent turbocharger efficiency above 73% using whenever possible
components taken from normal single-stage turbochargers. Since these
components have not been originally designed for the requirements of a two-stage
system, it is clear that some further potential would be available with a specific new
development.

190
2 POWER2 – SECOND GENERATION

Looking at the design of the two-stage turbocharging system for four-stroke


engines under the aspect of fuel efficiency only, the target pressure ratio might be
about 8. If we take into account the trade-off between efficiency and emissions, as
well as the requirement to improve power density and altitude capability, a single
figure compression ratio would not be the right solution. For the new generation
Power2 a wide range of pressure ratios from 8 to 12 was set as design target. In
this range, equivalent turbocharger efficiency over 75% with an intercooler
temperature of 70°C was set as another requirement.

Keeping in mind the overall performance to be achieved, additional objectives,


which will be illustrated below, were defined for the design.

2.1 Ease of service


Engine availability is a key factor for achieving optimal economic performance.
Consequently, the time required for service work needs to be reduced to a
minimum. During development of the new turbocharger generation, service
friendliness was considered from the very beginning. The designated goal was to
reduce service time of the complete two-stage system below the reference value of
current single-stage turbochargers.

Figure 1: Extractable cartridge concept

Today’s ABB turbochargers for medium-speed engines feature the cartridge group.
The cartridge consists of rotor, bearings and directly related casings. After removal
of the air casings it is possible to perform a quick exchange of the cartridge from
the cold side of the turbocharger, without touching the hot parts.

For the second generation Power2 turbochargers the extractable cartridge concept
was developed (Figure 1). The idea behind this concept is that the turbocharger has
an outer shell, consisting of air and gas casings, and a cartridge group, which
contains the entire interior of the turbocharger. In order to exchange the cartridge
during service, only the air inlet casing together with the insert wall of the
compressor needs to be removed. All other interfaces with the engine as well as the
insulation remain untouched. A dedicated tool was developed together with the
extractable cartridge concept, which enables quick exchange of the cartridge.

Another contribution to engine downtime reduction is the inspection of


turbochargers by endoscopy, which allows a status check of the components
without removing any part of the turbocharger.

2.2 Optimum performance


The overall performance of the turbocharging system highly depends on the design
of its core components, compressor and turbine stages. But for the global
optimization many other aspects must be considered.

191
All connections in the casings of
a two-stage system have to
guide a volumetric flow, which
is considerably higher than in a
conventional turbocharger.
Therefore, all the casings have
been optimized for low pressure
losses without exerting negative
influence on the stage operation
by means of CFD calculations.
a) b)
The result is not only a better
system efficiency, but also Figure 2: LP gas inlet casing
reduced dimensions, as the a) initial design, b) optimized design
example in Figure 2 shows.

Performance stability in operation, especially for engines burning heavy fuel oil is
another issue. Regular turbine and compressor cleaning are effective measures to
assure this performance stability. The location and shape of the injection holes for
water cleaning have been optimized by means of two-phase flow CFD simulations.

The high-pressure level of the HP turbocharger as well as the presence of two


turbocharger shafts would lead to a considerable increase of the air leakage into the
engine oil system (blow-by). This is a highly undesired effect of the turbocharging
system, so a new sealing concept was developed, which leads to a 75% reduction
of the air leakage compared to state-of-the-art labyrinth seals.

In order to optimize the bearing losses different thrust bearings have been
developed for the HP turbocharger, which are dimensioned for the specific thrust on
different applications.

2.3 Compactness
Almost every engine has its turbocharger mounted on it to form a complete unit.
Retaining this concept in the two-stage system was a key priority. Since one
turbocharger is replaced with two, it is obvious that the dimensions must be kept as
small as possible. For this reason very high specific flow was set as target for all
components. In order to illustrate the achievements the operating envelopes of the
new developed components was plotted in the diagrams showing the historical
development of ABB compressors (Figure 3).

6
2009 1-stage compressors
C 2009

2003
5 2004

2003
1996 LP-compressor
1996
2012
4
1992
1978

1989 1983
1970
3
1954-1964

1946 2010-2013
2
HP-compressor

1924
1
V298 [m3/s]

Figure 3: Compressor development

192
The position of the HP compressor,
working at high specific flow and
moderate pressure ratio has already
been shown (5), because the
development was necessary already
for the first generation. Completely
new is the LP compressor, which
reaches a range of pressure ratio
closer to that of single-stage
machines but with a very high
specific flow. The compressor map is
plotted in Figure 4 in comparison with
the compressor stage of the A100,
which reaches much higher pressure
ratios but at lower specific flow. A
second HP compressor stage was
developed for applications requiring
high overall pressure ratios.

On the turbine side the situation is


similar. Very high specific flow at
moderate expansion ratio is required Figure 4: Compressor performance
for the HP stage. The LP stage maps of A100-M (single-stage)
operates at high expansion ratio and and Power2 800-M LP (two-stage)
a high efficiency is required. Thanks at equal impeller wheel diameter
to the lower exhaust gas temperature
after the first expansion, the specific
flow requirements are lower.

An important decision was to adopt axial turbines for both stages to cover the
application on medium-speed engine with power between 3 and 10 MW per
turbocharging group. The reasons are the superior performance of the axial
turbines at high specific flow and moderate expansion ratio as well as the better
suitability for HFO application.

The first prototype of second-generation Power2 can be seen in Figure 5.

Figure 5: Power2 850-M

193
3 POWER2 APPLICATION

Possible thermodynamic achievements with high-pressure turbocharging have


already been shown in several theoretical studies. An example is given in Figure 6
(5) where the effects of the increase of boost pressure are shown for applications
with single-stage and two-stage turbocharging for a medium-speed diesel engine.
Every point of the curves represent an engine setting for constant output (bmep =
25 bar), firing pressure (200 bar), air-fuel ratio (C = 2.2) and exhaust gas
temperature (tTI = 520°C). The green area represents the range of results expected
for higher levels of bmep and turbocharging efficiency. It can be seen, that for the
design point there is a potential for improving the specific fuel consumption and
that the NOx emissions are almost linearly decreasing with the boost pressure.
These effects are due to the positive effects of reducing the cycle temperature by
means of the Miller process on the thermal efficiency as well as on the NOx
formation. Furthermore, the gas exchange cycle is improved due to the increased
turbocharging efficiency. Potential for power increase had not yet been considered
at this stage.

110 110
be NOx
1-stage_be_rel 2-stage_be_rel
[% ] [% ]
105 100
1-stage_NOx_rel 2-stage_NOx_rel
A
100 90
B

95 80
C
D
90 70

85 60

80 50
4 5 6 7 8 9 10 11 12
 C*

Figure 6: NOx- and specific fuel consumption (be) over pressure ratio
(C*) for single-stage and two-stage turbocharging systems

The diagram was built with simulations with partly idealized components and
considering the design points only. In order to better illustrate a possible
development path, the points A and B for single-stage turbocharging, and the
points C and D for two-stage turbocharging have been chosen and more detailed
simulations have been performed with performance maps of real components. The
turbocharger have only been scaled in order to work in every case with a
comparable loading. The diagram
200
in Figure 7 shows the evolution of LP Compressor diameter
the turbocharger dimensions, [%]
HP Compressor diameter
represented by the compressor 150
Volume
wheel diameter. It was assumed
that the volume is proportional to 100
Volume

the third power of the linear


Volume

Volume

dimension. It can be seen that the


Volume

50
pressure increase with single-
stage turbochargers, case A to
case B requires an increase of the 0
Case A Case B Case C Case D
dimensions. The change to two-
stage turbocharging at the same Figure 7: Turbocharger dimensions

194
pressure ratio implies a further considerable growth of the system dimensions, but
further increasing the pressure ratio, as from case C to case D, brings a reduction,
leading to comparable dimensions as case B. A two-stage system has additional
piping and a second air cooler, but by a consistent design taking into account the
air density it can be further assumed that the volume of the turbocharging system
is proportional to the volume of the turbochargers.

3.1 Case study medium-speed engine


The engine model is the same that was used for the simulations in Figure 6, i.e. a
medium sized medium-speed diesel engine with an output of 5 MW per
turbocharging group. For the configuration A the turbocharger is a TPL-C, the valve
timing is chosen for optimum filling, i.e. without any Miller effect. This configuration
was abandoned with the introduction of the IMO regulations on NOx emissions (IMO
Tier 1 and Tier 2).

The configuration B goes to the limits of single-stage turbocharging, utilizing a


turbocharger A100-M with radial turbine. Configurations C and D are realized with
two-stage turbocharging. The pressure ratio of Point C is below the design range of
Power2 and is considered only for comparison. The configuration of Point D includes
the combination with a continuously variable valve timing, as realized with the ABB
product VCM® (Valve Control Management). The operation without variable valve
timing is considered unfeasible, but included in some cases as demonstration.

3.1.1 Part load behavior


A turbocharging system without control is matched and operates well at constant
engine speed. For marine applications with highly turbocharged engines the fixed
pitch propeller curve (FPP) is a challenging requirement, because the engine speed
reduction at constant torque is always accompanied by a boost pressure reduction.
Several studies have been performed in the past to evaluate the possibility to run
on FPP (10).

In (10) a requirement parameter was defined in order to quantify the degree of


speed reduction for a given operating line:

nbmep10 (1)
n10  1 
n100%
Charge air pressure [bar]
A bmep level of 10 bar was chosen, because it is
6
just above the range of operation as natural TC = 0.65
Series7
aspirated engine, i.e. some boost pressure is TC = 0.75
Series8
required. For the FPP curve the degree of speed 5
variation is increased with the full load bmep. A
diagram based on experience first published in
(12) shows that the level of full load bmep with 4
allows to run the FPP curve without control is
about 23 bar. The situation today with single-
3
stage turbocharging is not very much different.

In Figure 8 it is shown how the pressure curves 2


over mass flow rate changes with efficiency and
pressure level. They are calculated for a free
running turbocharger with constant 1
temperatures and efficiency. For every pressure 0 20 40 60 80 100
Mass flow rate [%]
level and efficiency the turbine area can be
chosen for delivering the target pressure, but Figure 8: Pressure vs. mass
the combinations of efficiency and turbine area flow rate for a free running
have different effects at part load: the higher turbocharger

195
the pressure and the efficiency, the steeper is the decrease of pressure at part load.
It is evident that operation with pressure above 6 bar and high efficiency is very
challenging.

The different effects of efficiency and turbine area on the part load behavior of a
turbocharging system have been taken into account in a part load parameter (10):

TC , 1.5   T ,100% 


K teil  C   (2)
TC  
,100%   T , 1.5 
1.25
T

For a good part load behavior the 1.5

parameter Kteil should be well above =


one, for full load bmep = 25 bar about 1.4
Kteil Necessary range
1.3 (10). An equivalent part load
efficiency can be defined multiplying 1.3
the part load efficiency by the ratio of
the turbine flow coefficients. This 1.2
parameter cannot change very much,
A
therefore the part load parameter is 1.1
almost only a function of the full load
efficiency (Figure 9). The position of B
1.0
the points for the four configurations C D
considered are all below the required 0.9
range for running FPP without control, 0.5 0.6 0.7 0.8
TC,100%
even though with two-stage
turbocharging the equivalent part load Figure 9: Part load parameter vs.
efficiency is higher. A control is turbocharger efficiency
required in all cases.

3.1.2 Engine performance on the propeller curve


The main simulation results can be seen in Figure 10. As predicted by the diagram
in Figure 9, in all cases the minimum air to fuel ratio is below the acceptable limits
for operation with HFO (C = 4.5 650
1.7-1.9). The specific fuel Turbine inlet gas temperature
4 600
consumption is consistent t TI

with the reference values in 3.5 550 [°C]

Figure 6, the differences are 3 500


maintained in the range 50 2.5 450
to 100% engine load. 2
Air/fuel equivalence ratio
400
C
1.5 350
The possibilities to improve
the situation in case A are 1
140
300
200
well known: some NOx-Emission

overboosting, a part load


130 150
NOx

optimized specification, pulse


[%]
be
120 100
[%]
turbocharging or the 110 50
introduction of control Specific fuel consumption

options like exhaust gas 100 0

waste gate and air bypass. 90 -50


In any case a compromise is 80 -100
needed for the full load point 0 20 40 60 80 100 120
leading to a higher specific Engine load [%]

– C 4.3 1-st. – C 5.5 1-st.


fuel consumption. Case B A
beAAAA B
beBBBB
C – C 5.5 2-st.
beCCCC D – C 7.0 2-st.
beDDDD
gives the best part load
results without control, but Figure 10: Propeller curves without control

196
the situation needs to be improved with the same measures as in case A. More care
is required, because due to the high-pressure ratio the compressor map width is
reduced and the negative effect of a reduced turbocharging efficiency at full load is
larger. Since some Miller is present, valve timing variability can improve the
situation, but may not be sufficient. Even in case C the valve timing variability is
not sufficient to reach a minimum value of C = 1.9, other control devices in the
two-stage turbocharging system are not considered an attractive solution.

Completely different is the 4.5


Turbine inlet gas temperature
650

situation in case D with the 4 600


tTI
adoption of variable valve 3.5 550 [°C]
timing. The available charge 3 500
air pressure is considerably 2.5 450
higher than in the other Air/fuel equivalence ratio

cases and changing the


2 400
C

cylinder filling makes it 1.5 350

possible to control the air- 1 300


140 200
fuel ratio in a wide range. NOx-Emission
The simulation results are 130 150
NOx
excellent over the whole be
120 100
[%]

operating range (Figure 11). [%]


110 50
Another big advantage of this Specific fuel consumption

solution is that the 100 0

turbocharging system itself 90 -50


does not need any control.
80 -100
Variable devices on the gas 0 20 40 60 80 100 120
side would represent Engine load [%]

additional costs, complexity A – C 4.3 1-st.


beAAAA B – C 5.5 1-st.
beBBBB
C – C 5.5 2-st. D – C 7.0 2-st. VVT
and reliability issues with beCCCC beDDDD

HFO operation. All this can be Figure 11: Propeller curves –


avoided for case D. Case D with variable valve timing

3.1.3 The Miller effect at very low load


Further considerations lead to the conclusion that variable valve timing for case D is
very effective and necessary. Extreme Miller timing means a reduction of the filling
efficiency down to values of about 50%. This can be well compensated at high
engine load by the turbocharging system, which can be matched for the necessary
increase of the charge air pressure. But in the low load range, from idling up to
about 30% load this is not possible,
because the available exhaust energy is p
7
Case A prec
too low for an effective turbocharging. [bar]
Case A pac
This issue is illustrated in Figure 12, 6
Case D prec
showing the charge air pressure prec and Case D pac
5
the cylinder pressure at compression
start pac for the cases A and D on a
4
constant engine speed line. In the case A,
which is optimized for filling, the two 3
pressures are equal. In the case D with
strong Miller, the charge air pressure is 2
considerably increased, but the cylinder
pressure is even lower than in case A, 1
because the lower temperature allows to
reduce the pressure in the design point 0
for the same mass. In the load range 0 20 40 60 80 100
Engine load [%]
between 0 and 30% the cylinder
pressure is below 1 bar, which gives a Figure 12: Effect of reduced
comparable situation as a conventional cylinder filling

197
engine running at 5000 m altitude. In this condition it may be impossible to start
the engine, if the compression ratio is not high enough. Achieving a comparable air
to fuel ratio at idling as on a conventional engine would require a pressure ratio
about 2, which is probably only feasible with mechanical supercharging. It makes
much more sense to change the valve timing.

3.1.4 Operating range with variable valve timing


High-pressure turbocharging with variable valve timing offers even the possibility to
enlarge the operating envelope beyond the propeller curve. The lines in Figure 13
represent the possible speed reduction at bmep 10 bar in the different cases. In the
cases with Miller timing the 2.4
inlet valve timing is controlled C n2 2∙n3 n3
Case A
2.2
for keeping the air-fuel Case B
equivalence ratio C at a 2
Case C
constant value of 1.9 until the 1.8
Case D
condition of maximum filling 1.6
is reached. In the diagram
1.4
the required speed reduction
starting from bmep = 25 bar 1.2
at full load is shown for the 1
cubic propeller curve (n3), the 0 20 40 60 80 100
Engine speed [%]
double cubic propeller curve
(2∙n3) and the quadratic Figure 13: Air-fuel equivalence ratio
propeller curve (n2). vs. engine speed – bmep = 10 bar

In case A the possible speed reduction at C ≥ 1.9 is very limited. In the cases B
and C the cubic propeller curve is approached but not yet reached. In case D the
propeller curve can be run with a margin for 50% overload and even the double
propeller curve could be run, if a value of C = 1.6 can be accepted.

With the concept of case D, two-stage turbocharging with variable valve timing, the
resulting rule is: The higher the nominal charge air pressure, the larger the speed
range that can be covered.

3.1.5 Transient operation


The transient operation of a diesel engine is extremely complex, because it depends,
besides on the thermodynamic characteristics of engine and turbocharging system,
on many other parameters like load profile, engine governor settings and other
controls, mechanical inertia of engine, turbocharger and driven system, inertia of
the gas in the turbocharging system, thermal capacity of all system elements.
Looking at the turbocharger only, its contribution to the load acceptance depends
on two aspect: The part load performance, determining the available steady state
air excess and consequently the possible sudden load increase and the time
constant of the turbocharger, which controls the time required by the turbocharger
to reach a new steady state condition. But, as will be shown in the following, the
dominating influence for a fast load response is, how fast and how much the
exhaust gas expansion energy in the turbine can be changed.

The time constant of a turbocharger is defined as twice the rotational kinetic energy
of the rotor divided by the shaft power. The latter can be expressed with
compressor characteristics, leading to the formula:

4
= ∙ ∙
(3)

198
The time constant is a linear function of the inertia coefficient KJ, defined as the
moment of inertia divided by the fifth power of compressor diameter and of the
compressor diameter. On the denominator side of the fractions are the density at
compressor inlet, the specific work and the specific flow of the compressor stage.
The time constant depends on the operating point, therefore for comparison it can
be defined for a constant pressure ratio or a constant engine load. In Table 1 the
time constants are given for the four cases at 70% engine load.

Table 1. Time constants of the considered configurations

Time constant Case A Case B Case C Case D


[s]
 HP stage 1.82 1.12
 LP stage 1.64 1.48
 TC system 1.57 2.49 1.70 1.34

Transient loading of large


Engine 120
engines can follow different output
D+VVT
patterns, which can change the [%]
100
influence of different system D
A C
parameters. In the attempt to 80
B
find a representative case,
simulations have been performed 60
for a load increase from idling to
100% at constant speed. On 40 Case A Power
PIV 4.0
power plant applications this is Case B 1-st Power
PIV 5.5
made usually with a linear ramp, 20 Case C 2-st Power
PIV 5.5
whose gradient is dictated by a Case DPower
PIV 7.0
Case DVVT
PIV 7.0 + VVT
Power
specific critical load range. In the 0
simulations an ideal controller -5 0 5 10 15 20 25 30
Time [s]
was used, which is able to keep
the air fuel ratio on a constant Figure 14: Load ramp 0 – 100 % with
value. The results can be seen in constant C = 1.5
Figure 14.

Under the defined boundary conditions case A reaches 100% power in less than
10 s. Case B requires about 22 s, because the Miller effect reduces the power
margin at the beginning of the ramp and the turbocharger has a larger time
constant. With comparable Miller effect and two-stage turbocharging (Case C) the
load acceptance time is reduced by about 20%. In case D with constant valve
timing the initial power increase is further lowered, but this is overcompensated by
the superior transient performance of the two-stage system with lower inertia. In
all these cases with fixed valve timing the time to full torque is considerably higher
than in the reference case, but applying variable valve timing this time is reduced
to less than 6s. It must be remarked here that case D with constant valve timing
represents the behavior of the system under the assumption that the combustion
works well. Since the compression temperature is very low, it is possible that the
long ignition delay considerably reduces the combustion efficiency. This problem
disappears by applying variable valve timing.

The faster acceleration of two-stage systems can only partially be explained with
the lower time constant. In addition to this the HP turbine accelerates faster,
because it can use initially a very large part of the exhaust enthalpy. When the gas
pressure in the exhaust manifold is increased, a sudden increase of the expansion
ratio for the HP turbine results, whereby the LP turbine receives with some delay,
depending on the volume between the stages, the remaining enthalpy after
expansion in the HP turbine.

199
3.1.6 Effect of the pressure repartition
The repartition of the pressure ratio between the stages which gives the best
efficiency in the design point depends on the ratio between intercooler and ambient
temperature. With typical values of these temperatures, optimum values between
1.2 and 1.5 results for the ratio C,LP/C,HP. It has already be shown (3)(5) that a
value of 2 gives an efficiency, which is only marginally lower in the design point,
but leads to a smaller LP turbocharger. The thermodynamic effects of the change
have been studied, comparing
the results of case D (C,LP/C,HP = 2.5 0.8
Turbocharging efficiency
2) with a case D’ with C,LP/C,HP =
T
1.3.
2 0.6
For the case D’ the dimensions of
the stages are changed: The Series1
CLP/CHP = 2
linear dimensions of the HP Series2
CLP/CHP = 1.3

turbocharger can be reduced by 1.5 0.4


3%, those of the LP turbocharger Range for optimum efficiency
must be increased by 9%. The
turbocharging efficiency in the
1 0.2
design point is improved by about
2 points. But since the pressure Pressure ratio split
ratio of the LP stage decreases
much faster with engine load 0.5 0
than that of the HP stage, case D 0 20 40 60 80 100
Engine load [%]
has a better efficiency in the
middle load range (Figure 15). In Figure 15: Effect of pressure ratio
the part load region case D gives split on steady state operation
an improvement of 3% for the
air-fuel ratio and 13 °C exhaust
2 8 p
gas temperature reduction CLP/CHPnTLHD
510.87 =2 [bar]
against case D’. CLP/CHPnTLHD
510.87 = 1.3 6

1.5 4
Interesting is also the comparison
Charge air pressure
of the transient operation (Figure 2

16). By reducing the LP stage 1 0


pressure ratio the HP stage Normalised HP TC speed
makes more work, it runs faster speed -2
[-]
and accelerates faster. But the 0.5 -4
consequence is that the delay of LP TC speed
the LP stage is increased and the -6

pressure ratio for full power is 0 -8


reached later. The best system -5 0 5 10 15 20 25

performance is achieved when Time [s]

both stages can contribute and Figure 16: Effect of pressure ratio
this is realized better in the split on transient operation
reference case.

3.2 Case study high-speed engine


In this case a high-speed diesel engine with a power output of about 1.3 MW per
turbocharging unit was investigated by means of simulation. The four cases A, B, C,
D are defined in the same way as for the previous case study. Typical for high
speed engines is strongly reduced overlap and a wider operating envelope to be run
without control. This implies some over-boosting in the design point with wide
compressor maps and allowance for higher exhaust gas temperatures. All
turbochargers involved are with radial turbine.

200
3.2.1 The engine performance
at constant speed 100
A
Figure 17 shows the full engine be
power performance of cases A, B, C [%] B
98
and D in a trade-off diagram of fuel
consumption and NOx emission. The
nominal operating points feature 96
constant levels of air/fuel ratio and
maximum cylinder pressure. While C
94
in case A the medium-speed engine
D
is being operated at a pressure ratio
of 4.3, the HS engine at bmep 25 92
HS Engine
bar without Miller requires a MS Engine
pressure ratio of about 4.7, thus, 90
the distance between the cases A 0 20 40 60 80 100
NOx [%]
and B is reduced in the high-speed
engine case. Figure 17: be-NOx trade-off diagram

At the nominal load point a very similar pattern of specific fuel consumption
reduction as in Figure 6 can be observed. The tendency of NOx emission reduction
with increasing Miller timing turns out to be more pronounced in this case
compared to the medium-speed example. At a total pressure ratio of c,tot = 7, NOx
emission is reduced by 40% compared to the nominal point of case A.

3.2.2 Transient operation


The time constant of the resulting turbocharging systems for the high-speed engine
cases are shown in Table 2.

Table 2. Time constants of the considered configurations at


70% engine load

Time constant Case A Case B Case C Case D


[s]
 HP stage 1.12 0.69
 LP stage 2.53 1.83
 TC system 1.37 1.193 1.98 1.37

Transient simulations have been carried out in the same manner as for the
medium-speed engine (Figure 18). However, since the use of heavy fuel oil can be
excluded for this type of engine, the controller limit was extended corresponding to
a lower limit of excess air/fuel ratio
Engine 120
of C,min = 1.3. The initial increase of output
[%]
engine power is reduced from case A 100
to case D according to the reduction
of the charging efficiency (Miller 80

effect). Therefore, Cases B and C


with roughly the same Miller effect 60
Case A
show the same level of initial power
40 Case B
step. After this initial step, the
Case C
subsequent ramp up is dominated by
20 Case D
the TC system time constants. Case
Case D + VVT
D with VVT clearly outperforms the 0
reference Case A: while the reference -5 0 5 10 15 20 25 30

case reaches full engine power within Time [s]

roughly 7 seconds, Case D VVT Figure 18: Load ramp 0 – 100 %


ramps up in less than 4 seconds. with constant C,min = 1.3

201
Figure 19 shows the transient Engine 1.2
speed
behavior of the engine operated in [-]
island mode during a load 1.0

acceptance in two steps. In this


operation mode the advantage of 0.8

two-stage turbocharging is much


0.6
more pronounced than in the
preceding case. The reason is that Case A
0.4
in the long ramp both
Case B
turbochargers must contribute to 0.2 Case C + VVT
the load step. If the load step is
Case D + VVT
smaller, the HP stage, which reacts 0.0
much faster, can provide the 0 5 10 15 20 25 30
Time [s]
necessary pressure change also
without contribution of the LP stage. Figure 19: Engine load acceptance in
island mode with constant c,min = 1.3

CONCLUSIONS
Value 1-stage
A value diagram was presented in 2-stage

(3) showing qualitatively that there


is a range of pressure ratio, where
the value of single and two-stage
turbocharging systems is not
optimal. In this paper different
aspects of the system performance
have been studied for two specific
engine type cases.  C,overall

Figure 20: Weighted value function for


The results can be summarized as single- and two-stage turbocharging
in Table 3.

Table 3. Evaluation case studies

Case A Case B Case C Case D


Compactness 0 - -- -
Engine efficiency 0 + ++ +++
Part load 0 + - +++
Transient performance 0 - - +++
NOx emissions 0 + + ++
Potential for power
0 - + ++
increase

 Case B (A100) represents the last step in the development of single-stage


turbocharging; It gives an improvement in efficiency and emissions
(IMOII) using the potential of the single-stage technology.
 Case C, replacement of single with two-stage at constant pressure ratio, is
not considered attractive, because the added value is too small.
 Case D, replacement of single with two-stage turbocharging at much
higher pressure ratio in combination with Miller and variable valve timing
gives access to a large improvement potential.

The last generation of ABB A100 turbochargers was developed to exploit the full
potential of the single-stage turbocharging technology. The requirements set by
case B in the study can be fulfilled with excellent performance.

202
ABB’s second-generation Power2 and VCM are innovative products designed for the
efficient exploitation of the two-stage turbocharging technology on large engines. In
the study the engine output was kept constant for a better comparability. In a
further extension of the present work a case E with pressure ratio 8.5 was
considered, applying either more Miller at constant power for emission reduction or
a power density increase to bmep = 30 bar at constant Miller effect. In all cases the
fuel consumption, the part load and the transient behavior were comparable or
better than for case D.

Within the design range of second-generation Power2, the high pressure ratio and
the high efficiency are important: The engine designer is free to set a bmep target
in the development of new engines without considering any limitation from the
turbocharging system. An additional asset of two-stage turbocharging systems is
the possibility to compensate altitude without any need for de-rating.

REFERENCES

(1) Raikio, T., B. Hallbäck & A. Hjort, 2010, Design and first application of a two-
stage turbocharging system for a medium-speed diesel engine, 26th CIMAC
World Congress in Bergen (N).
(2) Haidn, M., J. Klausner, J. Lang & Ch. Trapp, 2010, Zweistufige Hochdruck-
Turboaufladung für Gasmotoren mit hohem Wirkungsgrad, 15.
Aufladetechnische Konferenz, Dresden (D).
(3) Codan, E. & Mathey, Ch., 2007, Emissions – a new challenge for
turbocharging, 25th CIMAC World Congress in Vienna, Austria.
(4) Codan, E., Mathey, Ch. & Vögeli, S., 2009, Applications and Potentials of two-
stage Turbocharging, 14. Aufladetechnische Konferenz, Dresden (D).
(5) Codan, E., Mathey, Ch. & Rettig, A., 2010, two-stage Turbocharging –
Flexibility for Engine Optimization, 26th CIMAC World Congress in Bergen (N).
(6) Millo, F., Gianoglio, M. & Delneri, D., 2010, Combining dual stage
turbocharging with extreme Miller timings to achieve NOx emissions
reductions in marine diesel engines, 26th CIMAC World Congress in Bergen
(N).
(7) Ruschmeyer, K, Rickert, C. & Schlemmer-Kelling, U., 2011, Potential des
Caterpillar MaK 6 M32 C mit zweistufiger Abgasturboaufladung, 16.
Aufladetechnische Konferenz, Dresden (D).
(8) Mathey, Ch., 2010, Variable Valve Timing – A necessity for future large diesel
and gas engines, 26th CIMAC World Congress in Bergen (N).
(9) CIMAC, 2007, Turbocharging Efficiencies – Definitions and guidelines for
measurement and calculation, Recommendation Nr. 27, Conseil International
des Machines à Combustion, Frankfurt am Mein (D), (www.cimac.com).
(10) Codan, E., Müller, G., 1997, Anforderungen an Aufladesysteme für zukünftige
Grossmotoren, 6. Aufladetechnische Konferenz, Dresden (D).
(11) Meier, E., 1983, Part-load operation of very highly turbocharged four-stroke
diesel engines, 15th CIMAC World Congress, Paris.
(12) Codan, E., 1993, Optimierung des Aufladesystems und Betriebsverhaltens von
Grossmotoren durch Computersimulation, 5. Aufladetechnische Konferenz,
Augsburg (D).

203
Electrical supercharging for future diesel
powertrain applications
P Newman, N Luard, S Jarvis, S Richardson
Powertrain Research & Technology, Jaguar Land Rover Limited, UK
T Smith, R Jackson
Lotus Cars Limited, UK
C Rochette, D Lee, M Criddle
Valeo Air Management UK Limited, UK

ABSTRACT

The desire to minimise fuel consumption and corresponding vehicle CO2 emissions
for future powertrain applications drives the need for advanced charging systems.
Worldwide legislative requirements and customer desires for improved efficiency
whilst maintaining vehicle drivability and performance are challenging automotive
manufacturers to investigate innovative concepts for air charge delivery.

This paper provides an outline of an innovative charging system concept developed


to support future Diesel engine applications. Under the UK Government funded
Technology Strategy Board project ‘Provoque’ the partners have collaborated to
develop a compound charging system incorporating a standard LP VGT turbo and
48V powered electric supercharger. Validated analytical results for the hybrid
compound charging system will be presented to show the application over both
current legislative and real world replicating drive cycles. Details of the advanced
control mechanism and interaction with the electrical system including load
management will be reviewed.

Application of the innovative charging concept to a premium vehicle will be


discussed with indication of both, the drivability and performance capability, as well
as the ability to support an aggressive down-speeding concept facilitated by the
electric supercharger to further improve efficiency.

ABBREVIATIONS

BSG Belt Starter Generator


EGR Exhaust gas recirculation
JLR Jaguar Land Rover
NEDC New European Drive Cycle

1 INTRODUCTION

1.1 Global Trends and Diesel Engine Boosting System


Increased customer awareness of environmental issues in the worldwide
automotive market has driven Governments and automotive manufacturers to
focus on developing technology to support vehicle CO2 improvements as well as
maintaining the historically challenging legislated criteria emissions [1].

____________________________________
© Jaguar Land Rover, Lotus and Valeo, 2014
207
In Europe, legislative bodies have introduced CO2 fleet average targets that
challenge automotive manufacturers to introduce new powertrain technology in
support of achieving much reduced fuel consumption over the New European
Driving Cycle (NEDC) drive cycle. The fleet average CO2 target for 2015 is
130 g/km CO2 (average vehicle weight), with a glide path to 2020 of a fleet
average CO2 target of 95 g/km CO2. Going forward, further challenges will be
introduce with the adaptation of more transient cycles such as those found in the
World harmonized Light-duty Test Procedures (WLTP) [2,3,4].

Additionally, the Criteria emissions legislation relating to EU6 Stage#1 for Diesel
vehicles from 2014 onwards outlines stringent limits for NOx emissions over the
NEDC drive cycle. These limits are envisaged to become more stringent with the
proposed introduction of an EU6 Stage#2 limit forecast to be around 2018 [5].
Discussions are also underway for further legislative actions relating to Real Driving
Emissions (RDE), ensuring automotive tailpipe emissions under ‘normal conditions
of use’ are regulated [3,4,5]. This presents further challenges to the automotive
manufacturer to ensure feedgas emissions and aftertreatment conversion
efficiencies are maintained over a wide speed / load, ambient temperature and
altitude range at the same time as minimising fuel consumption.

For North American markets the focus on CO2 has also grown with the introduction
of challenging Green House Gas (GHG) targets and corresponding Corporate
Average Fuel Economy (CAFÉ) legislation [2,6,7]. In similar time frames to the CO2
legislation in Europe, from 2017 to 2021, it is projected that a significant reduction
in vehicle fuel consumption will be required with improvements ranging from ~33-
41 to ~45-61 mpg dependent on vehicle foot print [6].

Given the recent focus in North America on fuel economy there is a growing
popularity for Diesel vehicles in the market adding further complications to the
automotive manufacturer in meeting stringent criteria emissions limits outlined in
LEVIII / Tier3 [6]. The near zero criteria emissions requirements [2] drive the need
for SULEV30 products in the 2020 timeframe where NMOG+NOx tailpipe emissions
must be below just 30 mg/mile over the transient FTP75 drive cycle.

Emerging markets in China, India and South America have followed the legislative
trends of both Europe and North America and hence it can be expected that these
markets will also drive automotive manufacturers to implement novel technology to
support both low criteria and CO2 emissions.

Hence, recent focus on legislation relative to criteria and CO2 emissions, together
with corresponding customer demand for improved fuel economy from automotive
manufactures have resulted in new technology and operating strategies being
researched and introduced. On Diesel powertrain applications, as well as the focus
on combustion efficiency and friction reduction, there has been key strategic
direction of powertrain downsizing applied industry wide supported by the
introduction of new charged air boosting system technology [8,9]

Examples of downsizing strategies and the introduction of Diesel powertrains with


low fuel consumption and high specific power outputs utilising advanced boosting
system concepts can be seen industry wide. Examples throughout the market from,
budget to premium, including Toyota [8], Volvo [10], VW [11] BMW [12] can be all
seen in literature.

Further steps to improve the CO2 performance of the Diesel powertrain are
currently under investigation with increasing focus on downspeeding [13,14] where
two methods of downspeeding are envisaged [9]:

208
1. Use of longer gear ratios (internal to transmission or final drive)
2. Carry over drive ratios – downspeed using gear shift strategy

Modifying drive ratios does have drawbacks and can result in compromise with
respect to other vehicle attributes such as gradability. Moving to downspeeding
utilising the gear shift strategy is preferable in maintaining vehicle capabilities,
however this does present further challenges in terms of transient performance of
the powertrain.

In order to support the downspeeding concept of carrying over drive ratios and
utilising shift strategy, charged air boost systems have been the further focus of
activities to support a powertrain wide system approach to CO2 reduction, whilst
maintaining criteria emissions performance. Examples of the development of series
stage turbocharged boost systems can be seen in literature [15,16] with focus on
maintaining top end performance, but also facilitating the low speed transient
capability that supports downspeeding. Further examples of the introduction of
compound boost systems utilising both turbochargers and superchargers can also
be seen in literature as a strategy for supporting improved transient performance
and further enabling downspeeding [9,10].

As powertrains become increasingly integrated with electrical architectures the


usage of electrical assist to support charged air boosting increases. The
development of electrically assisted turbochargers offers potential solutions [16,17]
as well as the development of electrically driven compressors to provide
instantaneous air charge delivery [18,19,20,21].

This paper describes the initial stages of work including analytical investigation
relating to the Technology Strategy Board UK Government funded, Provoque
project which looks to develop the technology of a 48V electric supercharger and
aggressive downspeeding strategy on a Diesel engine within a premium automotive
product.

1.2 The Provoque Project


To support technology transfer Jaguar Land Rover (JLR), Lotus Engineering, Valeo
and RaiCam joined as a collaborative consortium to develop 48V electrical
technologies and novel clutch by wire ‘e-clutch’ concepts to support the delivery of
a premium Sports Utility Vehicle (SUV) that would achieve minimum CO2 through
the use of aggressive downspeeding and other technological refinements.

Utilising a Jaguar Land Rover Range Rover Evoque vehicle the project aims to
develop on the latest Diesel engine architecture, a lightweight low friction
cranktrain, 48V electric supercharger, 48V Mild Hybrid Belt Starter Generator (BSG)
and e-clutch (manual clutch pedal with electronic rather than mechanical actuation
of the clutch), together with advanced active NVH technology to support delivery of
a 99 g/km compact premium sports utility vehicle. Starting in 2013 the project has
progressed with initial analytical assessment of the air charge boost system
configuration now completed.

2 ENGINE / BOOST SYSTEM DESIGN

2.1 JLR Next Generation Diesel Engine


As announced in the international press, to support JLR’s development as a stand
alone company, a new family of technologically advanced engines have been
developed to ensure future vehicle products deliver competitive attributes for the
worldwide customer base.

209
To ensure the research study and technology development described in this
publication are aligned to future JLR applications, a 2.0l four-cylinder Diesel
Research engine concept derived from the proposed new engine family was chosen
as the base powertrain for the Provoque project. High level details of the research
engine are given in Table 1. The JLR research engine features an advanced Diesel
combustion system utilising the latest high-pressure solenoid injection system,
together with high and low pressure EGR circuits.

Table 1: High Level Details of JLR Research Diesel Engine

Engine Parameter Specification


General Architecture Inline Four
Capacity ~ 2.0 l
Standard Boosting System Single VGT
Specific Power ~65 kW/l
Specific Torque ~ 200 Nm/l

As part of the Provoque project the research engine is planned to be fitted with
both a Valeo 48V electric supercharger and Belt Starter Generator device to support
attainment of the Provoque project vehicle demonstrator target of 99 g/km CO2.

Valeo 48V
Electric
Supercharge

Figure 1: Computer Aided Design image of JLR Research I4 Diesel Engine

2.2 Valeo electrical supercharger


The Valeo 48V electrical supercharger (VES) concept was first introduced in a 12V
variant and has been now further developed into the new 48V hardware for
evaluation through the Provoque project. The electrical supercharger concept
consists of the following sub-systems – centrifugal compressor, rotor assembly,
high-speed electric motor, power electronics and microcontroller.

Table 2: Valeo electrical supercharger specification

Parameter Specification
Voltage Platform 48V (nominal DC)
Motor Three phase Switch reluctance
Cooling Air Cooled
Peak Speed ~70,000 rpm
Peak Acceleration ~400,000 rpm/s
Peak Electrical Powers 7.5 kW
Compressor Type Radial

210
High level details of the 48V Valeo electrical supercharger (VES) concept are given
in Table 2 with an image of the unit presented in Figure 2. Control of the Valeo 48V
electrical supercharger is via CAN and this coupled with an e-hook to the Engine
Control Unit (ECU) for the air path control software, enables controlled demand of
the electric supercharger when boost is required.

Figure 2: Image of the Valeo electrical supercharger (VES)

3 ONE-DIMENSIONAL MODELLING RESULTS

Using a GT Power one-dimensional model of the JLR four cylinder Diesel Research
engine, a detailed study of the application of the Valeo 48V electric supercharger
was completed. A schematic of the one-dimensional model is shown in Figure 3
highlighting the final proposed layout of the innovative boost system configuration.

Figure 3: Schematic of the GT Power engine model developed to support


assessment of the electric supercharger application to a JLR Diesel engine

211
An assessment of the potential location for the Valeo 48V electric supercharger was
completed using the model to determine the optimum location for this
supplementary boost device and overall boost system layout. Figure 4 presents
results for the electric supercharger located pre (before) and post (after) standard
VGT turbocharger. The results show a small advantage in transient response with
location of the electric supercharger pre the standard turbocharger with
approximately a 60ms improvement in reaction time to maximum torque at 1500
rpm.

However, locating the electric supercharger pre the standard VGT turbocharger
does impact on compressor outlet temperatures as shown in Figure 4 with an
average increase across the speed range of approximately 50°C when compared to
the post location.

The corresponding compressor maps for both electric supercharger and standard
VGT turbocharger at each location are also shown in Figure 4. Results show that the
operation of the turbocharger is not significantly impacted by either location, with
the standard compressor map supporting both. However, locating the electric
supercharge pre Turbocharger does result in higher pressure ratios for the same
corrected mass flow rate which will impede the use of low pressure EGR and also
there is an impact on the width of compressor map for the performance required.
Additionally, the characteristic of the reducing mass flow rate, which reduces the
compressor performance for the pre-Turbocharger location can also be seen in
Figure 4.

Figure 4: Load Step transient comparison on JLR Research engine


investigating position of Valeo 48V electric supercharger

Based on the results presented, the Provoque project’s direction was to position the
Valeo 48V electric supercharger downstream of the standard VGT turbo and
develop a new electric supercharger compressor wheel to fully optimise the
operation. Figure 5 shows a comparison of compressor maps for both standard and
downspeeded operational cases using two different frame sizes.

212
Figure 5: One Dimensional analysis of Valeo 48V electric
supercharger compressor matching

Having determined the optimum position for the Valeo 48V electric supercharger
and developed an optimised compressor wheel to support the required operational
envelope window, further one-dimensional analysis was completed to understand
the benefit of the electric supercharger technology with respect to downspeeding.

Figure 6: One Dimensional simulation of transient load step comparisons


on JLR Research I4 Diesel engine with / without Valeo 48V electric
supercharger

Figure 6 shows the progression of full load torque at six different engine speeds
comparing the electric supercharger in combination with the standard VGT
turbocharger against the standard VGT turbocharger alone. The data shows that
although during the very early phase of load progression the turbocharger matches

213
the electric supercharger (approximately up to 40% full torque), significant
improvements in time to full load are achieved. As engine speed increases the
improvements in time to full load using the electric supercharger are still apparent,
although the standard VGT turbocharger does close the gap providing responses
closer to that of the new technology.

Further transient modelling was undertaken with the aim of estimating the energy
required by the electric supercharger during various vehicle drive-cycles when
utilising a ‘down-speeded’ gear shift strategy. Development of the one-dimensional
model to include representations of the JLR Diesel Research engine with the both
the Valeo electric supercharger and BSG within the chosen vehicle demonstrator
platform was completed.

Figure 7: One Dimensional vehicle simulation of JLR Research engine


operating with Valeo electric supercharger over an NEDC drive cycle

Following evaluation of the location of the electric supercharger, optimisation of the


compressor wheel to support both normal and downspeeded operation and
transient evaluations of energy consumption, vehicle simulations to estimate the
potential fuel economy offered by implementation of the technology and the ability
to downspeed were investigated. Figure 7 shows simulated engine fuel flow for the
baseline (standard VGT & vehicle/transmissions shift schedule) together with the
application of the electric supercharger and associated downspeeded operation. In
this instance, the downspeeded operation was characterised by shifting into a
higher gear during steady cruise operation. The results clearly show a benefit and
furthermore, simulations for the Provoque project demonstrator vehicle estimate a
4.5% improvement in CO2 over the NEDC drive cycle. Figure 7 also highlights the
potential benefit of torque assist from the application of the Valeo 48V BSG unit in
conjunction with downspeeding enabled by the electric supercharger.

Noting that the legislative requirement for the state of charge of the energy storage
devices is that it must remain neutral over the cycle, further work is required to
understand the optimum balance between the electrical energy for the electric
supercharger with downspeeding and the electrical energy for torque assist utilising
the BSG e-machine.

214
The results presented in this paper highlight the potential of implementing an
electric supercharger to support downspeeding thereby improving fuel
consumption. The next steps within the Provoque project will focus on developing
the control architectures required to support physical operation of the boosting
concept and validation of the fuel economy benefits through vehicle correlation
testing.

4 CONCLUSIONS

The early phases of the Provoque project investigation into the application of a 48V
electric supercharger have developed the optimum location for the concept in
conjunction with a standard single VGT turbocharger and indicated potential fuel
saving available with downspeeding, which is enabled by the transient response
improvement of an electric supercharger.

Optimisation of the electric supercharger and standard VGT turbo have been
completed using one dimensional analysis and the validation of this analysis is
planned during the next stage of the Provoque project.

5 ACKNOWLEDGEMENTS

The authors of this paper would like to thank all of the other members of the
Provoque project consortium for their involvement and the Technology Strategy
Board for their continued support, without any of whom this project would not have
been possible.

REFERENCES

0. Dr.-Ing. Detlev Schöppe, Dipl.-Ing. Stefan Lehmann, Dipl.-Ing. Nicolas


Nozeran, Dipl.-Ing. Friedrich Kapphan, “Next Generation of Common Rail
Diesel Injection System Featuring Piezo Injectors with Direct-Driven Needle
and Closed-Loop Control”, 22nd Aachen Colloquium Automobile and Engine
Technology, 2013
1. T. Johnson, Corning Incorporated, “Vehicular Emissions in Review” SAE Paper
2013-01-0538, 2013
2. Dr. Marc Uhl, Marcel Wüst, Dr. Ansgar Christ, Dr. Nikolas Pörtner, Alexander
Trofimov, “Electrified Powertrain at 48 V – More than CO2 and Comfort”, 22nd
Aachen Colloquium Automobile and Engine Technology, 2013
3. Dr. Volkmar Denner, “Shaping the Future – Innovations for Efficient Mobility”,
34. Internationales Wiener Motorensymposium, 2013
4. Dr.-Ing. Olaf Erik Herrmann, Dipl. Ing. Sebastian Visser, Dipl. Ing. Dirk Queck,
Ken Uchiyama, Katsuhiko Takeuchi, Koji Ishizuka, Dr.-Ing. Thorsten
Schnorbus, Dipl.-Ing. Joschka Schaub, “Combustion Improvement and
Emission Control Technologies Supporting the New Cycle Requirements for
Passenger Car Diesel Engines”, 22nd Aachen Colloquium Automobile and
Engine Technology, 2013
5. Dr B. Holderbaum, T. Korfer, Dr T. Schnorbus, M. Scassa, Dr D. Tomazic, H.
Nanjundaswamy, J. Schaub, “LEVIII and CAFÉ 2025 – Innovative Measures for
Compliance of Most Stringent Legilsative Demands”, 22nd Aachen Colloquium
Automobile and Engine Technology, 2013
6. D. Stanton, S. Charlton, and P. Vajapeyazula, “Diesel Engine Technologies
Enabling Powertrain Optimization to Meet U.S. Greenhouse Gas Emissions”,
SAE 2013-24-0094, 2013

215
7. J. Chisaki, K. Yoshijima, T. Kikuchi, S. Morinaka, K. Yamada, M. Okamoto, T.
Oda and K. Manabe, “Development of a New 2.0-Liter Fuel-Efficient Diesel
Engine”, SAE 2013-01-0310, 2013
8. P. Wetzel, “Downspeeding a Light Duty Diesel Passenger Car with a Combined
Supercharger and Turbocharger Boosting System to Improve Vehicle Drive
Cycle Fuel Economy”, SAE 2013-01-0932, 2013
9. Dr D. Crabb, M. Fleiss, J-E Larsson, J. Somhorst, “New Modular Engine
Platform from Volvo”, MTZ 09I2013 Volume 74, 2013
10. Dr.-H-J. Neußer, J. Kahrstedt, H. Jelden, H-J. Engler, R. Dorenkamp, Dr. S.
Jauns-Seyfried, A. Krause, “Volkswagen’s new modular TDI® generation”, 33.
Internationales Wiener Motorensymposium, 2012
11. P. Langen, “Future Mobility Solutions of the BMW Group”, 33. Internationales
Wiener Motorensymposium, 2012
12. D. Stanton, S. Charlton, and P. Vajapeyazula, “Diesel Engine Technologies
Enabling Powertrain Optimization to Meet U.S. Greenhouse Gas Emissions”,
SAE 2013-24-0094, 2013
13. Dr. T. Schmidt-Sandte, Prof. J. Hammer, “In Search of the Optimal Future
Powertrain”, MTZ 5 07-08I2012 Volume 73, 2012
14. Q. Zhang, C. Brace, S. Akehurst, R. Burke, G. Capon and L. Smith,
“Simulation Study of the Series Sequential Turbocharging for Engine
Downsizing and Fuel Efficiency”, SAE 2013-01-0935, 2013
15. J. Kang, J. Lee, H. Song and D. Lee, “Enhancing Power Density with Two-
Stage Turbocharger”, SAE 2012-01-0709, 2012
16. A. Darlington, D. Cieslar, N. Collings and K. Glover, “Assessing Boost-Assist
Options for Turbocharged Engines Using 1-D Engine Simulation and Model
Predictive Control”, SAE 2012-01-1735, 2012
17. N. Terdich and R. Martinez-Botas, “Experimental Efficiency Characterization of
an Electrically Assisted Turbocharger”, SAE 2013-24-0122, 2013
18. S. Tavernier and S. Equoy, “Design and Characterization of an E-booster
Driven by an High Speed Brushless DC Motor”, SAE 2013-01-1762, 2013
19. L. Eriksson, T. Lindell, O. Leufven and A. Thomasson, “Scalable Component-
Based Modeling for Optimizing Engines with Supercharging, E-Boost and
Turbocompound Concepts”. SAE 2012-01-0713, 2012
20. K. Nishiwaki, M. Iezawa, H. Tanaka, T. Goto and B. An, “Development of High
Speed Motor and Inverter for Electric Supercharger”, SAE 2013-01-0931,
2013
21. M. Forissier, D. Zechmair, O. Weber, M. Criddle, D. Durrieu, V. Picron, P.
Menegazzi, K. Surbled, Y. Wu, “The Electric Supercharger”, 34. Internationales
Wiener Motorensymposium, 2013

216
Electrically Assisted Turbocharger as an
enabling technology for improved fuel
economy in New European Driving Cycle
operation
T Suzuki, Y Hirai, N Ikeya
IHI Corporation, Japan

ABSTRACT

An Electrically Assisted Turbocharger (EAT) is one of the effective tools for improving
transient response and fuel consumption. This paper describes the effects of an EAT
mounted on a passenger car diesel engine (D/E) on fuel consumption during the New
European Driving Cycle (NEDC) and on engine transient response, as predicted by
means of engine simulation. The validity of the calculation model is confirmed by
comparing simulation results against results of transient and steady tests conducted
on an engine test bench.

1 INTRODUCTION

Automotive companies around the world are facing strict emission and fuel economy
regulations. In the EU, the Real Driving Emissions (RDE) regulation is proposed for
Euro 6-2, applicable from 2017[1]; vehicle performance improvement is required in
order to meet such regulations. The EU, US and Japan are further required to reduce
CO2 emissions by 20% to 45% due to increasing environmental concerns [2]. Engine
downsizing and downspeeding are important concepts within this process.

Downsizing is commonly applied in conjunction with turbocharging to improve


environmental performance while maintaining satisfactory drivability. Good
turbocharger transient performance is necessary, if vehicle transient response after
engine downspeeding is to be retained. Thus, turbochargers offering a good balance
between steady-state and transient performance play an important role in realizing
low emission transportation systems.

The Variable Geometry System (VGS) is a validated technology, mainly applied on


diesel engines (D/E), able to achieve good transient response at low engine speeds,
while offering low fuel consumption and good control of exhaust gas recirculation
(EGR) gas ratio. The VGS may, however, lead to increased back pressure and, thus,
degraded fuel economy.

IHI Corporation has developed simulation tools for turbocharger fuel economy effect
estimation during vehicle running. An EAT is known as the system which can improve
the transient response and Brake specific fuel consumption (BSFC) [3-4]. This paper

_______________________________________
© The author(s) and/or their employer(s), 2014
217
focuses on an EAT mounted on a passenger car D/E and examines the effects thereof
on fuel consumption during the New European Driving Cycle (NEDC) and on engine
transient response [4]. The EAT consumes electrical power during the assist period,
which is recovered by engine alternator generation during vehicle deceleration. The
analysis model validity is confirmed with 2.2L D/E by comparing simulative results
with experimental steady-state and transient results from an engine test bench. In
the meanwhile, 2.0L D/E is assumed as the target for EAT study.

A turbocharger properly sized for a 2.0L D/E with VGS and EAT is assumed herein and
an optimum combination of certain operating parameters, minimizing fuel
consumption and EAT power consumption is sought. The EAT potential for fuel
consumption reduction during mode operation and for transient response
improvement is then presented. The EAT control logic is designed so as to be a
minimum of required modifications for EAT adaptation by vehicles utilizing 12V
batteries; the results presented are thus within the bounds of this limitation.

2 VALIDATION OF CALCULATION TOOL

IHI Turbocharged Engine Simulation (ITES) software is a 0-D engine simulation code
used for this study [5]. The engine model validity is confirmed by comparing against
experimental steady-state and transient results. ITES requires reduced computational
time compared to commercial 1-D simulation tools and is based on the filling and
emptying method, enabling agile engineering from model tuning to execution [6-7]. It
is necessary to model both the turbocharger mechanical loss and the engine heat loss
when simulating engine bench experiments, especially at part load conditions.
Additionally, the turbocharger and engine heat capacity must be considered at
transient operation.

A comparison of simulation results from a well-tuned model and experimental data is


shown as confirmation of the calculation accuracy. EAT is not used in this validation.
The target engine and turbocharger specifications are summarized in Table 1.
Equipment which is used in this experiment is state-of-the-art and measuring
precision is sufficiently high.

Table 1 Engine specification for validation

Engine type Diesel Engine with HP-EGR


Displacement 2.2L
Cylinder Layout L4
Rating Torque Nm@rpm 340@2,000
Rating Power kW@rpm 103@4,000
T/C type REV4 (motor turned off)
Compressor wheel diameter mm 51
Turbine wheel diameter mm 44.5

2.1 Steady State calculation


Steady-state comparison is shown in Figure 1. The simulated engine torque, Specific
Fuel Consumption (SFC) and difference between boost and back pressure match well
to experimental results under full load conditions. Part load comparison at 3000rpm is
shown in Figure 2. Simulation results are sufficiently accurate at part load, despite the
presence of complicated engine systems such as variable air intake and EGR valves.

218
20Nm

5g/kWh

20kPa
20krpm

Figure 1 Comparison of experimental data and calculated data at full load

10kPa
5g/kWh

10krpm 50K

Figure 2 Comparison of experimental data and calculated data at part load

219
2.2 Transient calculation
A similar procedure is used for model validation under transient operation. For this
calculation, the engine speed and the acceleration pedal opening are used as inputs.
Those parameters are presented in Figure 3, while the comparison between
experimental and calculated data in transient mode is shown in Figure 4. The
simulated air flow rate, turbine speed and intake manifold pressure show same trends
as the experimental data. Some delay of the experimental exhaust manifold pressure
is observed, which is linked to the exhaust manifold pressure measurement system
delay.

500rpm

Figure 3 Experimental condition of transient mode

0.02kg/s 20krpm

20kPa
20kPa

Figure 4 Comparison of experimental and calculated data in transient mode

220
3 EFFECTIVENESS OF EAT

In this chapter, it is investigated the EAT impact on an assumptive 2.0L D/E, with a
focus on fuel consumption reduction and transient response improvement.

3.1 Calculation condition


The specification of the target engine and EAT is summarized in Table 2. The engine
utilizes Low Pressure loop EGR (LP-EGR). The EGR ratio is controlled by valves
installed at the recirculation route and exhaust pipe. Electrical power regeneration by
engine alternator is applied during vehicle deceleration and it should be noted that
only power generated in this manner is available for EAT operation.

Table 2 Engine specification for EAT study

Engine type D/E with LP-EGR


Displacement 2.0L
Cylinder Layout L4
Rating Torque Nm@rpm 375@1,600
Rating Power kW@rpm 110@4,000
T/C type REV4
Compressor wheel diameter mm 49
Turbine wheel diameter mm 39
Rating power of motor kW/rpm 1.5kW/70,000-128,000rpm

3.2 Effect of EAT on fuel consumption during NEDC


Vehicle speed is input as target in this study, while the acceleration pedal opening is
under feedback (F/B) control to follow vehicle speed. The driver model is tuned by
changing the PID coefficients of F/B control. The VGS control logic, the motor and the
generator operation logic of the EAT, as well as the gear set of the transmission are set
as parameters for which an optimum combination minimizing fuel consumption and
EAT power consumption is sought.

VGS nozzle control EAT motor control

Switch 1
Switch 1 Switch 2
Switch 2

Figure 5 Example of control map of the VGS nozzle

VGS nozzle and EAT motor are controlled as shown in Figure 5. The VGS nozzle and
EAT motor are under F/B control with boost pressure as target above and under open
loop control below the switch line. Two switch lines (Switch 1 and Switch 2) were
prepared. With respect to EAT motor control, three maps for F/B control are used:
Powerful operation, Proportional operation and Economy operation. The difference
between these maps is the maximum assist power of EAT motor. Powerful operation
assists in a proactive way, while economy operation assists in a power-saving manner.

There are also two modes of EAT motor open loop control: control with motor
assistance and control without assistance (motor off). The assist is always off at 5th
and 6th gear and during shift-up operation. Two patterns of transmission gear set
(T/M) are prepared: Gear Set 1 as the base gear set and Gear Set 2 for downspeeding.

221
The gear ratio of Gear Set 2 is smaller than the base one. The down-speed ratio at 1st
gear is smaller than other gear position to avoid start ability degradation, while
downspeeding is not applied at 6th gear, so as to maintain vehicle top speed. The gear
sets are summarized in Table 3.

Table 3 Downspeeding Ratios

Down-speed ratio from Gear Set 1 % Remark


Gear 1 2 3 4 5 6
Gear Set 2 8.5 10.7 8.7 4.6 4.1 0 Down-speed
version

The NEDC simulation is conducted with the 8 cases indicated by Table 4. In CASE2,
CASE3, CASE4 and CASE5, the motor assist turns on only while the vehicle is
accelerating; in CASE6, CASE7 and CASE8, the motor assist turns on not only when
accelerating but also during constant speed running, as later described. Fuel
consumption during NEDC is summarized in Figure 6.

Table 4 Calculation pattern for NEDC study

CASE VGS EAT Motor Motor Motor T/M Shift-up


switch Switch Open F/B Turn- gear timing
loop off set Set
1 (Base) SW2 - - - - 1 1
2 SW1 SW2 Off Pro X 2 1
3 SW1 SW1 Map Pro X 2 1
4 SW1 SW2 Off Eco X 2 1
5 SW1 SW2 Off Pow X 2 1
6 SW1 SW1 Map Pro - 2 1
7 SW2 SW1 Map Pro - 2 1
8 SW2 SW1 Map Pro - 2 2
SW1: Switch 1/ SW2: Switch 2
Map: Map control/ Off: motor turn off
Pow: Powerful operation for motor
Pro: Proportional operation for motor
Eco: Economical operation for motor
Shift up timing set 1: 1700rpm
Shift up timing set 2: The engine speed to keep the same vehicle speed as gear set1
Motor turn off: Motor turns off when current boost pressure is over target

2.2%

Figure 6 Fuel consumption during NEDC

222
In CASE2, CASE4 and CASE5, the motor assist turns on while the vehicle is
accelerating during mode. Among these, CASE4 (Eco operation) shows highest
electrical power consumption and maximum fuel reduction. This is due to the EAT turn
off switch, which is the function turning the EAT off when current boost pressure
reaches target boost pressure. The motor only resumes the assist when the deviation
reaches 5%. In case of abrupt acceleration, operations Pow and Pro can offer more
electrical power than Eco, which can lead to the motor turning off caused by boost
pressure reaching the target value. Despite Eco operation assisting with less power, it
continues to assist for longer because target boost is not achieved. As a result, Eco
operation consumes the most amount of electrical power, which leads to the highest
fuel reduction rate. The assist power difference between Pow and Pro operation is little
because NEDC has a high operating proportion at loads below 50%. This means that
the Pro operation would be sufficient for low load modes such as NEDC.

Therefore, CASE3 and CASE6, where the motor assist turns on not only while
accelerating but also during constant speed running, both use Pro operation. A
comparison between operation with and without turn off switch is conducted, but this
time avoiding motor turn off due to target boost pressure obtainment and, thus,
allowing the motor to operate for a longer time period. As can be seen, CASE6
exhibited lower fuel consumption, which is the reason it was qualified for comparison
against CASE7, The difference between these two is VGS nozzle controlling, as is
shown in Table 4.

CASE7 is shown to have improved fuel consumption. This is because the F/B area of
Switch 2 is broader than that of Switch 1. Thus, F/B control uses nozzle openings of
higher efficiency. With a suitable open loop control map, the difference between the
two will reduce. Also, a wider feedback area is better in light of performance
deterioration due to age. Eventually, CASE7, which exhibits best fuel economy of all
operating scenarios is qualified for further downspeeding and renamed as CASE8. As
expected, CASE8 shows the overall best fuel economy, achieving 2.2% decrease of
fuel consumption, with 24.8kJ used for the EAT, which corresponds to 0.1% of total
fuel energy. Figure 7 shows a number of parameters from CASE1 and CASE8 between
800s and 850s during NEDC operation.

The fuel consumption reduction of CASE8 is caused by a pumping-loss decrease and


engine down-speed. The degradation of EGR ratio cannot be observed herein. This
study reveals that the EAT motor can reduce the fuel consumption by motor assist not
only during vehicle acceleration, but also during constant speed running; however,
motor operation should be considered well. As mentioned previously, a 12V battery
vehicle is assumed. Fuel economy can further improve by increasing the assist power.

223
Figure 7 Detail behavior of each parameter

224
3.3 Effect of EAT on full acceleration
A comparison of vehicle acceleration performance is conducted next. 3 cases are
used: no assist (base), EAT assist and EAT assist with downspeeding (Table 5). The
initial condition is vehicle constant running at 2nd and 3rd gear. The acceleration
performance is assessed as average acceleration to target vehicle speed. The initial
speed and target speed are from 17.5km/h to 40km/h at 2nd gear and from 27.5km/h
to 60km/h at 3rd gear. The calculation results are shown in Figure 8.

Table 5 Calculation condition of full acceleration

CASE Gear set Motor assist Remark


1 1 - Base
2 1 X EAT assist
3 2 X EAT Assist + downspeeding

2nd gear 3rd gear

Figure 8 Comparison of acceleration

For CASE2, it is found that the acceleration performance is improved by 12.5% at 2nd
gear and by 8.5% at 3rd gear. In CASE3, acceleration performance is improved by
4.9% at 2nd gear but remained equal to base scenario at 3rd gear.

225
4 SUMMARY AND CONCLUSIONS

This study attempts to predict the effects of an EAT. Model validation is initially
conducted by comparing simulative results with experimental data from an engine
test bench, both in the steady-state and transient regime. An examination of EAT
impact on NEDC fuel consumption and on transient response follows:
- An EAT can reduce fuel consumption by 2.2% during NEDC, assuming optimal
EAT motor and VGS operation (CASE8 in table4).
- An EAT can improve vehicle acceleration by 12.5% at 2nd and 8.5% at 3rd gear
without downspeeding (CASE2 in table 5) and by 4.9% at 2nd gear with
downspeeding (CASE3 in table 5), while maintaining the base acceleration at 3rd
gear.

These conditions below are necessary to achieve the performance improvement


shown above:
- Vehicle with the engine and the transmission designed to be downspeeded.
- The combination of VGS nozzle control by feedback in a wide engine operating
condition with mild electrically assist in low engine load condition.
- Full power electrically assist when vehicles accelerate with engine full load.

To conclude, the EAT can simultaneously improve fuel economy and transient
performance, provided that the EAT is applied at a downspeeded vehicle. A vehicle
with 12V battery is assumed herein and EAT operates under this limitation. The result
will, however, change depending on the battery system of the vehicle.

5 REFERENCES

[1] Shigeo Furuno. (2013) Global Trends of Emission Regulation and Powertrain
Technologies for Vehicles, Journal of Society of Automotive Engineering of Japan.
[2] IEA (2012a), Technology Roadmap: Fuel Economy of Road Vehicles, IEA,
http://www.iea.org/
[3] Nicola Terdich, Ricardo F Martinez-Botas, Alessandro Romagnoli, Apostolos
Pesiridis (2013) Mild Hybridization via Electrification of the Air System:
Electrically Assisted and Variable Geometry Turbocharging Impact on an
Off-road Diesel Engine. ASME Turbo EXPO 2013.
[4] IHI Corporation (2011) Electrically Assisted Turbocharger the power electronics
which improve the good fuel economy. IHI technical Review, vol.51, no.1,
pp.14-15.
[5] N Ikeya, H Yamaguchi, K Mitsubori and N Kondoh. (1992) Development of
advanced model of turbocharger for automotive engines. SAE paper 920047.
[6] Watson N and Janota M.S (1982) Turbocharging the Internal Combustion Engine.
The MacMillan Press, Ltd.
[7] Georgios Iosifidis, Jason Walkingshaw, Bernhard Dreher, Dietmar Filsinger,
Nobuyuki Ikeya and Jan Ehrhard. (2013) Tailor-made mixed flow turbocharger
turbines for best steady state and transient engine performance. 1st
International Conference on Engine Processes.

226
Development of a high-efficiency
commercial-diesel turbocharger
suited to post Euro VI emissions
and fuel economy legislation
J Watson, R Lotz, D Grabowska, J Moscetti, T House, S Scott, R Vemula
BorgWarner Turbo Systems, USA

ABSTRACT

The development period post Euro VI / EPA ’10 finds diesel engine OEM’s seeking
combined combustion and aftertreatment solutions to the challenges posed by new
fuel economy-focused legislation regulating diesels. It is the intent of this paper to
discuss efforts underway, and results achieved to date, aimed at development of a
high-efficiency commercial vehicle turbocharger supporting these fuel economy
aims. This new high efficiency turbocharger is projected to achieve in excess of
60% overall efficiency and thus should support well any engine- or vehicle-level
effort targeting attainment of the new aggressive fuel economy standards.

NOMENCLATURE

BTE – brake thermal efficiency for h◦ – change in enthalpy across the
the engine stage
SCR – selective catalytic reduction Q – volume flow through the stage
EGR – exhaust gas recirculation  – rotational velocity of the stage
NOx – oxides of Nitrogen including  – efficiency of the stage, reported
NO and NO2 here on a t-s basis
PM – particulate matter P – pressure
Ns – specific speed of the stage M – mass flow
Ds – specific diameter of the stage T – temperature
D – physical diameter of the stage CSLA – constant speed load acceptance

1 INTRODUCTION / MOTIVATING FACTORS

1.1 Regulatory environment


External regulatory factors are largely responsible for driving a careful assessment
of approaches to commercial vehicle engine programs targeted for production in the
2017 and forward time frame. Pending governmental regulation in the US and
European Union mandates aggressive reductions in fuel consumption, and
consequently in CO2 emissions, with allowable emission of the traditional diesel
pollutants oxides of nitrogen (NOx) and fine particulates (PM) effectively holding flat,
at least for now. While the collective body of national and state regulators may
consider further reductions in allowable NOx near term (additional reductions in
NOx of up to 75% are possible beyond EPA ’10), the macroeconomic motivators of
(national) desire for energy independence and (owner operator) desire to minimize
fuel costs coupled with EPA’s newfound emphasis on fuel economy are expected to
cement fuel economy at top of the diesel agenda near term.

_______________________________________
© The author(s) and/or their employer(s), 2014
227
1.2 OE response – directions in diesel engine development
The fuel economy benefits of engine downsizing, the act of reducing engine
displacement to effect a reduction in friction (enabled by use of a turbocharger or
supercharger to augment engine brake power), and engine downspeeding, the act
of using transmission and related vehicle level changes to narrow the effective
speed range of the engine for purpose of reducing friction and enhancing
combustion efficiencies, have been explored for some time now and their
implementation is expected to accelerate with the pending legislation (1). Similarly,
use of Miller Cycle to modify the effective in-cylinder charge temperature with
positive consequences for NOx production (while enabling fuel economy friendly
injection timing changes) is also viewed as a viable technology for post Euro VI /
EPA ’10 operation.

Accompanying the potential introduction of the above strategies is a general trend


towards universal acceptance of SCR as the industry favored technology for
managing tailpipe NOx emissions against Euro VI and EPA ’10 mandates. The
relative success of SCR as a technologically effective and robust tool for reducing
NOx emissions is viewed as an enabler for the next round of fuel economy
enhancements, if for no other reason than it tends to better support more fuel-
economy friendly positive engine scavenging. Any new air handling system design
targeting this type of application must consider these facts in order to provide best
fit to the engine.

1.3 Air system architecture selection


It is well established that the targeted performance and emissions outputs of the
diesel engine (as prescribed by requirements of torque, air to fuel ratio, fuel
consumption, and rates of EGR) drive specification of the air system employed.
Elimination of or reduction in the usage of high pressure loop EGR, as enabled by
the relative success of SCR, tends to have a simplifying effect on the air system.
Namely it:

 Moderates the amount of boost air required, suggesting single-stage


turbochargers where two-stage units might have been required prior,
 Relieves the turbine stage from the need to drive high pressure loop EGR,
allowing for a reduction in the associated engine pumping work,
 Leads to selection of turbine stages with higher swallowing capacity (and
also likely better efficiency),
 If coupled with engine downspeeding as a strategy for friction reduction,
drives a reduction in required compressor and turbine operating ranges.
Technologies such as vane diffuser for compressor and fixed nozzle ring for
turbine thus suggest themselves for inclusion in the turbocharger as they
offer the promise of enhanced operating efficiency (though usually over a
limited range of operation),
 Reduces the necessity for variable geometry turbocharging for many
applications. Elimination of the VTG mechanism simplifies the
turbocharger, offering in the trade a more robust hardware solution and
one that should ultimately facilitate attainment of best overall turbocharger
efficiency.

The turbocharger suggested by the above is a single stage unit incorporating stages
capable of moderate to high speeds and pressure ratios, running fixed geometry for
turbine, and delivering best possible efficiency in the steady state.

228
1.4 Other considerations
Beyond top level architecture, the following must be considered in defining the
preferred turbocharger for the mission:

 Turbocharger mechanical elements (as they support the overall


performance proposition of the turbocharger), i.e. bearing and thrust
components – are ball bearings preferred over more traditional fluid film
journal bearings? How are thrust loads best managed such that adequate
thrust capacity is provided, but not to excess so as to adversely impact
performance?
 Match – i.e. selection not only of the preferred blading for each of the
compressor and turbine stages, but also selection of the proper sizes of
each in relation to both the anticipated operating ranges and in relation to
one another.
 The importance of fuel economy over a full drive cycle, driving increased
emphasis on delivery of improved stage efficiencies at part load operating
conditions.

It will be suggested herein that a thoughtful turbocharger match considering each of


the above points, informed not only by consideration of the performance and
emissions objectives of the engine but also by a thorough knowledge of the
performance sensitivities of the individual stages, will be essential to attainment of
best overall air systems performance.

1.5 The development approach: what have we done?


This paper describes a proactive turbocharger development program meant to
address the aspiration needs of a diesel engine for commercial vehicle targeting
compliance with post Euro VI / EPA ’10 emissions and fuel economy regulations. As
such the primary focus of the development is maximization of the turbocharger
overall efficiency while concurrently delivering on the requisite flow, speed,
durability, and package requirements normally associated with commercial vehicle
applications. The uniqueness embodied by the development lies in the holistic,
system level approach taken to ensure best overall performance.

System level optimization requires optimization of individual components as well as


knowledge of how those components function together as a system. The
turbocharger development effort described herein, and targeting the mission
described above in sections 1.1-1.3, has considered optimization on multiple levels
– at the impeller, at the stage, and for the turbocharger as a whole.

The first step has been to identify a baseline ‘off the shelf’ turbocharger best aligned
with the mission targets, and in the present case this means traditional high specific
speed turbocharger stages flowing the correct ranges for the engine application
targeted. The baseline turbocharger is constituted by a single stage employing a
fixed geometry wastegated turbine housing. Performance maps for the baseline
stages have been obtained and are used to reference suitability of both flow
capacity and pressure ratio for the developed stages. Efficiency targets for the
developed stages have been set in relation to the peak values shown on the
baseline maps, but consideration has also been given to the level of efficiency
increase representing a true step change when compared to today’s aero product
line.

We have here attempted to employ multiple approaches to optimization of


performance: the compressor performance has been optimized using both empirical
and state of the art numerical optimization approaches; the turbine stage has been
optimized using an empirical, experienced-based approach. The performance of the
turbocharger thrust components has been considered via CAE and geometrical

229
changes subsequently have been made to drive best combined turbine efficiency.
Ball bearings have been considered based on documented and clear evidence of
their capacity to improve performance of the overall turbocharger at low speeds.
Furthermore, the stages have been carefully size-matched based on directed test
outcomes suggesting the preferred compressor loading for the turbine developed.
In this way it is believed the best overall system level performance has been
realized.

Lastly, it should be noted that the impeller design direction to be followed here
involves a general increase in rotor group inertia as described below. These
impacts have been considered in relative terms by using engine cycle simulation
tools to assess transient engine response over defined events, both with the
baseline turbocharger and with the developed turbocharger.

It will be the intent of this paper to describe approaches applied and outcomes
observed at each level of development for this turbocharger.

1.6 Design direction for stages


It is common industry practice to employ non-dimensional parameters (e.g.
Reynolds number or Mach number) for classification purposes. Selecting the
appropriate non-dimensional parameters allows relation of candidate designs to
historical efficiency trends (5). In particular, a Balje diagram can be used to relate
single stage turbomachines categorized by specific speed, specific diameter, inflow
angle and efficiency, and as such is a useful tool for comparing potential designs,
both relative to one another and to known regions of high operating efficiency (2).
The Balje diagram presents the following non-dimensional parameters:

 specific speed (2), defined according to

 Q
Ns 
h 

3
4

 specific diameter (2), defined according to

Ds 
 
D h 
1
4

 inflow angle – i.e. as one of radial, axial, or mixed.

By making appropriate reference to the Balje diagram early in the design process,
trades of efficiency v. package size can be inferred at the outset and appropriate
decisions taken regarding preferred bounds of the design space. A Balje diagram
for turbines and expanders is depicted in Figure 1. Our applications will constrain
consideration of specific speeds to the range 0.1 to 1 in the units shown. For fixed
efficiency the trend in size with decreasing specific speed should be apparent.

In order to improve upon the efficiency of a baseline design, a change to one of


specific speed or specific diameter or both is required. With volume flow and
enthalpy change across the stage fixed as constraints, the equations defining
specific speed and diameter may be used to compute new speeds and diameters for
the candidate designs. The design geometry providing the best balance of
efficiency and size (package and rotating inertia) is usually selected for additional
detailed work.

230
Figure 1 Ns-Ds diagram for turbines and expanders, from Balje (2).
Regions of constant efficiency are plotted as contours.

Turbochargers historically applied to commercial vehicle applications involve radial


flow, high specific speed designs. In the present study radial flow, low specific
speed designs have been applied in an attempt to maximize operating efficiencies,
and it is understood that any performance uptick so realized will come at cost to
rotating inertia of the stages, and to packaging.

2 COMPRESSOR STAGE DESIGN

Radial compressors are used almost exclusively in turbochargers due to their


compact design, wide operating range and low manufacturing cost (4). A good
design combines efficiency at the target operating point (typically engine peak
power) with acceptable surge margin at low rpm engine conditions (peak torque)
and altitude margin (ability to deliver acceptable performance at higher pressure
ratios and rotation speeds). The design process is further constrained by needs to
maintain structural integrity over a long operating life, to present low inertia for
good transient response, and to be manufacturable at competitive cost.

For this development, the preliminary design of the compressor impeller was based
on empirical correlation (a Balje diagram for compressor) and simple 1-D analysis
tools (3,4). This set the overall dimensions and speeds, as well as preliminary
values for inducer and exducer blade angles and area ratios. Further refinement is
then accomplished with streamline curvature methods that allow determination of
blade loading, and more expensive 3-D CFD methods. All of these methods suffer
to some degree from not being able to represent reality faithfully, and the designs
must be tested to ensure that prediction aligns with tested outcomes.

A separate approach using numerical optimization techniques in combination with


3-D CFD was also employed. The numerical optimization employed a hybrid genetic
(7) and sequential quadratic programming (8) algorithm to vary impeller hub and
shroud contours as well blade angles to build a Pareto front of designs meeting a

231
multi operating condition and multi objective design problem. Three phases of
stand test were conducted where candidates of each approach were evaluated.

2.1 Structural optimization methodology


A conscious decision was made to define the aerodynamic shape of the impeller first,
and only subsequently to optimize the structure for both high and low cycle fatigue
loadings. This allowed the aerodynamicist more freedom in exploring different
aerodynamic concepts, but required somewhat greater effort to structurally qualify
the resulting designs. For this development impellers had to meet basic structural
integrity requirements in order to be stand tested.

Due to the relatively small inducer required by a low specific speed design and the
resulting shorter unsupported blade length, high cycle fatigue did not represent a
particular challenge for the compressor and could be accommodated with minimal
impact to aerodynamic performance. Structural integrity for low cycle fatigue
required the impeller to have peak stresses due to centrifugal loading below a
specific threshold at operating speed. An additional requirement for the impeller to
be flank millable for cost effective manufacture required a constant radius fillet at
the base of the blades.

Once the aerodynamic geometry was frozen, fillet radii and backwall shape were
seen to be the dominant variables influencing stress levels. Larger fillets and a
thicker backwall typically reduce stress in these components but also increase the
bore stress. The optimum configuration presents a balance of the lowest possible
stress in all regions. With four to six variables to describe the backwall shape and
fillets and three stress objectives for back wall, fillet and bore stress, the
optimization problem becomes too complicated to be solved efficiently using a DOE
approach. Rather, a numerical optimization scheme using a combination of a multi-
objective genetic and sequential quadratic programming algorithms was used to
build a three dimensional Pareto front characterizing the trade-off between the
three objectives. Typically 150 to 200 design evaluations were required to build up
a Pareto front of sufficient density to pick an optimum, the result being designs with
up to 10% lower stress than those based on standard dimensioning guidelines.

2.2 Results achieved – compressor performance


For this development both vaned and vaneless diffuser designs were executed and,
ultimately, a vaneless design was selected for inclusion in the high efficiency
turbocharger due to the attractive combination of stage efficiency, map width, and
pressure ratio. The ‘best’ such result achieved is as depicted in Figure 2.

It is believed that the map topology delivered supports well the aim of improved
fuel economy over targeted commercial vehicle drive cycles by improving upon
stage efficiency in regions of the map often important for part load operation, i.e. at
lower speeds and at higher mass flows.

It is also clear that peak stage efficiency is located away from full-load rated-power
operation for applications requiring more aggressive intake manifold pressures, and
so additional work will be considered to improve the map in this region.

232
Optimized Design
Baseline Design

Figure 2 Tested performance of a vaneless compressor stage developed for


this program. Efficiencies are normalized to those of the ‘off the shelf’
reference stage defining the flow capacity range. Data is measurement
on the BorgWarner test stands in Arden, NC USA. Compressor data is
corrected to 298K and 1000 mbar.

3 TURBINE STAGE DESIGN

Historically, turbocharger turbines have been designed to a high specific speed (1.0
to 1.2), frequently operating at sonic velocities and low specific diameter (1.8 to 2.0)
(4). This design philosophy yields a turbine with high through-flow, low inertia and
small space claim at the expense of efficiency. In keeping with the small space
claim, turbine volutes are frequently wrapped or folded away from the center body
to accommodate the high flow without increasing diameter.

The turbine stage selected as baseline for this development had a specific speed of
1.09 and a specific diameter of 1.85 at expansion ratio 2.4. The flow parameter
was 0.04 kg-√K / sec-kPa. The baseline stage performance compares well with
published results (4). Further, typical commercial vehicle operation drives
temperatures of the incoming exhaust gas in excess of 750°C and speeds in excess
of 500 m/sec for the turbine and the baseline turbine considered here supports both.

As indicated prior in 1.5, a more traditional approach was used in designing the
turbine stage with preliminary sizing supported by the specific speed-specific
diameter charts published by Balje (2). Having located the baseline turbine design
on the Ns-Ds diagram, new values of specific speed and specific diameter moving
the design in the direction of increasing efficiency were selected. From there, a one
dimensional impeller design using the method described by Whitfield and Baines
was performed (5). Once the preferred 1-D design was settled, a fully 3-D design
of the turbine impeller was articulated (with a variety of blade angle progressions
created), and the designs characterized using tools of CFD.

233
Of the 11 candidate impeller designs conceived, 5 were selected for stand test and
were also qualified for stress and frequency prior to prototyping and test.
Additionally, an accompanying symmetric turbine volute was designed to reduce
stator losses. Based on test results obtained, the 3 best stage designs were used
as seeds for the next design iteration. A total of 3 design iterations were completed.

3.1 Structural optimization methodology


The turbine impeller is subject to thermo-mechanical constraints which must be
accommodated in the design process. As with the compressor, an iterative process
is employed whereby thermodynamically attractive designs are analyzed in FEA with
subsequent adjustments made to the design in order to ensure viability of the
impeller. The process is repeated until both performance and durability goals are
achieved concurrently.

3.2 Results achieved – turbine performance


As shown in Figure 3, the final developed turbine stage achieved 109% of the
baseline combined turbine efficiency at 2.2 expansion ratio. The design presents a
specific speed of 0.96 at a specific diameter of 2.17. The larger diameter impeller
combined with the symmetric volute increased stage diameter by 14% and inertia
by 59% above the baseline high specific speed design.

The addition of a fixed nozzle ring has been reported to decrement performance in
some cases (6) and increment performance in others. In the event the developed
stage design did not achieve target efficiency, it was surmised that a nozzle ring
could potentially boost stage efficiency closer to target and so was considered for
inclusion in the high efficiency turbocharger developed here.

Figure 3 Performance of the low-specific speed turbine developed for this


program. The case with no fixed nozzle ring. Efficiencies are normalized
to the peak for the baseline stage. Data is measurement on the
BorgWarner test stands in Arden, NC USA. Turbine data is corrected
to 923K and 1000 mbar.

234
Five nozzle rings were designed for this study. The first four were the result of
computer optimizations of increasing complexity. The least complex design modified
only inlet and discharge angles and radii. The most complex allowed optimization of
the vane angles, radii, thickness profile and number of vanes. The final design
selected was a more traditional, experience based design.

CFD predicted a peak efficiency of 110% of the baseline combined turbine efficiency
for the moderately complex nozzle ring with a flow coefficient slightly below target.
Test results suggested little difference in performance between the stage with
moderately complex nozzle ring and the stage with the experience-based nozzle
ring design. Ultimately the lack of significant performance uptick from the case
with no nozzle ring and concern over increased high cycle fatigue risk led to the
omission of the fixed nozzle ring from the high efficiency turbocharger configuration.

4 TURBOCHARGER LEVEL CONSIDERATIONS

4.1 Optimization of turbocharger thrust components


An effort was made to quantify the potential turbocharger level efficiency gains
achievable by modifying the stock thrust component geometries. As the thrust
components themselves tend to impact mechanical efficiency of the turbocharger to
great degree, and because the geometrical parameters are easily modifiable in a
controlled laboratory setting, a decision was taken to evaluate these potentials
before moving on to consideration of more obvious alternative technology bearing
sets. The data is of interest in that it supports improved understanding of the
impact of thrust capacity on overall turbocharger performance and further supports,
for select applications, intelligent trading of excess thrust capacity for performance
gain.

Understanding that mechanical loss dwindles and bearing system loading increases
with operating speed, optimization is required to properly balance the competing
requirements of maximizing load capacity while minimizing friction power loss. In
the present study specification of thermodynamic and lubricant conditions
supported a fluid-film bearing system simulation used to predict the mechanical
friction. This prediction was then compared to measured performance data from
test stand. A sensitivity analysis involving the geometrical parameters of the
bearing system was then performed in order to determine efficiency improvements
attainable, balanced by an associated reduction in load carrying capability.

For this turbocharger tapered-land thrust bearings are considered, with the running
surface and associating annular bearing volume described by a designated height,
arc length, and radius as shown in Figure 4.

Arc Length Arc Length


Radius

Height

Figure 4 Tapered-land thrust bearing geometry optimized as part of the


present study.

Experimental results indicated a gain of up to 3% in measured combined turbine


efficiency. While these results were deemed encouraging, it must be understood
they are achievable via reduction in thrust capacity and so will be difficult to realize
for demanding engine applications.

235
Understanding that the typical operations profile for this turbocharger suggests
higher speed operation, and having satisfactorily quantified the potentials
achievable via modified traditional thrust component geometry, full attention turned
to consideration of ball bearings.

4.2 Ball bearings for improved mechanical efficiency


Ball bearings have been previously demonstrated by multiple investigators to
provide mechanical (and consequently combined turbine) efficiency benefits for
turbochargers. Typical results obtained by BorgWarner are as shown in Figure 5.

Greater benefit is seen at low speeds because the power consumed by the bearings
is a greater percentage of the total power of the turbocharger at lower speeds. For
each constant speed line, it is also surmised that there is greater benefit at low
expansion ratios because the thrust load is generally highest there. This results in
higher power consumption by the hydrodynamic thrust bearing whereas the ball
bearing cartridge power consumption is not as sensitive to thrust load.

Measurement with Ball


Bearing

Measurement with
Fluid-film Bearing

Figure 5 Impact of ball bearing on combined turbine efficiency. Data


is measurement on the BorgWarner test stands in Arden, NC USA.
Efficiencies are normalized to the peak combined efficiency attainable
by running the same turbine aero with journal bearings.

This reduction in bearing power consumption is expected to produce better transient


response, helping to overcome the increased inertia of the new impeller designs.
Given the obvious benefit to the present development ball bearings have been
slated for introduction to the high efficiency turbocharger.

236
4.3 Matching – assembling the pieces
For each stage developed, sensitivity studies were conducted via stand test to
quantify the performance variability of each with trim (ratio of inducer diameter to
exducer diameter for compressor, inverse for turbine), aerodynamic tip width of the
impeller blades, housing sizing and, for turbine, sensitivity to compressor loading.

For example, it is generally understood that combined turbine efficiency is a strong


function of compressor loading via the dependence of turbine efficiency on blade
speed ratio, and so the turbine stage developed here was tested with a range of
compressor blades and sizes and the impacts to turbine combined efficiency noted.

It is also understood that the size matching employed impacts thrust loading of the
turbocharger, which again impacts turbine combined efficiency, and so these same
tests help dimension those effects as well. Quantitative understanding of both
trends specific to the components in question supports optimization of overall
turbocharger performance.

4.4 Assessing the stage efficiency / inertia trades


In order to assess the engine response impacts of the increase in developed
impeller inertias, engine cycle simulations have been conducted. Performed early in
the development cycle, such simulations help steer the impeller design process by
suggesting, for a given engine application, how much steady map efficiency is
required to offset a particular increase in rotor group inertia. Later in the
development cycle actual achieved map efficiencies can be paired with known, as-
designed rotor inertias and engine response again checked for suitability to
application.

For each hardware configuration considered, both constant speed load acceptance
(CSLA) and acceleration events were simulated for engine models calibrated to
recent vintage (i.e. 2010 EPA or equivalent) emissions regulation. Time to target
engine torque, time to target engine speed, and time to target intake manifold
pressure are noted in each case.

Effectively, the engine cycle simulations allow the development team to quantify
differences in anticipated engine response whereby the off-the-shelf stages are
contrasted with the developed stages. Any deficiency in engine response then
motivates consideration of inertia reduction options (design or materials) for the
developed stages.

The following are noted:

 All else being equal, rotor groups of lower inertia will respond faster, but
the actual difference in engine response times provided by the lower inertia
(and lower efficiency) off-the-shelf stages and by the higher inertia (but
higher efficiency) developed stages is predicted to be at most on the order
of a few seconds for the engines and events simulated. For some
response-sensitive applications this could represent unacceptable delay,
but for a good many commercial vehicle applications a delay of this
magnitude will not be significant. In other words, the developed stages
may be best suited for genset, marine or similar applications where
response times are not paramount, but the stages will likely also have
utility for select over-the-road truck or row-crop tractor applications as well.

 Engine response is shown in these studies to be a strong function of


compressor map topology, not just peak compressor map efficiency (and
also not just the stage inertias). For this reason compressor stages
delivering a reasonable rotating inertia but wider performance maps better

237
encompassing the range of operating conditions simulated, for example a
vaneless compressor configuration with aluminum impeller, will tend to
fare better in the transient response simulations than will vaned, narrow
map counterparts.

Ultimately, results of these simulations are here meant to assess not just suitability
of the developed thermodynamics, but also to motivate consideration of the proper
material solutions for each impeller developed, and the decision has been taken
here to proceed with further consideration of material alternatives to Inconel 713C
for the high efficiency turbine impeller where intended for response-sensitive
applications.

5 RESULTS FOR FULL TURBOCHARGER

In consideration of all of the above, this program set as a developmental goal


attainment of overall turbocharger efficiency in excess of 60% as measured on test
stand. The final value achieved for the high efficiency turbocharger employing low
specific speed aerodynamics and running traditional fluid film bearings is 62.0%
overall efficiency as supported by the test data presented in Figure 6.

Projection with Ball


Bearing

Figure 6 Overall efficiency for a high efficiency turbocharger utilizing low-


specific speed stages. Data is measurement on the BorgWarner test stands
in Arden, NC USA. Compressor data is corrected to 298K and 1000 mbar;
turbine data is corrected to 923K and 1000 mbar. Ball bearing use is
expected to enhance the tested data as shown.

It is of note that this result does not (yet) include the contribution of ball bearings
to overall performance, and it is expected that the final outcome will improve,
based on prior experience, another 2-3 efficiency points at low speed and expansion
ratio (the full benefit apparent from Figure 5 is reduced here due to the

238
multiplication of the combined turbine efficiency and the compressor isentropic
efficiency, which is of course less than 1.0). Stand results for the high efficiency
turbocharger with ball bearings are pending introduction of ball bearing hardware in
the proper frame size, and are not available by time of final draft in February 2014.

6 SUMMARY / NEXT STEPS

The development period post Euro VI / EPA ‘10 finds diesel engine OEM’s seeking
combined combustion and aftertreatment solutions to the challenges posed by new
fuel economy-focused legislation regulating diesels. The proper air handling
hardware will be required to support attainment of those goals, and the authors
believe it is clear that relaxation of the negative pressure scavenging requirement
associated with the need to drive large amounts of high-pressure loop EGR, coupled
with the potential to narrow the range of operating speeds via the vehicle-level
strategy of engine downspeeding, presents opportunity for application of high
efficiency low-specific speed aerodynamics. It has been shown that low-specific
speed aerodynamics can tend to higher individual stage efficiencies, but also
increased inertia and package size, when compared to typical high specific speed
radial flow solutions.

Finally, it is the authors’ contention that a carefully-matched combination of low


specific speed stages can result in attainment of aggressive overall turbocharger
efficiency goals, in this case in excess of 60% overall turbocharger efficiency. Next
steps for the development involve incorporation of ball bearings into the high
efficiency turbocharger and also demonstration of stage and solution scalability.
Additionally, potential improvements to the delivered compressor map topology will
be considered.

REFERENCE LIST

(1) Such, C., Technology Roadmaps - Heavy Duty Diesel Engines, Ricardo plc,
RD.12/111701.1, 2012
(2) Balje, O., Turbomachines: A Guide To Design, Selection, and Theory, John
Wiley & Sons, New York, 1981
(3) Japikse, D., Centrifugal Compressor Design and Performance, Concepts ETI,
Inc., ISBN 0-933283-03-2, 1996
(4) Baines, N., Fundamentals of Turbocharging, Baines, Concepts ETI, Inc., ISBN
0-933283-14-8, 2005
(5) Whitfield, A., Baines, N., Design of Radial Turbomachines, Longman Group,
London, 1990
(6) Spence et al., 2007, A Direct Performance Comparison of Vaned and Vaneless
Stators for Radial Turbines, Journal of Turbomachinery, Vol 129, p53-61
(7) A Fast and Elitist Multiobjective Genetic Algorithm: NSGA-II, K. Deb, IEEE
Transactions on Evolutionary Computation, Vol. 6, No. 2, April 2002
(8) Adaptive Filter SQP, A. Turco, Lecture Notes in Computer Science Volume 6073,
2010, pp 68-81, 978-3-642-13799-0

239
Advanced boosting technology to meet
future heavy duty diesel engine
requirements
A Banks, R Cornwell, C Such
Ricardo UK Ltd, UK

ABSTRACT

Future Heavy Duty Diesel (HDD) engines will increase their specific power output,
which in turn will increase the overall pressure ratio requirements of the engine
boost system. This paper investigates the possibility of existing that the highly
efficient technology used in larger turbo machinery may be scaled down to HDD
frame sizes. The opportunity for a new, high pressure ratio, single stage
turbocharger is identified, as a way to satisfy future HDD engine requirements
without the cost and complexity of two stage systems.

DEFINITIONS AND ABBREVIATIONS

BSFC –Brake Specific Fuel Consumption


BTE – Brake Thermal Efficiency
CARB – California Air Resource Board
CO2 – Carbon Dioxide
DPF – Diesel Particulate Filter
EFTA – European Free Trade Association
EGR – Exhaust Gas Recirculation
EPA – Environmental Protection Agency (US)
EU – European Union
GHG – Green House Gases
GVW – Gross Vehicle Weight
HDD – Heavy Duty Diesel (~2L/cyl engine capacity)
MDD – Medium Duty Diesel (~1l/cyl engine capacity)
NOx – Nitrogen Oxides
OEM – Original Equipment Manufacturer
PM – Particulate Mater
PN – Particle Number
SCR – Selective Catalytic Reduction
VGC – Variable Geometry Compressor
VGT – Variable Geometry Turbine
VVT – Variable Valve Timing

1. INTRODUCTION

As a result of legislation in the USA and market pressures in all regions, future
heavy duty vehicles will need to decrease their Green House Gas (GHG) emissions
including CO2 by around 20% over the next 10 years. Changes in engine technology

_______________________________________
© The author(s) and/or their employer(s), 2014
241
make up a substantial part of this reduction and boosting systems have a
significant part to play. This paper summarises the expected developments in HDD
engine and associated boosting system technology.

Heavy Duty (HD) vehicles are typically defined by a Gross Vehicle Weight (GVW)
from 15 tonne up to 40 tonne (certain countries allow 60 tonne) and are fitted with
engines from about 8 litre to 16 litre displacement with power levels up to 750 HP
(559 kW). HD vehicles cover up to about 200,000 km per annum with a target life
before first major overhaul of up to 1.6 Million km. Medium Duty (MD) vehicles
have GVW < 15 tonne, are fitted with engines of about 4 – 8 litre, travel up to
150,000 km per year and have a life before first major overhaul of up to 800,000
km.

Within the EU, trucks and vans carry about 75% of freight transported over land,
delivering 18 billion tonnes of goods per year [1]. The expected growth of the
European Medium Duty (MD) and Heavy Duty (HD) vehicle market is shown in
Figure 1.

EU27+EFTA3 Sales and Parc Forecast

Commercial Vehicle Parc [Millions]


Annual Sales [Thousands]

450 40

400 38

350
36 MD
Sales
34
300
32
250
30 HD
200 Sales
28
150
26

100
24
Total
Sales
50 22

0 20
2004 2006 2008 2010 2012 2014 2016 2018 2020 Total
Parc
Year

Figure 1: European annual medium duty and heavy duty truck sales
including projection until 2020 [2]

The US annual sales volumes are similar to Europe; whilst China is producing in
excess of 1 Million HDD engines / trucks per annum, with growth rates exceeding
those of Europe and the US.

2. EURO VI LEGISLATION AND BEYOND

Euro VI legislation was introduced for heavy duty vehicles in January 2014. In the
next period, focus is likely to be on In Service Conformity, which is intended to
ensure that the requirements of Euro VI are met on the road.

In parallel, developments are in hand to introduce measures to control CO2 or fuel


consumption. The European Commission has sponsored the development of a
methodology for calculating CO2 emissions from a wide range of heavy-duty
vehicles [3]. This may lead to the promulgation of standards on CO2 emissions in a
timeframe indicated in Figure 2.

242
Although not the subject of this paper, another strand in engine development over
the next decade will be the utilization of alternative fuels and, in particular, natural
gas, which offers potential advantages in terms of CO2 emissions, fuel price and
energy security.

Figure 2: European emission legislation and expected fuel consumption


legislation

Due to the trade-off between NOx emissions and CO2 emissions, a major question
is whether a further round of NOx legislation will be introduced beyond Euro VI. It
is worth noting that the California Air Resources Board has proposed a voluntary
NOx limit which represents a 90% reduction below current US NOx limits i.e. down
to 0.02 g/kWh. In Europe, further reduction in NOx will depend on the results of air
quality monitoring. If as expected the Euro VI regulations cause a lowering in
ambient NOx, then no further tightening of NOx limits can be expected.

3. EURO VI ENGINE TECHNOLOGY AND BOOSTING SYSTEMS

The factors which most influence an engine’s boosting system requirements are:

 The highest engine power and torque levels over the speed range, which is
in turn defined by the engine’s design Maximum Cylinder Pressure (Pmax)
limit
 Exhaust Gas Recirculation (EGR) rate: the lower the EGR rate which is
needed to achieve the required engine-out NOx emissions, the lower the
boost pressure requirement
 Combustion system: this influences the minimum Air/Fuel Ratio which can
be used to achieve very low engine-out soot emissions

Other factors which have a secondary influence on the turbocharger are the
pressure drop across the engine (back pressure of the exhaust aftertreatment) and
whether the engine has open or closed crankcase breathing.

In recent years, new HDD engines have been designed to a Pmax limit of 200 – 210
bar (actual operation about 10% below the design limit) and this is expected to
increase towards 230 – 240 bar before the engine reaches a limit dictated by
increasing engine weight and cost.

Emissions control technology is common to all Euro VI HDD engines: Selective


Catalytic Reduction (SCR) for NOx control and Diesel Particulate Filters (DPF) for
control of particulate mass and particle number emissions. EGR is fitted to the
majority of engines. EGR rates have tended to reduce as SCR efficiencies have been
increased by thermal management of exhaust systems. This trend has enabled, in

243
one case, EGR to be deleted altogether in favour of high efficiency SCR (c. 96%
efficient) [4], and in another case, the EGR cooler has been deleted, to combine the
beneficial effects of hot EGR on exhaust temperature at part load and high
efficiency SCR at higher loads [5].

A range of boosting systems is employed on HDD engines, notably:

 Single stage turbocharger with wastegate and hot EGR [5]


 Asymmetric, single stage turbocharger used to drive EGR from one branch
of the exhaust manifold [6]
 Variable Geometry Turbocharger (VGT) with [7] or without cooled EGR [4]
 Two stage turbocharger with inter and after cooling and cooled EGR [8]

All on-highway HDD engines which use EGR employ the high pressure route where
EGR is taken from upstream of the turbine. The alternative, low pressure route
where EGR is taken from downstream of the DPF is used in some light duty engines
but is not considered sufficiently durable on HDD applications.

In terms of engine power rating, it is likely that specific power will increase from
the current level (typically 26 – 32 kW/litre) towards 34 – 36 kW/litre. This will
offer the potential to replace larger engines with more compact, lower displacement
engines in certain applications in the interests of CO2 reduction.

In summary, the future trends for HDD engines are shown below in Figure 3.

Specific ratings
Max cyl pressure
Engine rated speed
EGR rates
BSFC (long term)
SCR efficiency
Fuel injection pressure
Turbocharger pressure ratio >3.5
Increasing

Decreasing

Figure 3: Future HDD engine trends

4. OPTIONS FOR REDUCED CO2 EMISSIONS

Currently the highest Brake Thermal Efficiency (BTE) of a Euro IV/V engine at its
most efficient operating point is c. 45% (corresponding to about 188 g/kWh diesel).
The long term target is to increase this above 50%.

Turbocharger technology such as turbo compounding, which recovers some of the


exhaust energy by means of a turbine downstream of the turbocharger, can help to
reduce BSFC on long distance trucks. Around 5% improvement in fuel economy is
feasible for long distance trucks. The Scania DT11 engine originally launched in
1991 achieved a minimum BSFC of 186 g/kWh which equates to a BTE of 45.5%.
Other applications of mechanical turbo compound systems have been introduced on
a US 2010 engine [9] and a Euro VI engine [5].

244
Another method of reducing vehicle CO2 is down speeding the engine; Figure 4
shows an example from an ongoing project [10]. The effect of down speeding is to
reduce pumping losses and engine friction.

Figure 4: Down speeding on a medium duty diesel engine [10]

In this case it has been estimated that down speeding by about 18% has the
potential to improve fuel consumption by about 2% over a typical driving cycle of
regional long distance truck.

The down speeding is made possible by increasing the BMEP of the engine and
achieving constant power at a lower engine speed, and also the adoption of new
transmissions. The reduction in engine speed also reduces the flow range
requirements from the turbocharger, allowing higher pressure ratio and reduced
flow range compressor maps to be developed.

The down speeding concept has been introduced on a Heavy Duty vehicle [11], in
which case the transmission control in the vehicle maintains the engine in the most
economical speed range between 900 and 1400 rev/min.

The introduction of variable valve timing on the inlet valve closing (known as Miller
cycle) is being investigated at the research stage. Initial results appear promising
in terms of potential CO2 benefits [10]. If the Miller principle is used, this will
require an increase in boost pressure ratios above 3.5:1 to maintain the same
trapped AFR.

Despite its high cost and complexity, there may be a case for waste heat recovery
in applications which consume large quantities of fuel. Waste heat may be
recovered in the form of mechanical shaft power by turbocompounding, which has
proven durability in on-highway applications and has the added benefit of a power
increase. Alternatively, in the longer term, the Organic Rankine Cycle (ORC) utilises
heat from the EGR cooler and/or the exhaust downstream of the aftertreatment to
generate mechanical or electrical energy.

Assuming acceptable reliability and durability, the success of waste heat recovery
systems in developed markets will depend on payback times. A recent paper [12]
showed that, at current fuel prices, the payback time of an ORC system on a heavy
duty, on-highway truck was about 2 years in the USA and 1.5 years in Europe.

245
Waste heat recovery can also be converted into electrical power by electrical
turbocompounding or thermo-electricity. For generator sets in particular, the
electrical turbocompound approach [13] is a convenient approach.

5. TURBOCHARGER INDUSTRY REQUIREMENTS

The current solution for high boost pressure ratio requirements (>4.2) is to use 2
stage turbocharging, with or without inter-cooling. Compared with the single stage
turbocharger, this has a major impact on the system cost and introduces packaging
and complexity issues. Certain developing markets such as China are reluctant to
adopt solutions such as VGT and two stage which are seen as unacceptably
complex and expensive.

The challenge for the turbocharger industry is therefore to develop very high
pressure ratio single stage turbochargers.

If we look to the larger frame size machines used on non road applications, one
manufacturer produces a range of turbochargers which can achieve >5:1 pressure
ratio from a single stage aluminium compressor. The high pressure ratio
compressor is achieved using a vaned diffuser, and the use of an aluminium
compressor wheel is made possible by additional compressor blade cooling.

Figure 5 shows a very high pressure


compressor (140mm diameter) with a
peak of 80% compressor efficiency at
5.8:1 pressure ratio, developed for
larger non-road applications.

As noted above, the down speeding


trend reduces the speed range of the
engine which in turn reduces the flow
range requirements of the
turbocharger, which will allow similar
compressor design methods to be
introduced on HDD engines.

Until recently, on highway HDD


turbocharger development was
focused on driving EGR. As noted
above, future engines will have
reduced EGR levels or no EGR at all.
This will allow the turbocharger
designer to re-focus on overall
turbocharger efficiency. The reduction
in EGR will allow the complete use of
exhaust pulse energy, to maximise
turbine efficiency.

An interesting example of a new in-


house manufactured turbocharger is Figure 5: High pressure ratio
shown in Figure 6. The turbocharger compressor [14]
is a single stage unit, incorporating
an asymmetric turbine housing.

246
Figure 6: Current compressor and turbine maps for Daimler OM 471
engine [6]

The compressor has a very wide flow range and incorporates a ported shroud
housing design, the peak island efficiency of 80% is achieved at a low pressure
ratio (~2.0:1) and the maximum pressure ratio achieved is 4.2:1. The peak
turbine efficiency of 71% is achieved with a radial design wheel and asymmetric
turbine housing, used to drive EGR efficiently.

Figure 7 shows the future HDD engine trend for BMEP (at peak torque) and Pmax
as a function of compressor pressure ratio.

BMEP bar vs Comp P.Ratio Pmax bar vs Comp P.Ratio


4.4
4.4
4.2
4.2
4
4
3.8
3.8
3.6
3.6
3.4
3.4
3.2
3.2
3
3
24 26 28 30 32
190 200 210 220 230 240 250

Figure 7: BMEP trend at peak torque and P max trend plotted against
compressor pressure ratio requirements

Current BMEP levels of Euro VI engines at peak torque are about 24 – 26 bar. As
the rating increases up to 30 bar BMEP, the maximum cylinder pressure rises to
~240 bar, at which point the compressor pressure ratio is ~4.1:1. These data are
for sea level conditions only and assume a 13 litre HDD engine with standard valve
timings. The rating of 25 bar BMEP is currently the limit for a single stage
turbocharger, including an allowance for altitude operation to about 2000m without
engine de-rate.

6. FUTURE TURBOCHARGER TECHNOLOGY

6.1 Turbine Wheel and Housing Design


Turbine wheel design will focus on low specific speed wheels, driven by increased
mass flow (associated with lower EGR), and reduced engine speeds.

Turbine housing design may move away from VGT and back to divided designs
(both symmetric and asymmetric) to maximise pulse energy into the turbine. New
low cost variable turbine technology may be introduced also on niche applications.
Sector divided, rather than meridional divided, turbine housings, as used by larger

247
turbochargers, may improve turbine efficiency by better separation of engine
pulses.

Increasing the overall turbocharger efficiency will reduce the ability to drive EGR.
To offset this, alternative EGR concepts may be realised, such as low pressure EGR.
Ricardo’s investigations into a dual circuit EGR system [15] where EGR was taken
from both high and low pressure sources showed that good fuel consumption could
be realised whilst maintaining the ability to drive high levels of EGR.

6.2 Compressor Wheel and Housing Design


Compressor wheel designs will continue to develop low specific speed wheels, in
both aluminium and titanium. Ported shrouds will continue to be used, and further
developments of the port are being developed.

Abraidable coatings can assist in elevating compressor efficiency: this technology


has been developed for light duty applications. Wheel-to-housing clearance for
compressors will become more and more important as compressor pressure ratio
increases and wheel blade height reduces.

Variable Geometry Compressors (VGC) have been developed and offer


improvements in map width: this technology may also be applied to the vaned
diffuser [16]. Another technology which will continue to assist in increasing the
turbocharger overall efficiency is low friction bearings (including ball bearings)
which are already introduced on a medium duty application [17].

Higher pressure ratios will continue to drive developments of the seal design: this
will also assist in reducing crankcase pumping work on the engine.

7. COST OF TECHNOLOGY

Figure 8 shows the estimated trade-off of alternative boosting systems in terms of


pressure ratio capability and the cost compared with a wastegated single stage
turbocharger, based on typical HDD production volumes.

6 Opportunity new HP ratio single


stage unit
4 6
2
5
Pressure ratio capability

1 3 5
4

3 1 –W/G (single stage)


2 –New turbo HP (single stage)
3 – Low cost Variable turbine (single
2 stage)
4 – 2stage with fixed geom
5- VTG (single stage)
6 -2stage with VTG
1

0
0% 100% 200% 300%
Cost Increase Relative to w/g
* Estimated using Ricardo analysis

Figure 8: Turbocharger technology – compressor pressure ratio vs. cost


increase (baseline wastegate unit)

248
The current single stage turbochargers are limited to 3.5:1 pressure ratio at sea
level; above this, 2 stage solutions are needed with associated cost and packaging
issues. Technology level 2 in Figure 8 (high pressure ratio single stage
turbocharger) offers a high pressure ratio capability at a relatively low increase in
cost when compared to the baseline. Such a turbocharger would offer an interesting
potential solution to the problem of the engine manufacturer who wants to extend
the power range of his engine whilst maintaining commonality of the boosting
system.

8. CONCLUSIONS

New HDD engines will use very high BMEP >30 bar, and introduce high specific
ratings up to about 36 kW/litre. The new ratings will incorporate high P max
capability up to 230-240 bar.

These ratings will force further efficiency gains on the turbocharger, two stage
solutions will be required, and an opportunity exists for a high pressure ratio single
stage unit. This requirement is especially necessary if we consider turbo
compounding solutions.

The trend for boosting technology is to increase the overall pressure ratio
requirements > 4:1 in combination with very high overall efficiency. The new
generations of turbocharger will be combined with other technologies such as
variable valve timing to achieve high specific power and improved CO2 emissions.

High pressure ratio, single stage turbochargers are an attractive alternative to 2


stage turbochargers, providing significant benefits in terms of cost and package.

Technologies used on larger turbochargers are likely to be adopted for smaller


frame sizes. Significant single stage turbocharger efficiency gains are possible, and
are needed if we are to deliver higher specific engine ratings and >50% engine
Brake Thermal Efficiency (BTE).

9. REFERENCES

1. European Automobile Manufacturer’s Association ACEA publication “The


Automobile Industry Pocket Guide 2013”
2. LMC Automotive, ACEA, Ricardo Analysis, Automotive World
3. Hausberger, S. “Progress in CO2 and Fuel Efficiency Goals and Measurement
Methodologies” presented at 9th Diesel Emissions Conference & AdBlue Forum
Europe 2013, Duesseldorf, Germany, June 2013
4. Signer, M. “Meeting Euro VI and EPA10 Legislation without EGR” SAE HDD
Emissions Conference, Gothenburg, September 2012
5. Nakano, D. “Volvo Group Engine Evolution to Meet Euro VI Emission Standard”
SIA Conference, Lyon, November 2012
6. Gruden, D. et al “The 2013 Daimler DD15 Engine – Embracing NAFTA’s 2014
GHG and 2013 OBD Regulations with Benchmark Fuel Economy” Aachen
colloquium, October 2013
7. de Kok, P. et al “Common rail Integration on PACCAR Heavy Duty Truck
Engine” Aachen Colloquium, October 2013
8. von Hoerner, R., et al “Turbocharger Concept for Current and Future
Commercial Vehicles” 12th TU Dresden Technical Conference on
Supercharging, Sep 2007
9. Heil, B., et al “New Daimler Heavy Duty Commercial Vehicle Engine Series”
29th Vienna Engine Symposium, April 2008

249
10. Such CH, Edwards, S.P. “Potential reduction of CO2 emissions from Heavy
Duty Diesel engines for long distance transport” ATZ Live Conference,
Ludwigsburg 5 & 6 November 2013
11. Volvo iShift transmission - www.volvo.com
12 Stanton, D. “The Role of Waste Heat Recovery in Meeting Phase 2 US EPA
Greenhouse Gas Regulations” 9th Diesel Emissions Conference, Duesseldorf,
June 2013
13. Patterson, A.T.C, Tett, R.J., McGuire, J., “Exhaust Heat Recovery using
Electro-TurboGenerators”, SAE paper 2009-01-1604
14. Single stage high pressure ratio A100 compressor – www.ABB.com
15. Banks, A., Niven, M., Andersson, P., “Boosting technology for Euro VI and Tier
4 Final Heavy Duty Diesel Engines without NOx Aftertreatment”, presented at
Institution of Mechanical Engineers, London, May 2010
16. N Watson and MS Janota Turbocharging the internal combustion engine –
variable inlet guide vanes pg 135-136
17. http://turbo.honeywell.com/our-technologies/ball-bearing/ - use on Hino
medium duty truck

250
Considerations for gas stand measurement
of turbocharger performance
J B Schwarz, D N Andrews
BorgWarner Turbo Systems - North American Technical Center, USA

ABSTRACT

The SAE “Turbocharger Gas Stand Test Code” (SAE J1826) has historically defined
the best practices for testing of turbochargers on gas stands. First issued in 1989,
this specification no longer reflects the state of the art. To obtain consistent
thermodynamic mapping of turbochargers, additional practices beyond the scope of
J1826 must be followed, ensuring accurate, realistic results while also enabling a
fair comparison of product performance between manufacturers of turbochargers.
Using empirical gas stand test results as a basis, key issues are identified which
may produce misleading findings from turbocharger gas stand performance testing.

NOMENCLATURE

Compressor Parameters Turbine Parameters


Compressor Inlet Pressure Turbine Inlet Pressure
Compressor Inlet Temperature Turbine Inlet Temp
Compressor Discharge Pressure Turbine Discharge Pressure
Compressor Discharge Turbine Outlet Temp
Temperature Cp, ex Specific Heat of Exhaust Gas
Compressor Reference Pressure Ratio of Specific Heats for
Compressor Reference Exhaust
Temperature Turbine physical mass flow
Cp Specific Heat of Air Φ Turbine Flow Parameter =
Ratio of Specific Heats for Air
Compressor physical mass flow Π T-S Expansion Ratio =
Compressor Corrected Flow = , / ,
,
, T-S Turbine combined efficiency
Turbine Power

Π Compressor Pressure Ratio = Bearing Housing Parameters


/ Oil Inlet Temp
Compressor efficiency = Heat Flux at bearing housing
( )/ diffuser surface

Compressor Power

_______________________________________
© The author(s) and/or their employer(s), 2014
253
1. INTRODUCTION

The SAE “Turbocharger Gas Stand Test Code” (SAE J1826) was created as a
guideline to ensure consistent thermodynamic mapping of turbochargers across the
industry. (1) The field of turbocharging has evolved considerably since this
specification was outlined in 1989, and shortcomings in SAE J1826 undermine the
original purpose of ensuring consistent, repeatable, and accurate turbocharger
performance test results. Aspects of gas stand configuration and test operation are
not specified in sufficient detail in the specification. Stage performance and map
width can be altered through manipulation of these parameters.

1.1 Overview of Gas Stand Hardware


The gas stand provides a steady-state hot gas test bed on which measurements of
turbocharger stage performance can be made. The accuracy of performance figures
is sensitive to the design of the gas stand itself as well as the selection and location
of measurement devices. A more thorough discussion of modern gas stand design
and capabilities is provided by Young and Penz. (2) Additional reading is provided in
Mai’s study of significant parameters in gas stand turbocharger testing. (3)

Thermodynamic mapping depends on accurate measurement of temperature and


pressure changes across each stage as well as measurement of flow through the
stage. Figure 1 shows a typical layout of adapters and measurement sections
required for temperature and pressure measurement on the gas stand. Turbine
and compressor flow measurement devices are external to the components shown
in Figure 1.

Figure 1: Typical measurement sections and adapters found on a gas stand


used for turbocharger performance measurement

2. SURGE LINE AND MAP WIDTH

Surge is a compressor instability phenomenon marked by bulk flow reversal across


the stage. Practically, the surge line represents the minimum operable flow rate of
the compressor over a range of speeds. The surge event is triggered by stalled flow
conditions within the compressor stage; however, as a system instability, surge is
sensitive to numerous aspects beyond the compressor itself. The volume and
geometry of the compressor inlet and discharge ducting play a significant role in

254
determining whether the system will be destabilized, causing a surge. (4) SAE
J1826 makes no provisions regarding the design of compressor ducting.

Specifically, compressor inlet restriction due to ducting has been observed to have
a marked effect on the surge line, as seen in Figure 2. This result was obtained by
testing a single turbocharger with multiple levels of inlet restriction.

Similarly, it has been observed that a suction-side butterfly throttle valve can move
the surge line. (5) Pre-swirl or counter-swirl inlet vanes can also be employed to
shift the operating space of the compressor map. (6)

Given the dependence of surge on duct geometry, a more representative surge line
can be obtained by using application ducting. However, the only reliable method for
obtaining a consistent surge line is to maintain a consistent compressor ducting
setup across all turbocharger testing. This provides a repeatable surge line and
generally allows for a quantitative comparison between turbochargers.

As surge is triggered by a stall event, the onset of surge is particular to whether the
stall is rotating or stationary and whether it occurs at the impeller or in the diffuser.
This means that different turbochargers can experience surge at the same
compressor flow conditions for different underlying reasons. As a result, various
turbochargers may respond differently (in terms of surge) to changes in the system.

3.8

3.4 Surge line with compressor


inlet restriction
Pressure Ratio (Total-to-Static)

3.0

2.6

2.2

1.8

1.4

1.0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24 0.26

. .
Figure 2: Surge line shift due to change in system inlet restriction

3. HEAT TRANSFER EFFECTS

Turbocharger stage efficiency calculations depend on accurate calculation of


enthalpy change across compressor and turbine stages. Although the system is
assumed to be adiabatic, a range of non-adiabatic effects can occur, as seen in
Figure 3. Heat flux into or out of a compressor stage will cause a change in the
apparent efficiency of the stage. , is related through a power balance; therefore,
changes in result in an inverse effect on the turbine side.

255
Generally, these thermodynamic effects are found to have the greatest influence at
low flow rates and for small turbochargers. Here, non-adiabatic heat transfer
dominates relative to adiabatic compression and expansion. A number of
phenomena can change the thermodynamic balance, some dramatically affecting
stage performance results. SAE J1826 does not provide specific guidance for
constraining parameters related to heat transfer.

Heat transfer between bearing


housing & lubricating oil

T1 influence

T3 influence

Conductive heat Radiative, conductive


flux through heat flux from turbine
bearing housing housing
Convection at diffuser face

Figure 3: Examples of turbocharger heat transfer phenomena

3.1 Compressor-Turbine Power Balance


Compressor efficiency can be calculated by determining the stage enthalpy change,
a function of the pressure and temperature rise across the stage. To find the
turbine efficiency, a robust technique is to use the turbocharger power balance to
obtain the combined aerodynamic/mechanical turbine efficiency.

The compressor power is described by


/
Π 1

The power required by the compressor is taken to be equal to the power delivered
by the turbine. The turbine power equation takes a similar form,

256
= , , 1 − (1/Π )

, includes both turbine aerodynamic efficiency and bearing system frictional


losses. Assuming = , the equation can be solved for , . In this
representation, it can be seen that , is inversely dependent on . This
dependence on combined with the additional turbine side parameters ( , , ,
, etc.) forces the , uncertainty to always be higher than the uncertainty.

3.2 Insulation
Of the many phenomena that can affect heat transfer on a gas stand, insulation has
perhaps the greatest influence. Refractory insulation with a reflective backing can
significantly improve the accuracy and repeatability of turbocharger stage efficiency
measurements. Insulation prevents radiative heat transfer from hot surfaces and
reduces convective transfer of heat between turbocharger components and the test
cell, thereby greatly reducing the influence of ventilation systems on test results.
SAE J1826 specifies that the turbocharger should be insulated for testing; however,
many test facilities opt to test turbochargers without insulation.

Experiments were conducted to quantify the impact of insulation on several


turbocharger frame sizes. Baseline performance was determined for a turbocharger
with full insulation on the turbine inlet, turbine housing, and compressor cover.
Insulation was then progressively removed.

Insulation was found to have a reduced impact on medium to large turbochargers


typical of commercial diesel applications, but there was a pronounced effect on
smaller turbochargers. Figure 4 shows Total-to-Total (T-T) results for a small
turbocharger typical of passenger car and light-duty diesel applications. When
tested without insulation, the efficiency near surge increased by up to 4%, while
choke efficiency decreased.

2.6
3

2.4
2
Compressor PR (T-T)

2.2
1

2 0
Turbo without insulation
1.8 -1

1.6 -2

1.4 -3

-4
0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
Corrected Compressor Mass Flow (kg/s)
Turbo with insulation
Figure 4: Percent change in T-T compressor efficiency when run without
insulation, as compared with insulated case (left); test turbocharger
shown without insulation (upper right) and with insulation (lower right)

257
Of key importance, test results from insulated turbochargers cannot be compared
directly with un-insulated tests, particularly for small frame passenger car
turbochargers. From a repeatability standpoint, both insulated and un-insulated
test methods have their drawbacks. Un-insulated testing must be carried out with a
tight control of test cell temperature to reduce performance variation caused by
convective heat loss. Insulated testing shows little dependence on test cell
temperature, however the type and method of insulation has been observed to
cause and , variation.

There does not appear to be a clear consensus within the industry regarding
insulation practices. The authors’ recommend that a test facility tightly control test
cell air temperature and not use any insulation. Accurate control of test cell
temperature will require a dedicated cell HVAC system. This approach eliminates
variation caused by differences in operator methods of insulating the turbocharger.

3.3 Cooling Water


Water cooling can drive a significant change in apparent stage efficiency by
manipulating the enthalpy change across the stage. In order to project realistic
estimates of stage performance, turbochargers which feature water-cooled
components should be tested with cooling water flow rates and water temperatures
representative of the application. Figure 5 shows that for a compressor with a
water-cooled bearing housing, climbed from 58% to 70% with increased cooling
water flow. This effect diminished with increasing compressor mass flow rate. Since
, is computed via a power balance, this increase in results in a decrease in
, .

For aerodynamic development, the authors recommend that no cooling water be


supplied to the turbocharger. Such results are more indicative of raw aerodynamic
performance and can be compared fairly against a library of performance results for
non-cooled turbochargers. For application-specific development, the application
cooling water parameters can be replicated on the gas stand to provide a closer
match with application performance.

0.80
Compressor Efficiency (Total-to-Total)

0.75
100% Cooling
Water Flow
0.70

0.65

0.60

0.55
0% Cooling
Water Flow
0.50
0.00 0.04 0.08 0.12 0.16 0.20 0.24 0.28
(kg/s)
.
Figure 5: Influence of cooling water flow on compressor efficiency
.

258
Cold lubricating oil produces an effect on the compressor stage similar to that seen
for cooling water. Cold oil causes an apparent increase in a decrease in , due
to the compressor-turbine power balance. Additionally, a small decrease in , will
be seen due to increased bearing system drag due from higher oil viscosity.
Therefore, a consistent oil temperature should be maintained throughout all
thermodynamic testing.

3.4 Compressor Inlet Temperature


Measured compressor efficiency can be affected by changes in the compressor inlet
temperature, as seen in Figure 6. Consistent with other heat transfer effects, this
effect dominates at lower air flow rates and impacts small frame turbochargers
more so than larger units. This heat transfer effect occurs mainly in the parallel wall
diffuser section and is driven by convective heat transfer from the bearing housing
to the airflow. The bearing housing diffuser surface temperature is a function of ,
, and the inlet oil temp.

At low tip speeds, the diffuser surface will be significantly hotter than the air
leaving the compressor exducer. This convective heat transfer can be a significant
percentage of the turbo’s overall shaft power for small frame turbos at low tip
speeds. Testing on larger commercial diesel size turbos has not shown this effect
(i.e. large frame units are insensitive to changes). The effect of on can be
mitigated through control of the inlet air temperature to the compressor. The
authors’ recommendation is to tightly control T1 via a dedicated HVAC system,
especially if the test facility will be running small frame turbochargers.

RED: T1 = 50 C

BLUE: T1 = 35 C

BLACK: T1 = 25
Q

370 m/sec
230 m/sec

Passenger car turbo with


46mm exducer compressor

Figure 6: Influence of compressor inlet temperature and speed on


compressor T-S efficiency (left); compressor cross section showing
convective heat transfer through diffuser wall (right)

3.5 Turbine Inlet Temperature


Varying causes significant changes in calculated values of , , , Φ, and Π .

Variation in changes the heat flux into the bearing housing. This affects the
convective heat transfer into the airflow in the compressor diffuser section:

259
 As is increased for a given , , rises, and the resultant calculated
decreases.
 If a power balance is used to find , , the calculated value of , will rise
due to the higher calculation of compressor power.

These changes are most significant at low compressor speeds and flow rates. Small
frame passenger car turbochargers will show the most efficiency variation when
is changed.

When running turbine performance testing to a set of target turbine corrected


speeds, changing will cause significant changes in the turbine operating points.
As is raised, the Π range for a given speedline will increase, shifting to the right
on the turbine map. The Π shift causes the Φ values to slide up and to the right
along the Φ curve. In addition, the compressor operates at higher compressor tip
speeds and shaft power as is increased while operating at a certain set of turbine
corrected speeds. To minimize these various effects, it is recommended that a
consistent value be used for all thermodynamic testing.

3.6 A Final Note Regarding Compressor Heat Transfer Effects


When discussing heat transfer effects and their impact on , it must be noted that
none of these cause variation in the measured Π . These changes only impact
and the calculated – not Π .

4. TURBINE TOTAL TO STATIC PERFORMANCE AND TURBINE DISCHARGE


MEASUREMENT STACK SIZING

It is standard practice for automotive turbocharger manufacturers to report turbine


performance on a Total-to-Static (T-S) basis. This requires the total temperature
and pressure properties at the turbine inlet and only the static pressure at the
turbine outlet. It is challenging to accurately determine , and , at the
turbine discharge because of the variable rotational component of the flow velocity
and the large temperature gradient that exists in the flow; therefore the velocity
and density at the turbine discharge will not be known. Total measurement
techniques (such as pitot probes and total temperature probes) are not suitable for
high-volume automotive turbocharger testing.

The danger of reporting , on a T-S basis stems from its sensitivity to the turbine
discharge measurement section diameter. For a given turbo, the reported , can
be greatly manipulated by changing the discharge stack diameter. A large stack
diameter lowers turbine outlet velocities and creates more static pressure recovery.
This creates an increase in , , a reduction in Π , and an increase in the calculated
, . Figure 7 shows performance results with two different sizes of turbine
discharge stacks for a medium size commercial diesel turbo.

SAE J1826 does not provide guidance regarding sizing of turbine outlet
measurement stacks. For aerodynamic development, a measurement stack should
be selected so that a consistent ratio is maintained between the stack diameter to
turbine wheel inducer diameter. For example, the allowable range for measurement
stack diameter / turbine wheel inducer diameter could be 1.3 to 1.45. For
application-specific testing, a customer-specific outlet pipe diameter can be used.

260
0.675

89mm stack
0.650

68mm stack
Turbine Efficiency (Total-to-Static)

0.625

0.600

0.575

0.550

0.525
1.0 1.4 1.8 2.2 2.6 3.0 3.4 3.8 4.2 4.6 5.0 5.4
Expansion Ratio P3/P4 (Total-to-Static)

Figure 7: Influence of turbine discharge measurement stack diameter on


turbine efficiency

5. TURBINE DISCHARGE PRESSURE

Turbine efficiency can often be influenced by . Most gas stand testing is done at a
nominal that is not representative of the actual application. Application is
usually close to atmospheric pressure but can be elevated in specific cases due to
exhaust system piping, two stage architecture, and after-treatment. Therefore,
application should be considered when interpreting , results.

Testing conducted on a medium-sized commercial diesel turbocharger showed that


aggressive changes in turbine outlet pressure produced a change in , , as seen in
Figure 8. At lower flow rates, the change in caused the turbine operating point to
shift, following the general shape of the speedline curves for , and Φ. However,
at higher Π , the change in produced a marked shift in , and Φ . In this
scenario, the turbine showed elevated , as was increased.

The effect of varying on , is not universal for all turbine wheel designs.
Testing on smaller passenger car size turbochargers has shown little to no change
in , when varying .

For application-specific testing, control over is easily accomplished via an


exhaust throttle valve placed downstream of the turbine discharge measurement
section. Ideally the valve would be adjusted to approximate the levels expected
in the application. Testing for aerodynamic development necessitates that turbine
performance be obtained at consistent conditions to permit direct comparison
with an existing library of turbine performance data.

261
0.69

0.67

0.65
360 m/sec
Turbine Efficiency

tip speed
456 m/sec
0.63
tip speed

0.61

0.59

224 m/sec Normal P4 Moderate P4


0.57 tip speed
High P4
0.55
1 1.5 2 2.5 3 3.5 4 4.5

Turbine Expansion Ratio (T-S)

Figure 8: Effect of back pressure on turbine efficiency

5.1 Bearing System Losses Due to Oil Temperature and Viscosity


When computing turbine combined efficiency using a compressor-turbine power
balance, the bearing system losses are incorporated into the reported , . These
losses are generally less than 1%, but the effect is somewhat more for very small
turbochargers. In this circumstance, oil viscosity has a modest effect on , . In
addition, oil temperature can affect the system through two mechanisms: an
increase in temperature causes a reduction in viscosity and also changes the
bearing housing heat flux.

Decreasing Oil Viscosity


Turbine Efficiency

Passenger car turbo with


29mm exducer compressor

Expansion Ratio

Figure 9: Reducing oil viscosity causes an apparent increase in turbine


efficiency for a small turbocharger

262
As seen in Figure 9, , can be shifted by approximately ±1% for a very small
turbocharger as a function of the viscosity and formulation of the oil. This effect is
most pronounced at the lower speedlines and rapidly diminishes with increasing
turbine flow. Other frame sizes of turbochargers were tested, but oil viscosity
influence was found to be insignificant aside from small turbochargers typical of
passenger car applications.

For testing of small frame turbochargers, it is recommended that a consistent oil


viscosity and formulation be maintained. If the oil type is changed, thermodynamic
performance should be compared against reference data to evaluate the impact of
the change on stage performance.

6. CONCLUDING REMARKS

This paper has showcased several significant phenomena which can influence
turbocharger test results, and other effects exist beyond the scope of this paper.
When thermodynamic test results are obtained in accordance with the SAE
“Turbocharger Gas Stand Test Code,” there are many caveats not provided for in
this specification. Results must be taken with a degree of uncertainty, unless the
testing facility can provide specifics of the test procedure, hardware configuration,
and measurement techniques. Similarly, when comparing product performance
between turbocharger manufacturers, these caveats also apply.

Frequently, gas stand performance maps are used to produce engine power
simulations. Although on-engine testing currently provides the most realistic view
of turbocharger-engine system performance, the accuracy of gas stand mapping
has improved significantly due to knowledge of the factors discussed in this paper.
Future gas stand development promises greater accuracy for engine simulation.

A gas stand turbocharger testing facility generally serves two unique customer
types: aerodynamic development groups and application-specific customers. The
needs of these customers will often dictate certain testing conditions and practices.

Insulation has perhaps the most effect on apparent compressor and turbine stage
performance. Although SAE J1826 calls for the turbocharger to be insulated during
testing, the authors are aware of several facilities which test turbochargers without
insulation. A test facility must maintain a consistent practice to permit direct
comparison with a data library. Ideally, the test facility would use no insulation but
would employ a dedicated HVAC system to control test cell air temperature.

The compressor map shape and surge line are affected by the design and material
of the compressor ductwork, especially the resistance of the inlet ducting. To
ensure a consistent map shape, the compressor duct configuration must be kept as
similar as possible between multiple gas stands.

Several environmental factors impact the thermodynamic balance of the


turbocharger. To permit comparison of stage performance, oil inlet temperature
and T should be kept at constant values. Water cooling should not be supplied to
bearing and turbine housings during testing for aerodynamic development. Stage
performance obtained using water cooling cannot be compared directly with
existing non water-cooled performance data. Environmental heat transfer effects
are most significant at low flow rates and on smaller turbochargers.

, is sensitive to the diameter of the turbine measurement stack relative to the


wheel size. , increases as a function of the stack diameter. Additionally, higher
can cause an apparent increase in , , especially for large, higher-flowing

263
stages. Aerodynamic development work is best suited by maintaining a consistent
ratio of turbine stack diameter to wheel diameter for all turbine performance testing.
For application-specific testing, the turbine stack diameter should closely match the
application’s downpipe diameter.

Small passenger car turbochargers show a sensitivity to oil viscosity and


temperature, factors which affect the bearing system frictional losses. Where , is
found via a power balance approach, these factors will affect calculated , .

REFERENCE LIST

1. SAE Engine Power Test Code Committee. Turbocharger Gas Stand Test Code.
March 1995, 01. SAE J 1826.
2. The Design of a New Turbocharger Test Facility. Young, M.Y. and D.A. Penz.
Detroit, Michigan, USA : SAE International, 1990. SAE International Congress
and Exhibition. pp. 40-55.
3. Mai, Holger. Parameterstudie zur Turbolader-Kennfeldvermessung. Berlin : VFI /
Technische Universität Berlin, 2009.
4. Surge and Rotating Stall in Axial Flow Compressors, Part I: Theoretical
Compression System Model. Greitzer, E.M. s.l. : ASME J. Eng. for Power, April
1976, Vol. 98, pp. 190-198.
5. Practical Aspects of Centrifugal Compressor Surge and Surge Control. Boyce,
M.P., et al. College Station, Texas, USA : Texas A&M University, 1983.
Proceedings of the Twelfth Turbomachinery Symposium. pp. 147-173.
6. Stall, Stage Stall, and Surge. Japikse, D. College Station, Texas, USA : Texas
A&M University, 1981. Proceedings of the Tenth Turbomachinery Symposium.
pp. 1-13.

264
Critical aspects in turbocharger testing
H Bolz a, A Rinaldi b, A Kaufmann c
a
Kratzer Automation AG, Germany
b
CRITT M2A, France
c
Hochschule Weingarten, Germany

ABSTRACT

This article is focused on the quality criteria of the measurement results from Turbo
Charger Test Benches. The aim is to define clearly and precisely the quality criteria
for the repeatability and reproducibility of the measurement results of turbo charger
test benches, in order to give reliable measurement results for later use in engine
process simulation and engine control unit calibration. The testing facility can be
regarded as a complex measurement system. The results are affected by several
parameters and boundary conditions. Some of them can be influenced by the test
system itself; others are affected by the environmental conditions. An analysis of
these phenomena is presented. In particular this paper discusses the conditions
which can be influenced by the test bench (Measurement and Control) and defines
a test procedure to verify the quality of the test results. References are made to the
method of “Measurement System Analysis” as far as it is compatible to the
complexity of turbo charger measurement.

Intensive study of repeatability and reproducibility is reported and some


considerations of linearity and long term stability are presented.

This work gives an analysis of these aspects to support testing procedures as they
are known to be important in turbocharger development and comparison.
Limitations and possibilities of testing repeatability are explained and the critical
factors in the quality of turbocharger testing are discussed.

NOMENCLATURE

Symbol Physical value


, Compressor corrected mass-flow
u , Compressor corrected rotational speed
η , , Total-to-total compressor isentropic efficiency
Π , Total-to-total compressor pressure ratio
Volumetric-flow
Mass-flow
p Pressure
Q Heat flow
T Temperature
φ0 Relative ambient humidity

_______________________________________
© The author(s) and/or their employer(s), 2014
265
1 INTRODUCTION

The characteristic maps of turbochargers [TC], both for compressor and turbine,
are the main sources of information to quantify the performance of the machine.
They constitute the basic input for application of a turbocharger, for engine process
simulation, engine control unit calibration, research activities and several other
tasks in internal combustion engine development processes. This justifies the
importance of measurement and the corresponding boundary conditions.

These topics are well known issues in the turbocharger environment. The most
important entities are known and still little literature is available. Some
recommended procedures can be found in the Turbocharger test stand test code (1)
and related literature (2) (3). Following the increasing demand for turbochargers
and the speed up in research activities, some work on measurement uncertainties
has been reported by Brun and Kurz (4) and others (5) (6) Chapman and Schultz
(7) which provide information on gas turbine field test measurement uncertainties.
Recently E. Guillou (8) reported on the uncertainty and measurement sensitivity of
turbocharger compressor gas stands. Following the approach of E. Guillou, it is
intended to make public some studies with the specific goal of improving the
knowledge in this field.

This paper will discuss the impact of different parameters starting from the analysis
of the different parameters affecting the measurements. A mathematic approach is
discussed for comparison of two different test benches with identical equipment
with respect to comparability and repeatability. Results are obtained following a
specific procedure of using MSA (Measurement System Analysis). The overall study
focuses on some critical aspects of turbocharger measurement and works out basic
quality criteria on measurement results.

2 INFLUENCE PARAMETERS AND TEST SETUP

2.1 Basic Boundary conditions


All the experimental results presented in following pages have been obtained by
two different test benches both produced by Kratzer Automation AG. Care is taken
to compare results from exactly the same setup. One single set of turbocharger,
adaptation parts and measurement pipes are used in the study.

The test rig is a continuous flow apparatus particularly suitable for performing
investigations on exhaust turbochargers. It is characterised by the availability of
two independent supply lines for compressor and turbine. The turbocharger turbine
is fed with compressed air at defined temperature while the compressor can
operate at a controlled pressure level. Compressor back pressure is controlled by
an appropriate back pressure regulating system. The flow rate measurement can
vary from 0.01 kg/s to 0.25 kg/s. The turbine hot gas temperature can be
regulated from 150°C to 1200°C. The temperature is obtained by heat supply from
a 200 kW natural gas burner. A lubrication rig is used to ensure the correct
operation of the bearings. Three temperature probes type Pt100 are inserted in the
fluid vane both at compressor inlet and outlet. Three thermocouples type K are
inserted at the turbine interfaces. Standard pressure sensors are placed at the
interfaces on a stabilization ring which physically averages three pressure tappings
on the wall of the flow. Air humidity, test cell air conditions and turbocharger rim
speed are acquired with commercial sensors.

Figure 1 shows the principal points where the measurements used for the
thermodynamic calculations and gives a rough overview of the heat flow in the test
object.

266
Figure 1: Turbocharger measurement set up

2.2 Classification influence parameters


The influence of factors, affecting a turbocharger map, can be grouped in at least 5
different classes:
1 Environment of infrastructure
2 Environment of the test system
3 Specimen
4 Accuracy of the instrumentation
5 Test system control

Table 1: Explanation of each class of influence factors

Class Description Qualification


1 Influences caused by the ambient These parameters are generally
conditions of the complete test cell and given by the initial set up of the
the infrastructure for conditioning test cell. Modifications are
systems such as p0, T0, temperature mostly combined with
around the specimen, setup of venting reconstruction for the test cell
system, back pressure of exhaust gas
line, etc.
2 Influences caused by the piping of the These parameters generally
test system in the test cell such as construction factors and are
volume between TC and flow controller, correlated with the conditions in
pressure loss between ambient and the lab.
compressor input, etc.
3 Influence caused by production These influence parameters can
tolerances and runtime of the specimen be excluded by using a
dedicated specimen.
4 Measurement uncertainty in every Parameters mainly fixed and
sense, such as pressure, temperature are determined by the
and flow rate uncertainty instrumentation.
5 These parameters are generally given Can be influenced by the
by the strategy of controlling and correct control and data storage
storing results and have to be strategy
accounted for a repeatability study.

267
When comparing results from different facilities, all these effects need to be taken
into account. Concerning classes 1 and 2, few things can be done. Different setups
will bring differences in the results. As pointed out in Table 1, only some significant
modifications allow big improvements. These operations are generally not possible
in the running phase of a turbocharger development project. Concerning class 4 a
trade-off between sensor capabilities and cost gives the best solution. Regarding
classes 3 and 5 every effort should be made to define a standard setup between
different labs. With this general statement in mind, the strategy for the
repeatability and comparability study is defined with the aim of quantifying the
influencing parameters of classes 1,2 and 4.

2.2.1 Test environment


In order to rule out the influencing parameters of the environment conditions
corresponding to class 1 and 2 as much as possible, the following requirements
have been defined:

Strong requirement
 Level of intake temperature at compressor inlet Range of 20..25 °C
 Level of temperature variation over the test 0,5°C/h
< 1,5°C during test
Weak requirement
 Level of cell temperature around specimen Range of 30..40 °C
 Level of back pressure exhaust gas system 20 mbar (max.)
 Airflow control based on heat input in the
test cell, to avoid extreme airflow around TC

2.2.2 Specimen setup


In order to rule out the influence parameters of specimen corresponding to
influence parameter class 3 as far as possible, the following setup was adopted:
 One single physical Turbo Charger KP35 from Borg Warner
 4 x Adaptation parts between TC and measurement pipes
 4 x Measurement pipes with a diameter of 50 mm
 Fixed installed rim speed measurement
 Fixed installed temperature sensors (Pt100 on compressor side/
Thermocouples type K on turbine side)
 The test TC is prevented from entering surge or choke in order to prevent
high wear or damage.

2.3 Test setup


All results presented in the pages below are obtained by the use of the same test
sequence. The specimen is run on three consecutive measurement points on three
speed lines for ten times. The turbine working point is maintained at T3=600 °C
inlet temperature. Oil inlet conditions (temperature and pressure) are kept constant
during all the tests.

The following Table 2 summarizes the compressor working points.

Table 2: Measurement operation points

Cor. speed [m/s] 230 370 490


0,025 0,035 0,045
Cor. flow [kg/s] 0,050 0,070 0,085
0,080 0,095 0,108

268
3 SENSOR ERROR AND DATA EVALUATION

Experimental determination of the thermodynamic properties from measurement is


not precise. There are many sources of error as classified below. This makes the
measured value different from the real physical conditions. For the development
and comparison of turbochargers it is hence crucial to consider the possible errors
since they are fundamental to design evaluation.

3.1 Sensor Errors


A sensitivity analysis was run on some calculated variables in order to characterise
the influence of measurement uncertainties on the calculation of total-to-static
pressure ratio and total-to-static efficiency.

As the total error of the measured value largely depends on the installation, only
the possible sensor errors given in Table 3 are considered since the focus lies on
the error propagation. In addition to ambient temperature T0, ambient pressure p0,
relative ambient humidity φ0, inlet temperature T1, exit temperature T2, inlet
pressure p1, exit pressure p2, the recovery factor for temperature measurement rf
and the change in wheel speed uc on one target speed line are considered.

Table 3: Considered sensor errors

δT0 δp0 δφ0 δT1 δT2 δp1 δp2 δrf δuC


[K] [Pa] - [K] [K] [Pa] [Pa] - [m/s]
0.15 40 0.05 0.15 0.15 190 400 0.1 0.1%u

The error is computed using a Gaussian type error propagation analysis. A


numerical approach (central differences) is used to compute the error due to the
individual measurement error. The error computation for the efficiency is given
below.

∆ = ∑ ∆ ≈ ∑ Δ ( ) (1)

…, ,… …, ,…

Δ = …, + , … − …, − , … (2)

For the evaluation of the individual η … , x ± δx , … full data evaluation using


enthalpy and entropy correlations of the gas standards of humid air composition
was performed.

3.2 Compressor efficiency and pressure ratio maps


This error analysis has been applied to the previously discussed nine test points in
the compressor map. The computed error is quantified by an error bar in the y
direction. The error bar in the x direction corresponds to the 1% uncertainty of the
mass flow rate measurement.

269
Figure 2: Compressor map with error bars

Figure 3: Compressor efficiency map with error bars

In the case of the pressure map an additional error is present when two maps are
compared. Since the compressor wheel speed varies slightly along the iso-speed
line (δuC), the representation is not completely exact. As the pressure ratio
increases as the square of the wheel speed, this error becomes significant at higher
compressor wheel speeds and pressure ratios. This error can be quantified by a
simple analysis.

/( ) /( )
= = 1+ (3)

/( )
= 1+ (4)

For the circled point on the efficiency and pressure map a sensitivity analysis is
carried out. The results of the sensitivity analysis are summarized in Figure 4 and
5.

270
3.3 Sensitivity to individual sensors

Figure 4: Compressor pressure ratio sensitivity to measurement

Figure 5: Compressor efficiency sensitivity to measurement

The largest error due to sensor error in the efficiency map is given by pressure
measurement, followed by the temperature measurements before and after the
compressor. The measurement of ambient pressure, temperature, humidity, mass
flow rate and the recovery factor are of minor importance. The error in efficiency is
larger for the smaller iso-speed lines as the relative error in pressure and
temperature measurement are much larger. Investment in precise temperature and
pressure measurement would hence give the best benefit.

A significant error in the pressure map is due to the 0.1% compressor wheel speed
error. Longer holding time can cut down this error significantly at the price of
longer measurement times. When this error is neglected, the source of error is
dominated by the pressure measurement before and after the compressor. The
error due to data evaluation with dry air and constant heat capacities compared to
full thermodynamic evaluation of enthalpies and entropies contributes to roughly
0.5 % of the efficiency.

271
4 COMPARABILITY

In order to quantify the reproducibility error of the same test specimen the same
unit will be measured under the same conditions by two different test stands. The
comparability error is then defined as the differences between the results.

4.1 Comparability mathematic model


The typical map at a turbocharger test stand provides information regarding the
functional relationship

= , (5)

between the signals , and . For the compressor test setup the signals are

: corrected Mass flow


: corrected rim speed
: pressure ratio (total to total)

Corrected mass flow and corrected rim speed represent the independent
variables, and pressure ratio is the dependent variable that will take values
according to the functional relationship , . At the test stand, the independent
variables are set by controllers. This is the difference to the sensitivity analysis via
data evaluation in section 3 and will be an additional source of discrepancies. Note
that the function , mainly depends on the test specimen and is a priori not
known. The measured data is gathered at discrete support points. Therefore, data
samples , and are considered instead of one data point.

In order to provide information on how far the measurement data of one test stand
can deviate from another test stand due to the inevitable uncertainties given by the
limited precision of measurement equipment and control quality the following
considerations have been taken into account.

4.1.1 Measurement Uncertainties


The values , and contain measurement uncertainties [mu] ∆ , ∆
and ∆ due to the utilized sensors. Therefore the deviation can be assumed to
be limited by uncertainties.

μ , − ≤∆ (6)

μ , − ≤∆ (7)

μ , − ≤∆ (8)

Here μ , , μ , and μ , , are the real (unknown) signal values. In the given case, the
measurement uncertainties are obtained by means of error propagation as
discussed in section 3.

4.1.2 Uncertainties of independen Variables


The independent values u and v are each adjusted by controllers [contr]. Their
possible deviation from the ideal set point values [sp] u and v depends only on
the quality of the control. The maximum possible deviation of u and v from the set
points will be denoted as ∆u and ∆v in the following. The deviation is then
limited to the control precision.

− ≤∆ (9)

272
∆ (10)

The complete possible deviation of the real values μ , and μ , from the set point
values is limited by the sum of the possible control variation and measurement
error. With the respective overall uncertainties being:

∆ ∆ ∆ (11)

∆ ∆ ∆ (12)

4.1.3 Uncertainties of the dependent Variable


Uncertainty of the dependent variable is dependent on the function g:

∆ ∆ ,∆ (13)

From the uncertainties of the independent values. Note that function depends on
function f and remains to be determined. The measurement of the dependent value
itself has also an uncertainty ∆ with overall uncertainty

∆ ∆ ∆ (14)

4.1.4 Comparison of two test stands


When applying this methodology, one test stand has to be defined as the reference.
The reference measurement is then performed on this test bench. The set points
and for the reference measurements could be defined as in sections 2.3.
The definition of auxiliary set points follows the central difference scheme in every
parameter as given in figure 6:

Figure 6: Auxiliary points for comparability study

The uncertainty of ∆ , that results from the independent variable is not


determined. The basic formula , defined in section 4.1 can be written as

 f (u , v )  f (u , v ) (15)
yiuv   u i   vi
u v  v i
u  u v uv vui
i i

273
The partial derivative of the function is not known. However, a reasonable
estimate of the slope at the auxiliary points can be gained by applying a suitable
numerical differentiation method to the measured data , and .

To get sensible results from a numerical differentiation method from measured


points, some demands on the performed measurements must be fulfilled. On the
one hand, the distance between two auxiliary set points should not be too small,
since then the estimation of the slope would be strongly influenced by
measurement uncertainties. On the other hand, the auxiliary set points should not
differ too much, since the estimation of the slope at the considered specific point
can obviously become inaccurate. With the given distance in figure 6 a good
compromise is reached. The partial derivation of measurement- and control
uncertainty for the operation point results in the following numerical values for the
evaluated measurement point:

Table 4: Results of partial derivation

Reference test Error Control


∆ overall ∆y
bench Propagation Uncertainty

u= 0.0701 [kg/s] 1.08% 0.23% 0.00092


v = 370.12 [m/s] 0.01% 0.19% 0.74024
y=1.907 [-] 0.21% 0.016

Relative Error 0.86%

Two test benches were evaluated following the procedure previously exposed. As
Table 4 reports, for the point considered in Figure 6, the reference test bench
provide an uncertainty ∆ of ±0.86%. The absolute interval is then 1.72%. This
could appear important but all different contributions have to be taken into account.
For the same point the second test bench gives a measured value with a 0.72%
offset from the reference. The procedure hence facilitates the comparability test.

5 REPEATABILITY

The discrepancies which could be considered in a comparability study between two


gas test stands could appear too important. Therefore readers could consider the
previous paragraphs to determine how many factors to take in account and to get
an idea of how discrepancies can be accounted for. To avoid any problem a possible
strategy could be to try to perform all the tests that are planned for a development
process on the same bench.

The test sequence adopted for the testing campaign in this project was specially
designed for evaluating the test bench repeatability in reaching exactly the same
measurement points. During all test sequence the bench moves between the nine
different points of the compressor map, controlling at the same time corrected
mass flow-rate and corrected compressor rotational speed. Ten random acquisitions
of the same turbocharger working point have been made. Discrepancies are
analysed for total-to-total pressure ratio and total-to-total compressor efficiency
using the assumption of a perfect gas.

Table 5 reports the results concerning the analysed central point of the compressor
map, the same as considered in section 3.2. These results are obtained by four
different tests performed in a time span of three years. In each column authors
report the 2σ interval obtained with following equation (16):

274
∑ ̅
2 =2 ̅
(16)

The number of points is cut to n, ̅ for the average value and for the different
measurements.

Table 5: 2σ interval on four different tests on a target point of


, = / and , = , /

Test , , , , ,

Continental AG test1 0,08% 0,12% 0,14% 0,27%


Continental AG test2 0,19% 0,18% 0,29% 0,23%
CRITT M2A test1 0,12% 0,19% 0,20% 0,25%
CRITT M2A test2 0,08% 0,23% 0,20% 0,16%
Maximum value 0,19% 0,23% 0,29% 0,27%

Looking at the results in Table 5, the target mass-flow rate and target corrected
speed have two dispersions which are more than half of the final values. For the
final calculated variables, not only the combination of these two factors has to be
taken in consideration. It is also due to slightly different energy equilibrium in the
machine in test. The different equilibrium may be due for example to different heat
transfer, test cell and flow enty conditions. As the points are addressed in a random
sequence, the influence due to system hysteresis enters the measurement.

A big component of this dispersion is definitely due to the accuracy of the control.
Steady state parameters were set to 0,3%. Both of the two test facilities gave
reproducible results fully in agreement with the stabilisation criteria. Again, as in
the case of the representation of the compressor iso-speed line, longer holding time
can reduce errors significantly at the price of longer measurement times.

6 IMPROVED TURBOCHARGER TESTING

The origins for differences in turbocharger compressor maps are explained in the
previous sections. The documentation of the quantitative experiments concerning
the differences can be summarized with their principal findings.

1. Users of compressor maps need to be aware of the possible errors in


compressor maps. The dispersion due to sensor errors is not negligible.
2. Measurements of a turbocharger compressor map differ due to differences
in measurement equipment and boundary conditions.
3. The comparison of two identical test benches with one set of sensors and
measurement sections delivers a comparability quality that is acceptable if
all the tolerances are taken in consideration.
4. Repeated measurements of one operating point shows only very small
deviation in the measurement.

Measurements of turbocharger characteristics always have to be made having in


mind the usage of the measurement data. A back to back comparison of two
different compressors on a single test bench with one set of instrumentation gives
precise information on the relative difference. It can be used as a precise tool in
turbocharger development and for comparison of two competitor products. To avoid
any problem a possible strategy could be to try to perform all the tests planed for a
development process on the same bench.

275
Comparison of measured turbocharger maps and data obtained by computational
fluid dynamics (CFD) may differ significantly due to assumed boundary conditions.
Typically CFD computations assume adiabatic walls. This leads to large differences
in compressor efficiencies.

When the turbocharger maps are used as input data for engine simulation or engine
control unit calibration, the difference in boundary conditions may lead to
significantly different performance maps. The difference in boundary conditions in
measurement compared to CFD or to engine operation can only be overcome by
additional test bench features such as special conditioning units for compressor air,
identical inlet and outlet piping compared to the vehicle or other dedicated
measurement strategies. Comparison of the characteristics of different
turbochargers measured on different test benches with different instrumentation
under different boundary conditions is not sensible.

7 SUMMARY AND OUTLOOK

Turbocharger maps differ when measurement taken place on a different test bed
under different boundary conditions. Comparability of two test benches is
acceptable when identical instrumentation and measurement sections are used.
Repeatability of single operating points on the same test bench with identical
instrumentation can be very good. This makes turbocharger test benches an
important development tool for turbocharger development if used for comparison
tests allowing quantification of for example improvements in efficiency, in flow
range or in surge behaviour. The comparison of test results from different test
benches with different instrumentation under different boundary conditions may
show significant differences. These results should be used with the appropriate
assumptions. Measurement campaigns for assigned development aims for TC’s
should be performed on one dedicated test bench due to significant differences
between test benches.

All that has been said concerns the compressor stage. The authors plan to analyse
the turbine characterisation. This would give a complete characterisation of the
possible derivations in turbocharger testing. The definition of clear performance
criteria for measurement could in fact be a useful tool for everyone involved in
turbocharger characteristics.

ACKNOWLEDGEMENTS

Authors would like to thank Continental AG for their assistance in this project.

REFERENCES

1. International, SAE. Turbocharger Gas Stand Test Code SAEJ1826.


Warrendale, PA : s.n., March 1995.
2. Mai, Dipl.-Ing. Holger. Parameterstudie zur Turbolader-
Kennfeldvermessung. s.l. : Abschlussbericht über ein VFI-Vorhaben.
3. ASME. PTC 10, VDI 2045.
4. Brun, K. and Kurz, R. Measurement Uncertainties Encountered During Gas
Turbine Driven Compressor Field Testing. J. Eng. Gas Turbines Power, 123.
2001, S. 62-69.
5. Mothar, H., Chesse, P. and Chalet, D. Describing Uncertainties Encountered
During Laboratory Turbocharger Compressor tests. SEM Technical Article.

276
6. Brun, K. and Nored, M. Guidline for Field Testing of Gas Turbine and
Centrifugal Compressor Performance. s.l. : Gas Machine Research Council and
Southern Research Institute, 2006.
7. Chapman, K. S. and N. Schultz. Guidelines for Testing large-Bore Engine
Turbochargers. Manhatten, Kansas : The National Gas Machinery Laboratory,
Kansas State University, 2003.
8. Guillou, E. Uncertainty and Measurement Sensitivity of Turbocharger
Compressor Gas Stands. SAE International SAE 2013-01-0925. 2013.
9. Ingenieure, Verein Deutscher. Thermodynamische Stoffwerte von feuchter
Luft und Verbrennungsgasen. VDI Richtlinie 4670. 2000.
10. Shaaban, S. Experimental investigation and extended simulation of
turbocharger non-adiabatic performance. Hannover : Dissertation, Universität
Hannover, 2004.
11. M. Cormerais, P. Chesse and J.F. Hetet. Turbocharger Heat Transfer
Modeling Under Steady and Transient Conditions. Int. J. of Thermodynamics.
2009, S. 193-202.
12. Fesich, M. and Casey,M. The Efficiency of Turbocharger Compressors with
Diabatic Flows. J. Eng. Gas Turbines Power 132. 2010.

277
Containment simulation and validation of
turbocharger housing design

J M Ramamoorthy, S S Parikh, S Pandian, P S Kasthuri Rangan


Turbo Energy Limited, India

ABSTRACT

One of the basic requirements for turbocharger housing design validation is to


prevent the wheel fragments from penetrating the housing during wheel burst. This
paper describes FE simulation methodology development and validation of
containment test in order to obtain the optimum housing design at the early design
stage, ensuring “Design right first time”. This high speed dynamic simulation
methodology involves AWMHS method and explicit FEA. This simulation demands
elastic, plastic and fracture characteristics, which is defined by elastic modulus,
hardening parameters and ductile damage material model respectively. This
methodology is applied for compressor housings and also for high temperature
turbine housings. Simulation results are found to be in good agreement with actual
containment (rig) tests.

Keywords: Containment simulation, wheel burst, area weighted means hoop stress
method, explicit FEA, turbocharger housings, ductile damage.

ABBREVIATIONS

FEA – Finite element analysis


AWMHS – Area Weighted Mean Hoops Stress
UTS – Ultimate tensile strength
KE – Kinetic energy
IE – Internal energy
TMF – Thermo Mechanical Fatigue
ms – milliseconds

NOTATION

Σaie = sum of element areas


σihoop = element hoop stress

1. INTRODUCTION

The containment simulation along with TMF simulation helps to verify the optimum
housing Design. Housing Containment tests verify the operational limits of a
turbocharger's housings to contain wheel fragments in the event of a wheel hub
burst during operation. This paper describes the containment simulation and
validation of both compressor and turbine housings. For a containment simulation,
a wheel hub burst is required. The first part of this paper describes the wheel burst

_________________________
© Turbo Energy Limited, 2014
281
simulation and later the containment simulation and validation. The wheel burst
simulation initially involves the wheel burst speed estimation. This estimated burst
speed is used in explicit burst simulation and a material model is developed to
match the burst speed. In this simulation the wheel is naturally allowed to burst
similar to an actual rig test. Since the wheels are designed with a good speed
margin, a weakened wheel is considered for containment simulation. The following
chart (Figure 1) describes the workflow of the containment simulation & validation.
The various phases are explained in detail in the following topics.

Wheel burst simulation


Step 1: AWMHS method
Step 2: Wheel burst explicit simulation

Housing Containment Simulation

Results Comparison with physical testing

Figure 1: Workflow for containment simulation and validation

2. WHEEL BURST SIMULATION

A wheel bursts when the centrifugal force exceeds its internal binding forces. The
first part of containment simulation involves ‘wheel-alone’ burst simulation. The
wheel burst is initially estimated using AWMHS method. As the wheels are designed
for a much higher speed margin, it is required to weaken the wheel. Using AWMHS
method, various iterations were carried out to arrive at the final design of the wheel
so that a wheel hub burst occurs at the required speed. This estimated burst speed
is used for explicit simulation. Initially, wheel alone is considered (without the
housing) to quickly evaluate the wheel weakening dimensions and the material
model for wheel burst explicit simulation.

2.1. AWMHS method


Area Weighted Mean Hoops Stress method is
a standard method used to estimate the burst
speed of rotors. Here a sector model (Figure
2) of the wheel component is considered. In
wheels, material is removed at the bore to get
the hub burst effect. If a wheel is considered
without material removal, it may lead to over
design of the housings. Several iterations are
carried out to obtain the optimum material
removal dimensions for the required burst
speed. In the case of wheels under rotating
conditions hoop stress is of major concern.
Hence the maximum hoop stress is used to
estimate the burst margin.

The hoop stress is estimated using FE Analysis Figure 2: Sector


and burst speed margin is estimated using the model of wheel
empirical relationship mentioned below.

282
∑ 𝑎𝑒𝑖 ∗ 𝜎ℎ𝑜𝑜𝑝
𝑖
𝐴𝑊𝑀𝐻𝑆 =
∑ 𝑎𝑒𝑖

𝐵𝑢𝑟𝑠𝑡 𝑆𝑝𝑒𝑒𝑑 𝑀𝑎𝑟𝑔𝑖𝑛 = √(𝑈𝑇𝑆⁄𝐴𝑊𝑀𝐻𝑆)

The workflow to arrive at the weakened wheel design for containment simulation
from AWMHS method is shown in Figure 3.

Cyclic sector wheel model

FE structural analysis Modify material


removal
dimensions
Calculation of burst speed

Check if burst No
speed =
required speed

Yes

Conclude

Figure 3: Workflow to obtain material removal dimensions using AWMHS


method

2.2. Wheel burst explicit simulation


The turbocharger wheels rotate at high speeds and the wheel burst occurs over a
short duration of time. This high speed and short time duration simulation can be
performed only in explicit analysis. The uniqueness of this project is that the wheel
is allowed to fail by itself in simulation and that no restriction is imposed on the no
of failed segments.

The burst speed and weakening dimensions obtained from AWMHS method are
used herein as input. Since the failure involves fracture of the component, elastic,
plastic and also the fracture characteristics for the material are to be defined.
Different regions of the stress-strain curve specific to this project are shown in
Figure 4[1] [2]. The elastic region involves the elastic modulus, the plastic region
involves plastic strain and the fracture region requires the damage criteria which
include the failure strain. For turbine wheels higher temperature material properties
are used. In this project based on the burst speed obtained from AWMHS method, a
material model for explicit wheel burst simulation is chosen.

283
Plastic

Elastic
Fracture

ε
Figure 4: Different regions of stress-strain curve

There are various damage models available in commercial software [3] [6]. The
damage criteria in the material model were developed iteratively to match with the
burst speed obtained from AWMHS method, thereby reducing the time and cost for
extensive material testing. The wheels have an elongation of greater than 5%, thus
a model to capture its ductile nature is used. For turbine wheels high temperature
material models are used. The damage model is very critical as it defines the failure
of the wheel, similar to failure of wheel in actual operating condition. The workflow
to obtain damage criteria for explicit wheel burst simulation is shown in Figure 5.

Input: Weakened geometry & burst speed from AWMHS

Damage criteria

Explicit burst simulation Modify Damage


criteria

Burst @ Required No
speed

Yes

Yes
Burst @ 1% <
Required speed

No

Conclude

Figure 5: Workflow to obtain Damage Criteria

284
Thus, a suitable damage criterion is developed using the AWMHS method. Figure 6
shows compressor and turbine wheel burst in explicit simulation.

(a) (b)

Figure 6: Wheel burst explicit simulation (a) compressor wheel


(b) turbine wheel

3. HOUSING CONTAINMENT SIMULATION

Containment test along with TMF test decide the design criteria for the housings.
They also act as a safety standard for the turbocharger. The criteria for a
containment test are that, in the event of a wheel burst, no fragments of the
wheels must penetrate the housings [5]. Demonstration of this requirement is an
expensive and time-consuming process. Hence an upfront containment simulation,
in the early design stage, would lead to substantial time and cost savings. By
following this approach an optimum and robust design of the housing can be
arrived at.

For containment simulation, both the wheel and housing are considered which
makes the problem larger in size and more complex. Material models for the
housings are developed, similar to that of wheels, to visualize the actual damage on
the housings in simulation. A strain rate dependent material property is required as
the impact velocities of wheel fragments on housing are not constant. Additionally,
for turbine housings, temperature dependent material models are used as the
temperature on the turbine housing is non-uniform.

3.1. Compressor containment simulation


In this simulation compressor wheel, housing and central housing cover are
considered. For compressor side temperature effects are not considered. The
damage on the compressor housing can be visualized in Figure 7. These results are
for a time period of 0.3 ms.

0.1 ms 0.14 ms 0.16 ms

0.18 ms 0.24 ms 0.3 ms

Figure 7: Compressor housing containment

285
The total duration of the explicit simulation must ensure considerable reduction of
kinetic energy of the wheel [4]. This can be confirmed by plotting the energy
graphs as shown in Figure 8.

KE of compressor wheel = IE of Compressor wheel + IE of Compressor


housing + IE of Central Housing Cover + Friction Losses
Figure 8: Energy graph for compressor containment simulation

3.2. Turbine containment simulation


In this simulation turbine wheel and housing are considered. To include the
temperature effects on turbine side a steady state thermal analysis is performed for
the housing and the temperature distribution is mapped on to it in containment
simulation. The damage on the turbine housing can be visualized in Figure 9. These
results are for a time period of 0.6 ms.

0.1 ms 0.2 ms 0.3 ms

0.4 ms 0.5 ms 0.6 ms

Figure 9: Turbine housing containment

286
The energy graph for turbine side is shown in Figure 10.

Normalized Internal Energy


KE of Turbine wheel = IE of Turbine wheel + IE of Turbine housing + Friction Losses
Figure 10: Energy graph for Turbine containment simulation

4. RESULT COMPARISON WITH RIG TEST

The simulation results are compared with several rig tests and validated. In both
compressor & turbine side the wheel burst speed Simulation and rig test results
towards the Burst speeds is seen to be less than 2%. This validates our material
model developed using AWMHS method and wheel burst explicit simulation.

The Figure 11 shows the results comparison of compressor containment rig test and
simulation. A good agreement is seen between Simulation and Tests on the pattern
of ‘two-parts burst’ at same location observed in both rig test and simulation. The
housing contains the wheel burst in both rig test and simulation and no penetration
observed.

Figure 11: Compressor containment result comparison

287
The Figure 12 shows the results comparison of turbine containment rig test and
simulation. Similar pattern of hub burst at same location is observed in both rig test
and simulation. The housing contains the wheel burst in both rig test and simulation
and no penetration observed. Thus the simulation results are validated with that of
rig tests.

Figure 12: Turbine containment result comparison

5. CONCLUSION

Today the OEM market demands lighter turbochargers with high durability. The
housings contribute to more than 60% of the total weight of the turbocharger.
Generally the thickness of housings is decided with a high factor of safety, to
prevent containment failure in rig test. This leads to bulky housings and also in
case of turbine housings, it reduces the TMF life of the same. This simulation
methodology along with TMF simulation (for turbine housings) helps to design
housings with optimum thickness in the early design stage. It is also helpful to
identify unwanted material addition in the housing design. Substantial material
savings were observed by implementing this simulation methodology in housing
design., Optimum thickness of the housings is also likely to improve the life of the
housings considerably, which fulfils the need of the hour.

6. REFERENCES

1. Radioss Theory Manual 11.0 version, Copyright © 1990-2011. Altair


Engineering, Inc.
2. Analysis user’s manual, Abaqus 6.11 Documentation, © Dassault Systèmes,
2011
3. Abaqus explicit advanced Training manual, © Dassault Systèmes Simulia
Corp., 2012
4. Rajeev Jain, “Prediction of Transient Loads and Perforation of Engine Casting
During Blade-Off Event Fan Rotor Assembly”, Proceedings of the IMPLAST 2010
Conference October 12-14 2010 Providence, Rhode Island USA © 2010 Society
for Experimental Mechanics, Inc.
5. White paper No.2, Burst & containment : Ensuring Turbocharger safety, Garret
by Honeywell
6. Unai Hermosilla, José L. Alcaraz, Angel M. Aja, “Blade Impact Simulation
Against Turbine Casings”, 2004 Abaqus users conference

288
Engine crank angle resolved turbocharger
turbine performance measurements by
contactless shaft torque detection
B Lüddecke1, D Filsinger1, M Bargende2
1
IHI Charging Systems International GmbH, Engineering Division, Germany
2
Universität Stuttgart, Institut für Verbrennungsmotoren und Kraftfahrwesen,
Germany

ABSTRACT

The trend to use advanced simulation tools for engine performance prediction is
continuing and even emphasized due to shortening of development cycles. The
highly accurate prediction of steady and transient engine behaviour becomes
increasingly important. Complete drive cycle simulations (e.g. NEDC) help to assess
turbocharged engine performance at very early stages of complex engine and
vehicle development projects.

The turbocharger has recently developed away from an auxiliary part towards an
integral component of the internal combustion engine. Hence, the accuracy of
turbine and compressor maps becomes more and more relevant to achieve reliable
simulation results and predictions. High quality turbocharger performance data are
necessary over a wide range of operation conditions as input for engine simulation
programs.

Especially modelling the turbine stage efficiency for engine-like operating conditions
(pulsed flow) still is under research. In this context, many researchers raised the
question about unsteady effects within the turbine stage and whether the stage,
the volute and/or the wheel behave quasi-steadily or have to be considered as
unsteady devices.

In the paper at hand, a contactless shaft torque detection technique - that has been
integrated into an automotive turbocharger - is presented. It is possible to measure
the turbine shaft torque with high accuracy and time resolution. A turbocharger
equipped with the detection system has been adapted to a modern four cylinder
gasoline engine with direct injection. Engine cycle resolved torque data has been
gathered in order to assess the crank angle resolved, on-engine performance of the
turbine stage.

The available measurements represent an excellent basis for advancements in the


modelling and simulation of turbocharger turbine stages with engine simulation
tools. Furthermore, the results give a clear indication about the significance and
magnitude of unsteady effects within the turbine stage under pulsed flow
conditions.

_______________________________________
© The author(s) and/or their employer(s), 2014
301
NOMENCLATURE

A Area [m^2]
∆t time difference
k kilo, 1,000
T torque [Nm]
mass flow rate [kg/s]
n rotational speed [rpm]
P power [W]
PI pressure ratio [-]
p pressure [Pa]
T temperature ([K] for calculations,
[°C] for tables)
u blade tip speed (turbine or compressor) [m/s]
. blade speed ratio [-]

Abbreviations
BDC Bottom Dead Centre
CA Crank Angle
CFD Computational Fluid Dynamics
DFT Discrete Fourier Transformation
ECU Engine Control Unit
MFP Mass Flow Parameter [m s K^0.5]
NEDC New European Driving Cycle
TDC Top Dead Centre
WOT Wide Open Throttle

Greek Symbols
∆ difference [-]
 number pi
ɵ moment of inertia [kg m^2]
Engine cycle duration (720 °CA)

Indices/Subscripts
1 compressor stage inlet
2 compressor stage outlet
3 turbine stage inlet
3.5 turbine impeller inlet
4 turbine stage exit
aero corresponding to aerodynamics
air / a values related to air
C compressor (stage)
Cycle related to the full engine cycle
coolant state of coolant
corr corrected
eff/effective relevant/operative part
fric/friction related to friction
gas / g values related to burnt gas
GE related to gas exchange
i crank angle index during engine cycle; i ∈ {0-719}
inertia related to rotor inertia
is isentropic
journal related to journal bearing
lubricant state of lubricant
meas measured value
r radial

302
rotor corresponding to the rotor
s static
ss static-to-static
T turbine (stage)
t total
tt total-to-total
ts total-to-static

1 INTRODUCTION

In [1] a novel, contactless turbine shaft torque detection system, based on inverse
magnetostriction was introduced. It was employed for steady state performance
measurements of a mixed flow turbine on a hot gas stand. One main advantage of
this approach is its robustness towards occurring heat flows into the compressor
stage during testing (compare [2]). This enables a reasonable turbine performance
measurement especially at low pressure ratios, mass flows and rotor speeds.
Furthermore, the relative influence of bearing losses on turbine performance
measurement is reduced significantly.

In the present work, the detection system is employed for unsteady torque
measurements under pulsed flow conditions on an internal combustion engine. The
standard serial production turbocharger of a modern four cylinder gasoline engine
with direct injection has been equipped with the shaft torque detection. Engine
cycle resolved torque data has been gathered in order to assess the crank angle
resolved, on-engine performance of the turbine stage.

Especially the modelling of turbine stage efficiency for engine-like operating


conditions (pulsed flow) still is under research. Of course, turbine map extrapolation
methods are state of the art, however they still suffer from non-perfect input (test)
data. In this context many researchers raised the question about unsteady effects
within the turbine stage and whether the stage, the volute and/or the wheel behave
quasi-steadily or have to be treated unsteadily.

The available measurements represent an excellent basis for advancements in the


modelling and simulation of turbocharger turbine stages with engine simulation
tools. Furthermore, the results give a clear indication about the significance and
magnitude of unsteady effects within the turbine stage under pulsed flow
conditions.

2 SENSITIVITY STUDIES

The influences of varying distances between the primary sensor (magnetically


coded shaft) and the secondary sensor (coils + electronics) were investigated to
prove the robustness of the system and its suitability for the conducted
investigations.

In Figure 1 the rotor and sensor of the torque sensing system are shown in the
lower right corner. Furthermore, the results of a detailed sensitivity study regarding
sensor/rotor relative displacement are presented. The direction of the investigated
movement is indicated by the white arrows. The dashed grey lines indicate the
maximum potentially possible movement that could theoretically occur considering
geometrical clearances and tolerances of the bearing design. However, the
occurring displacements and movements of the rotor under real conditions are
usually significantly smaller due to oil film thickness and the damping effects of the
oil film onto shaft motion.

303
Figure 1: System configuration and results of sensitivity studies regarding
rotor/sensor relative displacement

Hence the influence of the rotor motion in any direction becomes quite negligible
within the range that the rotor can potentially move during operation - even under
worst case scenario assumptions.

Furthermore - as the differential signal of the coils is used to record changes in the
outer stray magnetic field of the primary sensor - the system is very robust against
parasitic magnetics fields (like geomagnetic fields or other sources). The change of
an external magnetic field will influence the absolute signal level of every coil, while
the difference between them remains constant, as described in [1]. This robustness
becomes especially important at engine test stand, where many residual sources of
electromagnetic fields exist, e.g. the ignition system of the engine.

3 SHAFT TORQUE AND BEARING FRICTION

To better understand the relevance and interpretation of the measured torque data,
Figure 2 illustrates how the torque is transferred from the turbine to the
compressor through the shaft for steady state conditions (no rotor acceleration or
deceleration). The minor losses by seal ring friction as well as the ventilation losses
on surfaces are not considered.

During expansion, the gas flow causes pressure forces on the blade surfaces and
hence transfers torque to the impeller. This is referred to as the aerodynamically
available torque. The torque sensor is located between the journal bearings of the
turbocharger. Consequently, the measured value is lower than the actual
aerodynamic torque, as the turbine side journal bearing causes a torque loss. The

304
compressor side journal bearing as well as the thrust bearing create additional
losses. The remaining torque is available to the compressor wheel for work input
into the fluid, resulting in pressure rise. It is important to draw the attention to
these relations in detail, as below in chapter 7 the unsteady case under pulsed flow
(with alternating rotor speeds) is considered, which significantly differs from the
steady case.

Figure 2: Schematic of torque transfer from turbine to compressor and


occurring losses for steady state condition

4 HOT GAS STAND MEASUREMENTS

Prior to engine testing, a detailed experimental study on the hot gas stand was
elaborated under steady state boundary conditions.

The turbine inlet temperature T3 was varied in six steps (40, 200, 400, 600, 800,
1000°C) and for every case several speed lines have been recorded. The
temperature of coolant and lubrication were set to 30°C and 35°C, respectively, to
keep the thermal loading for the measuring system low. This leads to higher values
of bearing friction, of course. However, as these losses are modelled and corrected
based on available bearing friction measurement data, this has no implication on
the relevance and accuracy of the data. The supply pressure of the lubrication was
set to a constant value of 4 bars absolute.

In Figure 3 an excerpt of the measured turbine torque for T3=600°C is plotted


against the total to static pressure ratio. The red curves with squared symbols
represent the measured “raw” torque. As mentioned above, the torque is measured
between the two radial bearings, Figure 2. Hence, to assess the corresponding
aerodynamic torque – that is basically the one of interest – the friction torque of
one radial bearing has to be added to the measured torque, eq. (1).

, = + (1)

Based on available experimental data, a polynomial function (of the rotor speed)
has been employed to account for the journal friction torque. The result is given by
the black curves with white circles in Figure 3. Obviously, the correction causes an
offset from the measured value, while the slope remains constant.

305
Figure 3: Measured and calculated torque trend vs. pressure ratio

Additionally, torque data from CFD results are shown as solid blue curves.
Obviously, for every speed line, the measured data show a linear dependency of
turbine rotor torque versus stage pressure ratio within the measured range. This is
to be expected and has also been reported by several researchers, compare e.g.
[3], [4], [5]. The CFD data also show a fairly linear trend but as the computed
range is much wider than the measured one, smooth changes in the slope can be
seen, if very low and very high pressure ratios are compared. The reason for this is
believed to be related to the strongly changing flow regimes and mach numbers, if
for constant rotor speed the pressure ratio is varied in a wide range. However, it
can be stated that measured, friction corrected and computed data match well.

Furthermore, the measured data and the CFD results in particular show that the
turbine rotor torque would become zero far before the pressure ratio equals one.
This is related to the centrifugal force field of the rotor (compare [6]) and is
important for the interpretation of the engine results shown later.

The available recorded data for different rotor speeds as well as turbine inlet
temperatures have been employed to gain a wide map of the turbine stage in terms
of blade speed ratio and pressure ratio, Figure 4, top. The data points of the
measured speed lines are marked with different symbols, corresponding to the
turbine inlet temperature. In the middle of Figure 4, the corresponding turbine
stage efficiency – based on measured shaft torque - is plotted. The white circles
represent exactly the data points from the plot at the top, which have been used to
interpolate a contour of turbine stage efficiency. The iso-lines have a distance of
one per cent efficiency. The plot shows the efficiency after journal bearing friction
correction (eq. (1)) as demonstrated in Figure 3 and hence represents aerodynamic
efficiency. At Figure 4, bottom, a contour plot based on CFD data is shown. The
white circles accordingly mark the location of the numerical results that the plot is
based on. The comparison between the (bearing friction corrected) measured data
and CFD results shows good agreement. The absolute level of efficiency though is
still highest for the computed data.

306
Figure 4: Turbine stage efficiency contour plots based on measured torque
and radial bearing friction corrected (top) and comparison with CFD data
(bottom)

This can be attributed to miscellaneous minor and partly superimposed influences.


These are for instance: geometry simplifications and turbulence modelling for the
numerical results; deviation of real geometry from nominal shape as well as the
inevitable uncertainties within an experimental data set. It is also conceivable that
the radial bearing friction model does not perfectly represent the occurring bearing
friction for all hot gas temperatures and all speed lines as the model is based on a
limited set of experimental data.

However, the data shown above give a very high level of confidence into the
presented method for steady state conditions.

5 ENGINE MEASUREMENT SETUP AND INVESTIGATED OPERATING


POINTS

Based on the serial production turbocharger specification of the test engine, a


prototype turbocharger was built up, incorporating the described contactless shaft
torque detection system. The standard serial production turbocharger was then
replaced by the prototype. Compressor and turbine specification were not changed.
Hence, regarding engine-turbocharger matching, the engine can be operated safely
within its full operating range.

307
Furthermore, as a partly programmable ECU was available, certain conditions of
operating points could be manipulated independently, e.g. valve timing, which was
used to arrange special parameter variations for detailed investigations of the
interaction between the combustion engine and the charging system.

For the time resolved turbocharger turbine shaft torque measurements, a recording
frequency of over 100 kHz was used. By DFT, the spectrum of the measured signal
was investigated regarding its bandwidth and maximum relevant frequency. The
raw data were then filtered appropriately and resampled to crank angle (0.1 °CA
resolution). This workflow assures high quality crank angle resolved data and
reasonable final file sizes. For the shown steady operating points, about 200
consecutive engine cycles were recorded, post-processed, filtered and then an
average engine cycle was calculated.

In Table 1 an overview of the engine operating points presented in this paper is


given. All points have been recorded at an engine speed of 1250 rpm. Four
stationary stable operation points are presented, where only the cam phasing for
intake and exhaust valves was varied to adjust engine load, while the throttle was
kept at WOT conditions. The load is given as percentage, related to the full load
torque of the serial production engine application at 1250 rpm.

Table 1: Engine operating points

No. Speed Pedal Load Cam_int Cam_exh Lambda


-- rpm % % °CRK °CRK --
116_00 1250 WOT / 100% 63.6 110 -110 0.99
116_01 1250 WOT / 100% 78.5 85 -85 1
116_02 1250 WOT / 100% 91.6 82 -80 1.09
116_03 1250 WOT / 100% 112.9 82 -72 1.18

It is obvious how significantly the valve overlap timing influences engine torque.
Any engine load variation is solely caused by intake and exhaust valve timing
variation and thus strongly related to the so called “scavenging” mechanism that
(additionally to the conventional turbocharging) boosts the engine. By this well-
known operating strategy ([7], [8], [9]), engine torque can be almost doubled.
Even the serial application full load torque can be exceeded, what is believed to be
related to two main reasons:
 The waste gate had been mechanically blocked to minimize leakage – a
state that can surely not be reached in serial production engine under
pulsed hot gas conditions. The increased mass flow over the turbine wheel
leads to increased turbine shaft power and hence compressor power.
 The OEM follows a conservative scavenging strategy to assure durability of
the engine as well as the definite avoidance of pre-ignitions under any
circumstances in the field.

For the investigated case the engine had been well conditioned and was operated
under surveillance of a monitoring and control system, hence the limitations
mentioned above can be exceeded. All four stationary points were run at thermal
persistency and close to engine knock limit.

For the operating points listed in Table 1, a combined combustion and gas
exchange analysis was conducted exemplarily for cylinder four, employing the
commercially available software Tiger [10]. The corresponding results are plotted in
Figure 5. Although – due to the complex flow regimes – scavenging operation is
hard to analyse accurately by a zero-dimensional or limited one-dimensional code,
the results clearly indicate the fraction of scavenged mass. The valve effective

308
areas are given by the dashed black lines. The intake and exhaust channel
pressures of cylinder four are drawn as solid blue and red curves. The
corresponding calculated intake and exhaust mass flows are drawn as dotted blue
and red curves.

Figure 5: Results of gas exchange analysis

Obviously, the shift of the valve lift curves causes two changes: First of all, it
enables scavenging at all, as intake and exhaust valves are allowed to be opened
simultaneously with a certain overlap. Secondly it also moves the relative position
of pressure pulses and valve overlap in the desired direction. For scavenging,
cylinder inlet pressure (~compressor outlet pressure) has to be higher than cylinder
outlet pressure (~turbine inlet pressure).

At the end of the scavenging process (close to exhaust valve closing), a negative
mass flow can be observed. This is caused by the absolute length of the (exhaust)
valve timing as the opening event of the next cylinder’s exhaust valve pushes back
some mass flow while the exhaust valve of the actual cylinder is still opened. In a
four cylinder engine, a shortened and/or variable exhaust timing length can help to
avoid this, as shown in [7], [8]. This effect proves the imperfect flow separation of
the exhaust channels in especially four-cylinder engines, where the exhaust valve
opening duration is longer than the distance between two exhaust strokes. This is
also one main driver for twin scroll or double scroll concepts, where the flow
separation is realized within the turbine housing. An alternative is a variable
exhaust valve opening event, realizing this flow separation within the cylinder head.
However, a short exhaust valve opening duration can only be used effectively for
the low end torque area, as for high speeds and loads the longer duration is needed
(together with gas dynamic effects) to realize the exchange of the cylinder gas
mass in a very limited period of time.

309
6 CYCLE AVERAGED ENGINE AND TURBOCHARGER OPERATION POINTS
AND TOTAL VALUE CORRECTION

As mentioned above, the waste gate was mechanically blocked to assure that the
complete gas mass flow is fed to the turbine wheel. Moreover, the actual position of
the waste gate system was controlled by an optical distance sensor to prove the
blocking forces were high enough to resist the opening forces caused by the
exhaust gas pressure pulses acting on the waste gate surface.

The avoidance of any significant leakages (internally or to the test cell


environment) is mandatory to achieve reliable power balances, allowing for
sensibility crosschecks of the measured data.

In Table 2, the turbocharger conditions corresponding to the engine operating


points of Table 1 are listed.

Table 2: Turbocharger operation points corresponding to Table 1

N_red T_ p_ T_ T_3 Wastegate


No. PI_C PI_T
Turbine lubricant lubricant coolant AVG position
rpm K^-0.5 °C bar abs °C -- -- °C --
116_00 1970 22 3.9 29 1.18 1.15 649
closed /
116_01 2454 23 3.9 29 1.30 1.23 689
mech.
116_02 2996 24 3.9 29 1.49 1.34 722
blocked
116_03 3773 25 3.8 29 1.85 1.57 758

The turbocharger speed develops as expected for increased engine load, as the
average turbine and compressor pressure ratios increase.

In Figure 6 an overview for 116_00 to 116_03 (Table 1 & Table 2) is given. Distinct
colours are used for each operating point within the graphs (116_00=blue,
116_01=green, 116_02=red, 116_03=black/white).

Figure 6: Graphical overview of investigated operating points: compressor


map (left) and turbine map (right)

In the left graph the compressor map – obtained from hot gas stand tests - is
shown and the four compressor operating points are depicted. The four single data
points are based on averaged measured pressures (fast sensors, signal time
averaged), temperatures (slow sensor principle, no averaging) and turbocharger
rotor speeds (fast, time averaged signal) from engine test stand. The same is true
for the right graph, showing the steady state hot gas stand turbine map
(T3=600°C) and averaged turbine data during engine operation.

Although measurement positions, due to packaging, piping and bends are different
on engine compared to the gas stand measurement, a good match is obvious.

310
7 ENGINE CYCLE RESOLVED TURBOCHARGER AND ENGINE
MEASUREMENT RESULTS

The phase shift between turbine stage inlet, impeller inlet and turbine stage outlet
is a frequently discussed topic when analysing time or crank angle resolved turbine
performance data. Many authors raised the question how to correctly define and
calculate a stage (and/or wheel) pressure ratio that is comparable to steady hot gas
stand results, [11], [14], [5]. The phase shift is depending on the volute size and
design, the pressure tapping positions as well as the engine operating point with its
rapidly changing flow and temperature regimes during the engine cycle.

Hence it was decided to conduct a sensitivity study by applying phase shifts onto
the measured inlet and outlet pressures. The results are presented at the end of
this paper in chapter 8 after all other relevant phenomena have been discussed and
corrections have been introduced.

It is to note that all cycle resolved pressures and pressure ratios from engine
measurements are static and static-to-static, respectively. No correction is
computed as it has been done for the presented steady hot gas stand maps. For the
compressor, due to the complex packaging situation and especially because of the
unsteady flow conditions (consecutive filling of the cylinders), there is no
straightforward method available to extract the total values from the available
measured data. The potential error introduced by a possibly not fully valid
correction procedure could be higher than the potential gain in accuracy. Anyway,
the difference between static and total values amounts to only few thousandth of a
bar. The same is true for the turbine side, with the difference that at the turbine
inlet the variations/pulsations are much larger and the temperature- and velocity
level is higher. Especially during blow down phase (shortly after exhaust valve
opening) the flow velocities and temperatures as well as their gradients raise
significantly.

Figure 7: Crank angle resolved pressures, turbine stage pressure ratio


and turbine rotor torque

311
In Figure 7 the crank angle resolved pressure data for turbine inlet (red solid line)
and outlet (blue solid line), compressor outlet (green solid line) as well as the stage
pressure ratio (black dashed line) and the rotor torque (black solid line) are shown
for all four operating points described in Table 1 and Table 2.

It is obvious how the amplitudes and gradients of all plotted values increase with
engine load and that due to valve timing variation also the location of the pulses
changes. A clear dependency of measured turbine torque and stage pressure ratio
can be seen, while the comparison with turbine inlet pressure pulses exhibits a shift
in time.

In Figure 8 the measured, unsteady turbine rotor torque is plotted versus turbine
stage static-to-static pressure ratio. The values have already been corrected for the
journal bearing friction torque. Additionally, the steady torque data (also bearing
friction corrected) obtained at hot gas stand for T3=600°C (grey squares) and
T3=800°C (grey triangles), are shown. The corresponding reduced turbine speeds
are given as well (grey numbers for steady state, coloured numbers for unsteady
results). As the measured, average turbine inlet temperature during engine
operation is in between these two values, this is the best comparison that can be
made between pulsed and steady flow conditions. As mentioned by other authors
([3], [5]) and as expected, due to filling and emptying effects of especially the
volute, no straight line with constant slope, but an orbit is observed. The unsteady
data is cycling in the vicinity of the steady state measurement data. Compared to
the usually presented MFP values (comparison of steady and unsteady case, see
[3], [11], [12]), the deviations as well as the observed orbit are quite small,
however.

Figure 8: Measured unsteady shaft torque, energy averaged values and


steady-state torque data obtained at hot gas stand; all values plotted
vs. turbine stage pressure ratio

As a third group of values, the energy averaged torque and pressure ratio are
plotted as coloured asterisk symbols. In this context, the employed averaging
procedure is very important in terms of achieving results that are comparable to
the steady case. The averaged values shall represent a process that transfers the

312
same average power to the compressor stage (and the bearing system) as a steady
state operation on hot gas stand would exhibit.

First of all, the overall amount of energy that is transferred from the turbine wheel
over the shaft to the compressor stage (and the bearing system) during one engine
cycle is calculated, eq. (2).

,
= 2 ∆ = ∆ (2)
60

The value ∆ is not quite constant during engine cycle, as small variations in
engine speed can be observed. Please note that the journal bearing friction
corrected torque values are employed.

Based on the cycle energy and the engine cycle duration (720°CA), the average
turbine shaft power can be computed, eq. (3).

∑ ∆ (3)
, = =
∑ ∆

Together with the cycle average turbine shaft speed, the energy equivalent, cycle
averaged shaft torque is calculated, eqs. (4) & (5).

, , 2 , ∆
, , =2 = (4)
60 60 ∑ ∆

,
, = (5)
, ,

Finally, the energy equivalent, relevant cycle average turbine stage pressure ratio
can be determined by eq. (6).

∑ , ∆
= (6)
, ,
∑ ∆

It is obvious that the cycle-average values of torque and pressure ratio are close to
the steady state speed lines for T3=600°C and T3=800°C. More precisely, they are
almost in between or in extension of the steady measurements with comparable
reduced speed.

Some minor deviations exist, hence the asterisk symbols do not match exactly with
the (extended, if necessary) reduced speed lines at the energy averaged pressure
ratio. This could be related to unsteady turbine operation. However, it seems more
likely that the “slow” temperature measurement method and its positioning are of
major importance in this context. As a matter of principle the measured
temperatures behave more like a time averaged quantity and hence do not
represent energy equivalent values.

One also observes a difference in the slopes for unsteady and steady state
measurements in Figure 8. For the steady state results it is clear that the torque
will become zero before the pressure ratio equals unity due to the centrifugal force
effects of the rotor, [6]. This characteristic is not exhibited for the measured
unsteady torque data. However, this is an aerodynamic matter of fact that should
not change for radial turbomachinery under any condition.

The reason for this unexpected slope is the transient speed change of the rotor.
Figure 9 shows the measured stage pressure ratios as already shown in Figure 7,

313
amended by the measured rotor speeds as well as the torque values that are based
on the rotor speed change and the rotor inertia, according to eq. (7).

= ∙ 60 ∙ 2 (7)
, ,

It is obvious that the rotor speed change corresponds to a significant amount of


torque (in the order of magnitude of the measured signal) and hence has relevant
influence on the measured shaft torque and certainly cannot be neglected.

Figure 9: Measured pressure ratio, rotor torque and rotor speed amended
by the calculated torque based on speed variation and inertia

The crank angle resolved characteristics of the inertia based torque look similar to
the contactlessly measured values with the difference that a change in sign occurs,
related to rotor acceleration/deceleration. The practical interpretation is that during
the acceleration phase, the rotor “swallows” some of the aerodynamic torque
(hence it can’t be measured) and releases it again during deceleration (hence more
than the aerodynamic torque is measured).

For calculation of this torque, only the inertia of the turbine wheel and that portion
of the shaft until the measuring position has to be taken into account, not the full
inertia of the rotor. Thus it is addressed with the index “effective”. The described
behaviour is illustrated in Figure 10, for acceleration as well as for deceleration. The
correct value of (aerodynamic) torque can be determined by superimposing both
torques as well as the bearing friction correction for the radial journal bearing, eq.
(8).

= + + ∙ 60 ∙ 2 (8)
, ,

314
An equivalent approach has been presented in [13], where the turbine stage
performance has been investigated under pulsating conditions by the use of a
special test rig and instantaneous measurements.

Figure 10: Rotor shaft torque distribution for unsteady case: deceleration,
acceleration and offset influence onto shaft torque measurements

In Figure 11 the results before and after inertia torque correction are shown again
versus static-to-static pressure ratio. In the upper left graph the measured torque
during steady and unsteady operation for case 116_03 is shown (compare Figure
8). In the upper right corner, the calculated torque, based on eq. (7) is given. The
lower graph shows the superimposed curves of the two upper graphs. While
measured torque shows invariably positive values, the inertia based torque can be
positive or negative, according to acceleration and deceleration, respectively. The
superimposed values look much more reasonable than before correction. If the
slope of the resulting curve is compared with the steady values, a very good
agreement is obvious, especially for low pressure ratios. The extrapolated curve
would attach zero torque before pressure ratio equals one, as anticipated. For high
pressure ratios, however, the orbit has grown and henceforth encases the steady
state values from hot gas stand.

Regarding the (energy) averaged values (asterisk symbols), an important note is to


make. For the upper right graph in Figure 11, the transferred energy portion is zero
(apart from decimal places). Hence it is not sensible to calculate an energy average
pressure ratio that belongs to the blue orbit, as no net energy is effectively
conducted via the shaft to the compressor. An energy equivalent value that shall be
compared with steady state values from hot gas stand only makes sense, if it based
on the (measured) shaft power that is in fact transferred to the compressor.

315
Figure 11: Measured torque after bearing friction and inertia correction

In Figure 12 the correction results of all curves from Figure 8 - analogous to the
method from Figure 11 – are plotted. Obviously, the slopes of the torque orbits are
much more reasonable and for low pressure ratios show good match with the
steady state data (grey squares and triangles).

Figure 12: Superimposed torque curves (measured + inertia based),


energy averaged values and steady-state torque data obtained at
hot gas stand; all values plotted vs. turbine stage pressure ratio

316
So far, the phase shift between stage inlet, rotor inlet/outlet and stage outlet has
not been considered, as already mentioned in chapter 7. It was necessary to
discuss the steady and unsteady turbine stage characteristics as well as the
corrections that have been introduced above, prior to addressing this topic.

8 SENSITIVITY STUDY REGARDING PHASE SHIFT CORRECTION

As explained, a proper (and even cycle resolved) phase shift correction is hard to
calculate analytically. Hence, based on the known nominal geometry, average
measured mass flows, temperatures and pressures, the gas flow velocity as well as
the speed of sound was estimated by compressible flow equations.

A sensitivity study was undertaken for a phase shift of p3 and/or p4 solely as well as
simultaneously, taking into account gas flow velocity, speed of sound or a
superimposition of both, Table 3. As characteristic length within the volute
geometry, half of its circumferential length has been employed, as recommended
by other researchers, e.g. [14].

Table 3: Results of phase shift correction study for measured turbine stage
inlet and outlet pressure considering gas flow velocity, speed of sound and
their superimposition

Phase shift correction study results for Phase shift correction study for turbine
turbine stage inlet pressure p3 (delay) stage outlet pressure p4 (advancement)
Gas flow Speed of Super- Gas flow Speed of Super-
velocity sound imposition velocity sound imposition
°CA °CA °CA °CA °CA °CA
116_00 20.3 2.7 2.4 19.7 1.1 1
116_01 16.1 2.6 2.2 14.9 1 1
116_02 13.2 2.6 2.2 11.4 1 0.9
116_03 10.2 2.5 2 8.2 1 0.9

It is believed that the most reasonable approach takes into account both gas flow
velocity and speed of sound and that p3 is delayed while p4 needs to be advanced.
The result of this variation is plotted in Figure 13. However, it is not claimed that
the presented phase shift results represent a perfect reproduction of real flow
physics.

Obviously, the observed orbits after phase shift (red solid curves) become
significantly smaller compared to the cases without phase shift (black dotted
curves). This is to be expected if the fundamental hypothesis is, that the rotor
behaves quasi-steadily and the orbits are caused by volute filling and emptying
effects, an assumption that is supported by the findings in [14] for an automotive
size turbocharger turbine stage. After correction of all relevant influences any orbit
should then collapse to a line and this line should be congruent to a measured
steady state speed line obtained at the hot gas stand.

In Figure 13 also the steady state values from hot gas stand that fit approximately
in terms of average reduced speed are plotted additionally (blue squares and
triangles). Especially for the low and mid pressure ratios a very good match for the
slopes and absolute values is obvious.

317
Figure 13: Torque vs. turbine stage pressure ratio before (dotted black
curves) and after (solid red curves) phase shift (parallel delay of p3
and advancement of p4) and comparison with steady-state data

The remaining orbit surface areas (discrepancy between rising and falling pressure
ratio limb of the unsteady torque measurements) can be explained by:

 Remaining, not fully captured volute filling and emptying effects (due to
simple correction procedure with gas flow velocity and speed of sound,
especially since being based on average measured values and not cycle
resolved).
 “Unsteady” / not captured effects due to the rapid change of reduced
speed / highly instantaneous temperature fluctuations shortly after engine
exhaust valve opening.
 Fractions of the rotor speed change that are related to an alternating
compressor mass flow (and hence its power consumption) caused by the
intake strokes of the four cylinders and not related to turbine power
fluctuation. This influences the result of eqs. (7) and (8).
 Real unsteady effects in the turbine wheel / volute-wheel interaction for
very high pressure gradients as well as mass flow gradients.
 No consideration of transiently changing degrees of reaction during the
pulsed flow admission.
 Scavenged mass flow during gas exchange at valve overlap timing that
alters the usual, expected relationship between pressure ratio and rotor
torque.
 Engine speed variations during one complete cycle.

However, it is to note that after phase shift correction, the turbine torque trend
exhibits almost steady characteristics, especially for low stage pressure ratios.

318
9 CONCLUSION AND OUTLOOK

A novel and well validated direct shaft torque measurement technique has been
employed for turbine stage performance assessment under steady state as well as
under engine conditions. For steady state conditions, the achieved experimental
results were compared with CFD data, exhibiting a good qualitative as well as
quantitative agreement.

Under pulsed flow conditions, the obtained unsteady results initially differed
significantly from the measured steady state values from hot gas stand. While the
energy averaged values match reasonably, especially the characteristics in terms of
the slope of the curves and the expected pressure ratio for zero torque showed
deviations. In a step-by-step process, the relevance of bearing friction and
especially rotor speed change was pointed out.

Finally, the role of phase shift correction was discussed exemplarily, based on
simple correlations for average gas flow velocity and the speed of sound. From the
results presented it can be concluded that the aerodynamic behaviour of the
investigated turbine stage can be assumed to be quasi-steady. Unsteady effects
seem to be rather small and therefore are negligible.

The obtained results represent a fundament for the deeper understanding of


turbocharger turbine behaviour not only under engine-like or pulse flow conditions,
but under real engine conditions. The data can be utilized to judge the quality of
engine simulation results of the same operating points, if these are “rebuilt”
virtually within an engine simulation environment. The results can be employed to
validate or (in case of significant deviations) improve state of the art turbine map
extrapolation methods and strategies.

ACKNOWLEDGMENTS

The authors would like to thank Mr Philipp Nitschke from ICSI in Heidelberg for his
valuable contribution to this work. Furthermore, the authors are grateful for the
excellent support of Mr Frederik Haußmann and Mr Marco Leonetti from the FKFS in
Stuttgart.

REFERENCE LIST

[1] Lüddecke, B.; Filsinger, D.; Ehrhard, J.; Steinacher, B.; Seene, C.; Bargende,
M. (2013), “Contactless Shaft Torque Detection For Wide Range Performance
Measurement of Exhaust Gas Turbocharger Turbines”, Proc. of ASME Turbo
Expo 2013: Power for Land, Sea and Air; San Antonio, Texas, USA, GT2013-
94538
[2] Lüddecke, B.; Filsinger, D.; Bargende, M. (2012), “On wide mapping of a
mixed flow turbine with regard to compressor heat flows during turbocharger
testing”, Proc. 10th International Conference on Turbochargers and
Turbocharging of the IMechE.
[3] Cao, T., Xu, L., Yang, M., Martinez-Botas, R., F., (2013) “Radial Turbine Rotor
Response to Pulsating Inlet Flows”, Proc. ASME Turbo Expo 2013: Power for
Land, Sea and Air; San Antonio, Texas, USA, GT2013-95182
[4] Newton, P., Romagnoli, A., Martinez-Botas, R., Copeland, C., Seiler, M.
(2013) “A Method of Map Extrapolation for Unequal and Partial Admission in a
Double Entry Turbine”, Proc. ASME Turbo Expo 2013: Power for Land, Sea
and Air; San Antonio, Texas, USA, GT2013-95815

319
[5] Chiong, M. S., Rajoo, S., Costall, A. W., Bin Wan Salim, W. S.-I., Romagnoli,
A., Martinez-Botas, R. F. (2013) “Assessment of Cycle Averaged Turbocharger
Maps through One Dimensional and Mean-Line Coupled Codes”, Proc. ASME
Turbo Expo 2013: Power for Land, Sea and Air; San Antonio, Texas, USA,
GT2013-95906
[6] Lüddecke, B.; Filsinger, D.; Ehrhard, J.; Bargende, M. (2012), “Heat transfer
correction and torque measurement for wide range performance
measurement of exhaust gas turbocharger turbines”, Proc. 17th
Supercharging Conference, Dresden
[7] Wurms, R., Budack, R., Böhme, J., Dornhöfer, R., Eiser, A.,Hatz (2008), W.
“Der neue 2.0l TFSI mit Audi Valvelift System für den Audi A4 – die nächste
Generation der Audi Turbo FSI Technologie”, 17. Aachener Kolloquium
Fahrzeug- und Motorentechnik
[8] Wieske, P., Lüddecke, B., Ewert, S., Elsäßer, A., Hoffmann, H., Rückauf, J.
(2009), “New Concepts for Optimising Transient Torque Response and Fuel
Economy of Turbocharged Gasoline Engines”, Proc. 18th Aachen Colloquium
[9] Wieske, P., Lüddecke, B., Ewert, S., Elsäßer, A., Hoffmann, H., Taylor, J.,
Fraser, N. (2009), “Optimisation of Gasoline Engine Performance and Fuel
Consumption through Combination of Technologies”, MTZ worldwide Edition:
2009-11
[10] Tiger Version 2, Thermodynamic Package for Internal Combustion Engines,
Enginos GmbH, www.enginos.de
[11] Costall, A., Szymko, S., Martinez-Botas, R. F., Filsinger, D., Ninkovic, D.
(2006) “Assessment of Unsteady Behavior in Turbocharger Turbines”, Proc.
ASME Turbo Expo 2006: Power for Land, Sea and Air; Barcelona, Spain,
GT2006-90348
[12] Capobianco, M., Marelli, S. (2010) “Experimental investigation into the
pulsating flow performance of a turbocharger turbine in the closed and open
waste-gate region”, Proc. 9th International Conference on Turbochargers and
Turbocharging of the IMechE.
[13] Reuter, S., Koch, A., Kaufmann, A., (2010) “Extension of performance maps
of radial turbocharger turbines using pulsating hot gas flow”, Proc. 9th
International Conference on Turbochargers and Turbocharging of the IMechE.
[14] Aymanns, R., Scharf, J., Vedder, R., Wedowski, S., Uhlmann, T., (2011) “A
Revision of Quasi Steady Modelling of Turbocharger Turbines in the Simulation
of Pulse Charged Engines”, Proc. 16th Supercharging Conference, Dresden

320
Performance and flow-field assessment
of an EGR pulse optimised asymmetric
double-entry turbocharger turbine
M Sakai, A Romagnoli, R F Martinez-Botas
Department of Mechanical Engineering, Imperial College London, UK

ABSTRACT

In this paper it will be presented a novel turbine concept specifically designed for
exhausts pulse flow energy conservation and EGR control. In order to combine both
these features, an asymmetrically divided double-entry turbine was developed to
respond to the imbalance of mass flow due to EGR extraction from one side. The
EGR side was equipped with variable geometry vanes in order to control the EGR
rate and to optimize the flow entering the turbine wheel, whereas no vanes were
contemplated in the other turbine entry. A detailed analysis on the design and
efficiency of the asymmetric turbine is provided in this paper.

Keywords: Double-entry turbine, Variable Geometry Turbocharger, Exhaust Gas


Recirculation, Pulse turbocharging, Computational Fluid Dynamics, Unequal
admission

NOMENCLATURE

Π Pressure ratio ߙ Absolute angle (°)


݉ሶ Mass flow (kg/s) ߚ Relative angle (°)
‫ܣ‬ Cross sectional area (mm2) ‫݋‬௡௩ Vane throat distance (mm)
ܴ Radius to area centre (mm) ‫ݒ݊ݏ‬ Vane pitch (mm)
ܿ Absolute flow velocity (m/s) ݈݊‫ݒ‬ Vane chord (mm)
ߩ Density (kg/m3)  Efficiency
ߠ Circumferential angle (°)
ܾ Inlet width (mm)
߰ Azimuth angle (°)
VGT Variable Geometry Turbocharger
EGR Exhaust Gas Recirculation
SPR Scroll Pressure Ratio
CFD Computational Fluid Dynamics
NOx Nitro oxygen

Subscript
0 Total/Stagnation condition L Large scroll
1 Volute inlet S Small scroll
4 Vane exit D Double-entry
‫ݎ‬ Radial direction ts Total-to-static
ߠ Circumferential direction

_______________________________________
© The author(s) and/or their employer(s), 2014
321
1 INTRODUCTION AND PROBLEM DEFINITION

Due to the increasingly stringent emission regulations and demand for high fuel
economy, turbocharger is inherently one of the most promising enabling
technologies towards achieving engine design for low emissions and fuel
consumption. Turbochargers are not only expected to provide efficient exhausts
energy recovery, but they are also used in order to support engine operation by
controlling engine back-pressure for enhanced EGR rate. In large multi-cylinder
engines, single-entry VGTs are normally used due to their ability to support EGR
and instantaneous boost pressure response. However, the disadvantage of single-
entry VGTs is that they do not enable to maximize energy extraction out of the
exhausts pulses (pulse turbocharging) since there is no pulse separation within the
turbine volute (as in multiple-entry turbines). Hence, aim of this research, is that of
combining the advantages of exhausts pulse energy extraction and variable
geometry turbocharging in one single turbine design.

Pulse turbocharging (i.e. conserving pulse energy until the rotor entry) is a well-
known technique to utilise the maximum energy of the exhausts. In order to avoid
exhausts pulse interference, these are isolated with multiple-entry turbines. Two
types of multiple-entry turbines are currently available in the market: meridionally
divided “twin-entry” turbines and circumferentially divided “double-entry” turbines
(Figure 1). The twin-entry turbine is divided meridionally and each incoming flow is
fed into the entire rotor circumference. In contrast, double-entry turbine feeding
area is divided circumferentially and incoming flows are guided with radially
doubled scrolls.

Inner Outer
limb limb

Entry 2 Entry 1
(a) (b)
Figure 1: Comparison between multiple-entry turbine designs (1):
(a) Twin-entry and (b) Double-entry turbine design

In order to merge the requirements of imbalanced mass flow and the pulse
turbocharging advantage, asymmetric vaneless twin-entry turbine design has been
proposed by Müller et al. (2) and lately optimized by Brinkert et al. (3). Their work
showed that it is possible to achieve remarkable EGR-rates in some regions of the
engine map, even though the average exhausts back-pressure is lower than the
charge air pressure thus limiting the operations for this design. A different approach
than that provided by (2) and (3) is proposed in the current paper, with the design
of an asymmetric variable geometry double-entry turbine. The reasons for the
selection of this turbine configuration are explained as follows.

In order to extract exhaust gases and recirculate it into the intake side, exhaust
manifold pressure has to be higher than the intake manifold pressure. Therefore a
VGT is necessary to increase exhaust manifold pressure corresponding with widely
changeable transient operation. Besides supporting EGR control strategies by
changing exhaust manifold pressure, VGT also offers the additional advantage of
varying the inlet area to the turbine wheel thus maximizing exhausts flow energy
extraction at different engine operating conditions. The choice of asymmetric

322
double-entry geometry in place of a twin-entry one can be explained by considering
that adopting symmetrically divided multiple-entry turbines in EGR engines, would
lead to an imbalance of mass flow caused by EGR extraction from one side of the
exhausts manifold. This flow imbalance can hardly be controlled in twin-entry
turbines since the flow leaving the two entries mix together before entering the
wheel. This is not the case in double-entry turbine configuration since the flows
from the two entries are completely isolated and introduced into the turbine wheel
separately, so that the flow controllability is believed to be more effective than in
twin-entry turbine. A typical arrangement for the asymmetric variable geometry
double-entry turbine is given in Figure 2. More details are provided in the next
paragraph.
EGR Vaneless
valve Large scroll
EGR
cooler Vaned
Small scroll

Figure 2: The system of the Asymmetric double-entry turbine with


a 6-cylinders engine

2 PRELIMINARY ANALYSIS

The asymmetric double-entry turbine design started with preliminary off-design


performance analysis. The two turbine inlets were treated as two independent
turbine scrolls. Variable geometry vanes were only contemplated at the end of the
scroll in which exhausts are extracted (small scroll in Figure 2) whereas in the other
scroll no vanes were included. The main purpose of this partial vane arrangement is
to minimise the losses due to the presence of vanes (e.g. vane pressure loss and
vane clearance leakage loss) but still supporting EGR operations. With the
asymmetric variable geometry double-entry turbine housing, the turbine wheel will
mainly be driven by the large scroll flow which is optimised for higher efficiency
over wide turbine operating conditions (no vanes, less losses). In the small scroll
instead, the presence of variable geometry vanes is made necessary in order to
support EGR operations. As the EGR rate varies (hence varies) the variable
geometry vanes will optimize the inlet flow area to the turbine wheel (i.e. the inlet
flow angle) in order to maximize the exhausts flow energy extraction. In order to
calculate the asymmetric double-entry turbine performance, the efficiency of the
small and large scroll turbines were individually calculated using a mean-line model
developed at Imperial College (4). Since the impact of EGR in the large scroll is
negligible, the performance of the large vaneless scroll was calculated as
conventional fixed geometry turbine. In contrast, the small vaned scroll was treated
as single-entry VGT. The overall efficiency of the asymmetric double-entry turbine
was then calculated as mass flow weighted average of the efficiencies of each
turbine scroll.

+
= (1)
+

323
where, ߟௌ and ߟ௅ are individually calculated turbine efficiency of small and large
scroll.

In order to choose the optimum circumferential division1, the efficiency of several


configurations were assessed and compared with that of an equivalent single-entry
VGT with same turbine wheel. The analysis showed that the optimum
circumferential division varies depending on different EGR rates and the final choice
fell on a 160:200 arrangement (refer to Figure 4). The mean-line method results
showed that it is possible to achieve better turbine efficiency than conventional
VGTs when EGR is being operated widely.

1.2 Conventional VGT


Single-entry VGT
Asymmetric double-entry
Double-entry 160:200 turbine
1
Mass weighted efficiency t-s

30% 20% 10%


EGR 40%
0.8
3%

0.6
EGR 0%

0.4

0.2
Asymmetric double-entry preliminary design features
Circumferential division: 160:200 - EGR  10%
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
Normalized Inlet Mass flow to the Turbine

Figure 3: The turbine efficiency advantage of asymmetric double-entry


turbine against conventional VGT when EGR is being operated widely
(The mass flow was normalized with the design-point mass flow rate of the
single-entry VGT used for comparison)

This is shown in Figure 3 where for the selected turbine division, an improvement of
3% in efficiency could be found for a wide range of EGR rates (from 0% to 40%).
This can be achieved by varying the vanes angle in the small scroll in order to
match the best flow angle at the inlet to the turbine wheel and hence maximize
exhausts flow energy extraction. After a preliminary analysis run on a typical duty
cycle of an off-road engine, an EGR rate of 10% was chosen as initial design value
(refer to the text box in Figure 3).

3 ASYMMETRIC DOUBLE-ENTRY DESIGN

The design of the asymmetric variable geometry turbine scroll started with the
selection of the EGR rate and the circumferential division to be considered at the
design-point (10% and 160:200 respectively as described in the previous
paragraph). Once these two parameters were fixed, the design procedure revolved
around the assumption that the incidence angle at the end of small scroll should be
identical to that at the end of large scroll.

1
In an asymmetric double-entry turbine, there are same radii and widths of the flow path but different two
circumferential flow areas for small and large scrolls (Figure 4).

324
From the free vortex correlation, the relationship between the centroid radius of the
area and flow velocity is:

∙ = ∙ (2)

The continuity equation for incompressible flow provides the relationship between
radial flow velocities at the inlet and exit to the volute (station 1 and 4
respectively);

∙ ∙ = ∙ ∙ (3)

By substituting equation (2) and (3) into the following absolute turbine inlet angle,

= (4)

We obtain,

· ·
= (5)
· ·

For incompressible flow, the equation (5) can be simplified,

= ⁄ · (6)

In case of single-entry turbine, = 2 since all the incoming flow is distributed


around the circumference of the turbine wheel, whereas in a double-entry design
turbine the inlet angles should be considered separately and identical to each other,
( + = 2 ).

= ⁄ · = ⁄ · (7)


= (8)

From equation (8), it can be gathered that for a set EGR rate of 10%, the ⁄ ratio
between the large and small scrolls can be expressed as in equation (9).

⁄ %
= = (9)
⁄ %
Then it is now possible to separate circumference into two, and set the ratio
between the two ⁄ s identical to circumferential division.

In order to perform a comparison with a variable geometry single-entry turbine, a


turbine wheel designed and tested at Imperial College by Abidat (5) (and lately
used by Rajoo (6) for single-entry VGT design) was chosen. In order to run a one-
to-one comparison, the sum of the small and large volute ⁄ values and their
ratio against azimuth angle (Figure 4) were designed to be identical to that of the
single-entry VGT designed by Rajoo (6). In other words, despite the asymmetric
double-entry turbine volute is divided into two, the ⁄ values were set to obey to
the free vortex condition following the single-entry VGT designed at Imperial
College. The key features for the VGT asymmetric double-entry turbine are shown
in Table 1 and Table 2.

325
Table 1: Design conditions
Parameter
Pressure ratio 2.91
Inlet temperature 344 K
Mass flow rate 0.678 kg/s
Rotational speed 60000 rpm
Target EGR 10 %

40
Small scroll Large scroll
35
30
A1/R1 [mm]

25
20
15
10
5
0
0 40 80 120 160 200 240 280 320 360
Azimuth Angle [deg]

Asymmetric double-entry turbine Single-entry turbine by Rajoo 2007


(a)
Station 4: vanes exit (downstream)
Small scroll Station 3: vanes exit
Stationedge)
(trailing 3
Station 2: vanes inlet

160

200

Large scroll
Station 1: volute inlet
(b) (c)
Figure 4: Asymmetric double-entry turbine volutes A/R design
(a) Comparison of A/R change along the volutes from its tongue between
asymmetric double-entry turbine and single-entry turbine
(b) Asymmetirc double-entry turbine layout
(c) Single-entry turbine layout for comparison

Table 2: Geometric details of asymmetric double-entry turbine


Geometric feature
Asymmetric circumferential division 160 : 200
A/R Small scroll 13.33 mm
Large scroll 16.67 mm
Radius Tongues 70 mm
Turbine wheel 42.07 mm
(reference diameter)
Number of blades 12
Number of vanes 9
Vanes angle (standard vanes angle) 67.65 °
Vane pitch angle 20 °

326
Unlike single-entry VGTs in which the vanes are arranged around the entire turbine
wheel circumference, in a double-entry turbine configuration, the number of vanes
is strictly related to the location of the two tongues. If the vanes are equally spaced
around the periphery of the turbine wheel, the angle between two vanes is required
to be a common divisor of the asymmetric circumferential division angles in order
to have identical distance from tongue to vane. In the 160:200 circumferential
division, the angle between two vanes should be a common divisor between 200°
and 160° (i.e. 40°, 20° and 10°). In addition to this, the vane pitch is also
dependent on the exit flow angle and the Zweifel’s criterion (7). The former is given
as function of vane geometries (8),

= ⁄ (10)

whereas the latter is a function of the optimum tangential lift coefficient which in
the Zweifel’s criterion (7) is suggested to fall between 0.75 ~ 0.85,

= | − | (11)

Unlike the single-entry turbine housing designed by Rajoo (6), the vanes shape was
changed from straight to curve in order to maintain the vane angle from the leading
edge to the trailing edge and therefore obtain the same flow angle (in station 4) as
that in the vaneless section.

4 CFD ANALYSIS

4.1 Computational analysis and discussion


In order to understand the turbine basic behaviour of the asymmetric double-entry
turbine, the simulation results have been obtained with CFD analysis. The CFD
analysis was conducted using commercial software ANSYS-CFX and in Table 3 it is
provided the mesh characteristics of the whole turbine domain.

Table 3: Mesh characteristics of turbine domain


Number of Number of
Region Element type
Elements Nodes
Unstructured
Volute 575,323 120,868
(Tetrahedra, Wedges)
Structured
(Hexahedral) and
Vane section 78,180 75,684
Unstructured
(Tetrahedra, Wedges)
Structured
Rotor domain 1,246,416 1,372,956
(Hexahedral)
Unstructured
Exit ducting 39,045 8,121
(Tetrahedra, Wedges)
Total 1,938,964 1,577,629

The turbine performance analysis was run under equal (same pressure ratio within
the turbine inlets) and unequal admission (unbalance of pressure ratio between the
inlets) conditions. For ease of discussion a parameter, here defined Scroll Pressure
Ratio (SPR), has been introduced (equation 12) to provide the rate of imbalance
between the small and large scroll;

= = (12)

327
1.0 1.0
30000rpm (50% speed) 60000rpm (100% speed)
0.9 0.9

0.8 0.8

Turbine Efficiency t-s


Turbine Efficiency t-s

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 (a) 0.1 (b)


0.0 0.0
1.0 2.0 3.0 4.0 1.0 2.0 3.0 4.0
Pressure Ratio t-s Pressure Ratio t-s

1.2 1.2

1.0 1.0
Normalized Mass Flow rate

Normalized Mass Flow rate

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

(a) (b)
0.0 0.0
1.0 2.0 3.0 4.0 1.0 2.0 3.0 4.0
Pressure Ratio t-s Pressure Ratio t-s

Figure 5: Turbine performance of the Asymmetric double entry turbine at


(a) 50% speed and standard vane position,
(b) 100% speed and standard vane position
(The mass flow was normalized with the design-point mass
flow rate of the single-entry VGT used for comparison)

The analysis started by setting the vane angle at design-point (refer to Table 2),
considering tow rotational speeds of 60000rpm (100% design-speed) and
30000rpm (50% design-speed, in which it is believed large amount of EGR is likely
to be required), and five different values for the SPR. The simulation results are
given in Figure 5 and show that the highest turbine efficiency occurs under equal
admission conditions with an efficiency value of 76% at 1.3 pressure ratio for 50%
speed and 78% at 2.43 pressure ratio at 100% speed. This is consistent with the
initial design assumption of peak efficiency point occurring at 2.91 pressure ratio
(5) (6) (refer to Table 1) and also with previous available literature showing that in
multiple-entry turbines the peak efficiency point occurs under equal admission (9)
(10) (11). As the rate of imbalance between the two inlets increases, the turbine
performance decreases significantly, independent of which limb is flowing less mass
flow (either SPR 0.5 or 1.5 present large efficiency drop).

328
4.0
30000rpm (50% speed)
available under pulsating flow
condition only
3.5 Blocked flow in small scroll
small scroll mass flow
Blocked flow in large scroll cannot exceed large scroll
mass flow
SPR 0.5
3.0
SPR 0.8
PRsmall

SPR 1.0
2.5
SPR 1.2

SPR 1.5
2.0

B C
A
1.5
C
Backflow in
Backflow in
small scroll D D large scroll

1.0
0.0 0.5 1.0 1.5 2.0
SPR

Figure 6: Small turbine volute flow interaction:



A : Steady-state operation, ○
B: ,○
C: ,


D : 0

In order to understand the limit of mass flows through which the asymmetric
double-entry turbine can operate, typical mass flow patterns are illustrated in
Figure 6 for 50% turbine speed. In the figure have been identified four flow areas,
from A to B, trying to include also the effects of pulsating exhausts flow conditions.
Starting with steady-state flow assumption, it can be noticed that in contrast with
standard flow type A, the mass flow in the small scroll can never exceed that in the
large scroll due to the cylinder distribution (3 cylinders connected to the small
volute and EGR circuit, and other 3 cylinders connected to the large scroll:
). Hence the flow type B (blue shaded area in Figure 6) cannot be
obtained in the steady-state condition of asymmetric double-entry turbine design.
As the pressure in the small scroll keeps decreasing as consequence of larger and
larger EGR rates, the mass flow in the small scroll would experience a blocked flow
condition (dashed line, flow type C) and in an extreme case some backflow could
occur (flow type D). In the large scroll instead the mass flow can never be
exceeded by that in the small scroll since in the large scroll there is not EGR flow
extraction. Hence flow type B, flow type C and flow type D in the large scroll can
only be obtained in an experimental lab set-up but it would not occur during
standard engine operating conditions since there is no EGR flow extraction.
However, it is worth noting that the results of Figure 6 are true only under the
assumption of steady-state flow. This is not the case in a real engine since the
exhausts flow are instantaneously pulsating. Therefore over an entire engine cycle,
due to the firing order of the engine, the mass flow in large scroll can be exceeded
by that in the small scroll as shown from the black marks in Figure 6. However this
will not be discussed further within this paper which is currently looking at steady-
state efficiency assessment for the asymmetric double-entry turbine design.

329
4.2 Turbine performance and optimisation with vanes angle
From the previous analysis it was found that the highest efficiency point occurs
under equal admission conditions. If the vanes angles were not optimized to
achieve equal admission condition, the turbine efficiency would drop due to the
imbalance of mass flow between the two inlets. By changing the vanes angle when
target EGR rate is changed, it is possible to keep higher turbine efficiency. In other
words, the optimum EGR rate can be adjusted by varying the vanes angle position.

1.0
Turbine Efficiency t-s

0.8
0.6
0.4
0.2
(a)
0.0
0% 10% 20% 30% 40% 50% 60% 70% 80%
EGR rate

1.0
Turbine Efficiency t-s

0.8
0.6
0.4
0.2 (b)
0.0
0% 10% 20% 30% 40% 50% 60% 70% 80%
EGR rate

1.0
Turbine Efficiency t-s

0.8
0.6
0.4
0.2 (c)
0.0
0% 10% 20% 30% 40% 50% 60% 70% 80%
EGR rate

Figure 7: Comparison of optimum EGR rate:


(a) at 50% speed and -10° of vanes angle
(b) at 50% speed and standard vanes angle (design-point)
(c) at 50% speed and +10° of vanes angle

In Figure 7 it is reported a comparison between efficiencies at 50% speed for


different vanes angle position, EGR rates and SPR values. EGR rate is expressed as
in equation 13.

330

= = (13)
+ + ×

Two additional vanes position (±10° of the design-point vanes angle) were
considered in the analysis, as shown in Figures 7a and 7c. At design-point vane
position (Figure 7b), the optimum EGR rate is approximately 20%2 as shown by the
black bold line. As the SPR decreases slightly (SPR 0.8) in order to maintain the
same level of efficiency within the turbine, the EGR rate should increase
significantly (more than 30% EGR). In some engine operating conditions, large EGR
rate is not required, and therefore in order to maintain optimum turbine flow
conditions (corresponding to equal admission conditions), the vanes angle need to
be varied. By slightly varying the vanes position (±10° of the design-point vanes
angle), an equal admission condition can be obtained for no excessive EGR rate
values (10% and 30% for -10° and +10° vane position respectively). It is obvious
that the equal admission condition line, which gets always maximum efficiency, is
totally depending on the variable geometry vanes angle in Figure 7. However, it is
notable that the equal admission condition is not always available since the intake
manifold pressure and the exhaust manifold pressure are always changing on
engine operation map.

5 CONCLUSION

In this paper, the design and performance assessment of a novel asymmetric


double-entry variable geometry turbine was discussed. This novel turbine housing
design was conceived for improving engine EGR operation and enhancing exhausts
pulsating flow energy extraction in large multi-cylinder engines. The design started
with preliminary mean-line analysis in order to fix the basic turbine geometrical
parameters. Then the complete turbine design moved into 3-D modelling and CFD
performance assessment. As a result of this process, the final turbine housing
arrangement comes as asymmetric double-entry with circumferential division of
160:200 for 10% EGR rate (design-point) and variable geometry vanes only in the
turbine inlet where EGR flow extraction occurs.

Simulations were run for a number of different turbine speeds (100% and 50%
design-speeds), SPRs (from 0.5 to 1.5) and vanes angles (-10°, ±0°, and +10°).
The simulation results showed that the peak efficiency point occurs under equal
admission (SPR 1.0) with 78% turbine efficiency at 100% design speed and 20%
EGR rate. Thanks to vanes position adjustment no much penalty in efficiency was
observed at 50% speed, with peak efficiency value of 76%. The benefit of variable
vane configuration could be appreciated for different EGR rates where the
possibility to optimize the flow condition at the inlet to the turbine wheel showed
that it is possible to retain an equal admission conditions (and hence optimum
turbine efficiency) for a wide range of EGR rates.

As final remark about this project, it is worth saying that a prototype of the
asymmetric variable geometry double-entry turbine will soon be tested at Imperial
College London. The test programme will focus on the validation of the presented
CFD results as well as in pulsating flow performance assessment.

2
Despite 10% target EGR rate the optimum EGR rate calculated by CFD was 20%. This may require some
further investigation and analysis for the vanes design and profile.

331
REFERENCE LIST

1. Watson, N. and Janota M. S., 1982, “Turbocharging the Internal Combustion


Engine”, pp. 60.
2. Müller, M. et al., 2011, “The Asymmetric Twin Scroll Turbine for Exhaust Gas”,
Proceedings of ASME Turbo Expo 2008, GT2008-50614.
3. Brinkert, N. et al., 2011, “Understanding the Twin Scroll Turbine-Flow
Similarity”, Proceedings of ASME Turbo Expo 2011, GT2011-46820.
4. Romagnoli, A. and Martinez-Botas, R. F., 2011, “Performance Prediction of a
Nozzled and Nozzleless Mixed-Flow Turbine in Steady Conditions”, International
Journal of Mechanical Sciences, Vol. 53, pp. 557-574.
5. Abidat, M., 1991, “Design and Testing of a Highly Loaded Mixed Flow Turbine”,
Ph.D. Thesis, Imperial College of Science, Technology and Medicine.
6. Rajoo, S., 2007, “Steady and Pulsating Performance of a Variable Geometry
Mixed Flow Turbocharger Turbine”, Ph.D. Thesis, Imperial College of Science,
Technology and Medicine.
7. Hiett, G. F. and Johnston, I.H, 1963, “Experiments Concerning the
Aerodynamic Performance of Inward Flow Radial Turbine”, Proceedings of
Institution of Mechanical Engineers, 178, Pt 31(II).
8. Japikse, D. and Baines, N., 1984, “Introduction to Turbomachinery”, pp. 6-17.
9. Copeland, C. D., Martinez-Botas, R. F. and Seiler, M., 2008, “Unsteady
Performance of a Double Entry Turbocharger Turbine with a Comparison to
Steady Flow Conditions”, Proceedings of ASME Turbo Expo 2008, GT2008-
50827.
10. Copeland, C. D., Martinez-Botas, R. F. and Seiler, M., 2011, “Comparison
Between Steady and Unsteady Double-Entry Turbine Performance Using the
Quasi-Steady Assumption”, ASME Journal of Turbomachinery, Vol. 133, 031001.
11. Romagnoli, A. et al., 2012, “Comparison between the Steady Performance of
Double-Entry and Twin-Entry Turbocharger Turbines”, ASME Journal of
Turbomachinery.

332
Comparison of the influence of
unsteadiness between nozzled and
nozzleless mixed flow turbocharger
turbine
M Y Yang1, M H Padzillah1,2, W L Zhuge1,3, R F Martinez Botas1, S Rajoo2
1
Imperial College London, UK
2
Universiti Teknologi Malaysia, Malaysia
3
State Key Laboratory of Automotive Safety and Energy, Tsinghua University,
China

ABSTRACT

Nozzled and nozzleless volutes are the most commonly used stator configurations
for the turbocharger turbine. In this paper detailed experimental investigation was
carried out to compare the performance of a mixed flow turbine with nozzled and
nozzleless stator under steady and pulsating conditions which replicates the
pulsating exhaust flow from the internal combustion engine. The results show that
the turbine with nozzled volute has higher efficiency at low load conditions in both
steady and unsteady conditions. The steady peak efficiency is nearly the same as
the nozzleless one but it shifts to higher velocity ratio. The improvement of the
cycle average efficiency for the nozzled volute under pulsating conditions is
influenced on the pulse frequency. The hysteresis loops of the swallowing capacity,
reflecting the filling and emptying as well as the wave action effect in the turbine,
are similar for two configurations. The comparison shows that the volute
configuration has no direct influence on the ‘unsteady’ behaviour in the turbine.
Instead, the results of the 1-D model show that the swallowing capacity curve has
an evident influence on the unsteadiness in the turbine. Furthermore, the
unsteadiness of a turbine is enhanced by reducing the swallowing capacity or
increasing the turbine loading.

NOMENCLATURE

IC Internal combustion
Cr Radial component of velocity m/s
Cϑ Tangential component of velocity m/s
C velocity m/s
U Rotor tip velocity m/s
PR Pressure ratio
P pressure Pa
MFP Mass flow rate parameter kg/s. √ /Pa
m Mass flow rate kg/s

_______________________________________
© The author(s) and/or their employer(s), 2014
333
Greek letters
α absolute flow angle deg.
η efficiency
β relative flow angle deg.
torque N.m

Subscripts
0 ambient parameter
ave cycle average
in Inlet
out Outlet
V nozzled (with vane)
VL nozzleless (vaneless)

INTRODUCTION

The increasing demand of vehicle with low CO2 emission has forced the industry
towards highly efficient engines. A turbocharger turbine plays a key role in
recovering the energy from the exhausted gas of an engine and hence attracts
intense researches targeting at the performance improvement. Generally there are
two types of the turbine identified by the configurations of the volute, including
nozzleless turbine and nozzled turbine. The nozzleless turbine is usually applied for
the turbocharger of the IC engine in passenger vehicle due to its simplicity (cost)
and reliability. On the other hand, in relatively larger IC engine (10-15L), a nozzle
ring is usually employed to turn the flow in limited space in the turbine stator, and
hence results a turbine with smaller size.

The performance comparison under steady conditions between the turbine with
nozzled and nozzleless volute has been proceeded extensively by experimental or
CFD method in literatures. The turbine with nozzled volute was claimed in most of
the researches to have better peak efficiency than the one with nozzleless stator.
Baines and Lavy [1] indicated that nozzled turbine offers superior efficiency
compared to nozzleless turbine near the peak efficiency point. The efficiency of the
nozzled one drops gradually below its nozzleless counterpart as the operation point
deviates from the peak point. The experimental investigation of the flow pattern at
the exit of volute showed that highly distorted flow appeared in the nozzleless
volute especially at the region near the tongue. While the flow distortion caused by
the volute tongue is damped by the nozzle, thus produces a more uniform flow
distribution at the rotor inlet. The different flow pattern at exit of the volute was
considered to be the main reason for discrepancy of the turbine efficiency [1, 2, 3].
A direct performance comparison between the two volute configurations in a radial
turbine was conducted via experimental and CFD method [4, 5]. The swallowing
capacity is remained similar for two configurations by adjusting the vane
geometries. The results showed that the nozzleless volute always has superior
efficiency at all operating conditions. It was implied that the different observation
from conventional conclusions was attributed mainly to different surface finishing
when the comparison was made. CFD results showed that the flow was less uniform
in the nozzled volute due to wake and jet caused by nozzle vanes, and higher
pressure loss was produced in the nozzled volute due to the introduction of vanes.
Except for the investigations on the radial turbine, the experimental comparison
was also carried out on a mix flow turbine [6]. The volute with a removable nozzle
was used on a mix flow rotor in the testing. The efficiency of the nozzled turbine
shows constantly better efficiency than the nozzleless one.

334
Few investigations have been conducted on the performance comparison under
pulsating conditions between the nozzled and nozzleless turbine. Romagnoli [7]
firstly experimentally compared the unsteady performance of a turbine with
nozzleless and nozzled volutes. The results showed that the hysteresis loops of the
swallowing capacity were quite different between two configurations and the vanes
of the nozzle were considered to be the reason for the discrepancies. The cycle
averaged efficiency of the nozzleless volute superiors the nozzled one. It is
noteworthy that the swallowing capacity of the two configurations was not matched
for the comparison and the turbine with nozzled volute was not working at the
optimized vane angle (60 degrees, [8]). Rajoo [9] conducted detailed experimental
comparison of unsteady performance of a nozzled turbine and the results
demonstrated that both the shape and the area covered by the hysteresis loops
varied with the nozzle vane angle. The nozzle was considered to shield the rotor
from the pulsating flow and produced a different hysteresis loop when the opening
of the nozzle changed.

The current research experimentally investigates the influence of the volute


configuration on a mix flow turbine performance under steady and pulsating flow
conditions. Furthermore, the influence of the swallowing capacity on the
unsteadiness in the turbine is demonstrated via the well calibrated 1-D unsteady
method.

1 TEST FACILITIES AND EXPERIMENTAL METHOD

The test was carried out on the rig in Imperial College London. The layout of the rig
and the turbine for investigations are shown in figure 1. The turbine is driven by the
compressed air and heated to 65 degree centigrade by an electrical heater. The
pulsated air flow is produced by the pulse generator which produces a similar pulse
shape as the engine exhausted gas. The output power of turbine is absorbed by
electrical eddy current dynamometer, of which the maximum output power and
speed can be achieved at 60 kW and 68 kRPM respectively.

Figure 1. Layout of the test rig, volute and the mix flow rotor

The investigated mixed flow turbine with the leaned vane nozzle was developed by
Imperial College [8, 10]. The main geometries of each component are listed in table
1. Both vane angle and number of the nozzle are adjustable. The nozzled and
nozzleless volute configurations for comparison are shown in figure 2. For the
nozzless turbine testing, the vanes are removed and strut holes are sealed by the
same base of vane, as shown in sub-figure (b).

335
Table 1. Main geometries of the mix flow turbine

Stator
A/R of the volute 33 mm
Vane number of nozzle 15
Rotor
Mean inlet radius of rotor 41.8 mm
Inlet blade angle at middle of blade 20 deg.
Cone angle of blade 50 deg.
Hub Radius at rotor exit 13.5 mm
Tip radius at rotor exit 39.3 mm
Exit blade angle -52 deg.

(a) nozzled volute (b) nozzleless volute


Figure 2. Configurations of nozzled/nozzleless volute

The comparison between the configurations with nozzle and nozzleless is carried
out under the same swallowing capacity characteristic. It is considered to be a fair
comparison under this constraint since the inlet conditions imposed at the turbine
inlet is the same for two configurations given the same velocity ratio or inlet
pressure. In order to achieve the similar swallowing capacity, the performance of
the turbine with nozzless volute will be tested as the baseline firstly. Then the
swallowing capacity of the nozzled turbine is adjusted to match with the baseline
via changing the vane angle. The same vane angle is set for the nozzled volute
during the unsteady test.

For the steady test, the performance is measured at two speed lines as 30 kRPM
and 48 kRPM. The chopper plates are set to be fully open. The mass flow rate is
measured by McCrometer V-Cone Flow meter (dP meter) of which the accuracy is
±0.5%. The pressure upstream the turbine inlet and downstream the turbine exit
are evaluated by a scanivalve and the temperature is measured by T-type
thermocouples about 600mm upstream the volute inlet (measurement plane). The
output power is evaluated by the torque from dynamometer as well as the rotor
rotational speed. For unsteady test, the performance is measured under three
different frequencies (20HZ, 40HZ and 60HZ) and three different loads conditions
(low, middle and high loads) at the two speed lines. The instant static pressure are
measured at the measurement plane and turbine exit by Schaevitz pressure
transducers P700 and instant temperature are evaluated by T-type thermocouples
together with the instant pressure. Instant torque was evaluated by the rotor
acceleration plus the load cell reading. The instant mass flow rate is measured by
Dantec 1-D hotwire 55P16 and the device is calibrated by the V-Cone flow meter
under steady conditions over the range of the mass flow and temperature covered
by a pulse.

336
2 COMPARISON OF THE STEADY PERFORMANCE

Figure 3(a) shows the swallowing capacity of the two volute configurations at 30
kRPM and 48 kRPM. It can be seen that the swallowing capacity for the nozzled and
nozzleless volutes at both speeds is matched properly over the whole measured
pressure ratio range. The maximum difference of mass flow parameter is lower
than 4% which exists at the highest pressure ratio at 48 kRPM turbine speed. As
discussed before, the matching procedure is conducted by adjusting the nozzle
vane angle. The correspond vane angle of the nozzled volute is found to close to
the optimized value (60 degrees), thus the efficiency of the nozzled turbine can be
expected to be good, which is going to be demonstrated following.

The total-static efficiency for both volute arrangements at the two speeds is shown
in figure 3(b). The performance superiority of the nozzled volute over its nozzleless
counterpart at the two rotational speeds is clearly demonstrated in the figure. The
peak efficiency for two configurations is similar, but the advantage of the efficiency
for the nozzled one becomes obvious as the velocity ratio increases. Moreover, the
efficiency improvement by the nozzled volute is more evident as the speed
increases from 30 kRPM to 48 kRPM: the peak efficiency is 0.69 for both
configurations at 30 kRPM, and it is slightly higher for the nozzled volute than the
nozzleless one at 48 kRPM, which is 0.738 and 0.731, respectively. As the velocity
ratio increases to the right side of the peak efficiency point, taken the point at 0.9
for instance, the nozzled volute shows 5% efficiency improvement at 30 kRPM while
7% at 48 kRPM.

(a) swallowing capacity (b) total-static efficiency


Figure 3. Steady performance comparison of two volute configurations

Another important observation from figure 3(b) is the locations of the peak
efficiency point. It can be seen that the employment of nozzle shifts the peak
efficiency point towards high velocity ratio region. At 30 kRPM speed, the peak
efficiency point moves from velocity ratio of about 0.55 to 0.68 as the nozzled are
installed. The shifting can still be observed but less obvious for 48 kRPM speed
partly due to the smaller measured range. This behaviour of the efficiency
performance for the two configurations could be explained by the changes of the
velocity triangle at the inlet of the rotor, as discussed below.

By inducing the nozzle vanes to the volute, the actual flow area at the rotor inlet
will reduce due to the blockage of the vanes compared with the nozzleless volute.
In order to maintain similar swallowing capacity as the nozzleless configuration, the
radial component of the velocity Cr has to be increased, as shown in figure 4. The
peak efficiency of the turbine with either volute is considered to be achieved at the
same optimized incidence angle β at the rotor inlet. Under constraint of the same
swallowing capacity (CrVL < CrV), the vane of the nozzle should be increased (αV >

337
αVL) to achieve the optimized incidence angle (β V = β VL). As a result, the tangential
component of the absolute velocity at the peak efficiency is lower for the nozzled
volute (CϑV < CϑVL), and hence the lower available power at the rotor inlet according
to the Euler work equation. It consequently results in a lower isentropic velocity for
the nozzleless configuration thus the higher velocity ratio at peak efficiency point.

Figure 4. Velocity triangle at rotor inlet for nozzled and nozzleless volute

The efficiency comparison between the nozzled and nozzles volute shows that the
conventional claim of the benefit at peak efficiency for the nozzled volute is not
quite valid, at least for the investigated turbine. Instead, the peak efficiency for two
configurations is very similar. It was usually considered that the uniform flow at the
nozzle exit contributes to the higher peak than the nozzleless volute. However, the
nozzleless volute in the experiment is produced by removing the flexible nozzle
vanes from the nozzled volute, thus leaving large interspace region. This could
benefit the turbine performance in such a way that flow distortion caused by the
volute tongue is damped significantly in the large interspace region. On the other
hand, the pressure loss induced by the nozzle vanes reduces the turbine efficiency
and therefore balances the benefit of the uniform flow. As a result, the peak
efficiency achieved by two configurations is at similar level. At off-peak conditions,
the nozzled volute shows the efficiency advantage at most of the steady conditions,
especially at high velocity ratio conditions. Different from the method utilized by
Simpson [11], the swallowing capacity manipulation in current research is achieved
by adjusting the vane angle instead of the vane profile. The shifting of the peak
efficiency results from the vane angle adjustment and therefore the efficiency at
low velocity ratio can be benefit from it.

3 UNSTEADY TEST RESULTS

3.1 Cycle average performance


The convenient way to compare the performance of different configurations under
pulsating conditions is cycle average parameters via which the influence of the
phase shifting methods can be avoided. The cycle average efficiency is evaluated by
the integration of actual power and isentropic power over a pulse cycle:

( ∙ )
= (1)
∙ ∙ ∙( ( )

The cycle average pressure ratio and mass flow rate are defined to evaluate the
mean load of the turbine and the mean swallowing capacity under pulsating
conditions, respectively:

= (2)

338
= (3)

Figure 5 shows the plots of cycle average efficiency at different loadings and
frequencies at two rotational speeds. At 30 kRPM the nozzleless turbine works more
efficiently at high-load-low-frequency condition than the nozzled one (subfigure
(a)). Specifically, the efficiency at 20HZ high load condition (PRave=1.24) is 6.4%
higher for the nozzleless one (location ‘A’). As the load reduces and the frequency
increases, the nozzled turbine gradually shows the advantage over the nozzleless
one. For instance, the efficiency of the nozzled turbine is 11.1% higher at 60HZ low
load condition (location ‘B’). As the speed increases to 48 kRPM, the advantage of
the efficiency for the nozzled turbine is more obvious than at 30 kRPM, as shown in
subfigure (b). The nozzled turbine has the biggest advantage at low frequency low
load condition where the efficiency is 13% higher (location ‘C’).

(a) 30 kRPM (b) 48 kRPM


Figure 5. Cycle average efficiency at different frequencies/pressure
ratios at 30/48 kRPM

In generally it can be observed from both subfigures that the nozzled turbine is
more efficient at low load conditions, although the magnitude is obviously
influenced by the frequency. It is noteworthy that the similar trend can be seen in
the steady performance shown in figure 3(b). Moreover, the advantage of efficiency
for the nozzled turbine under pulsating conditions is more evident at the higher
speed, which is the similar as the phenomenon at steady conditions. It is thus
implied that the cycle average performance of the investigated turbines has similar
behaviour with the steady performance, but obviously the frequency imposes an
evident influence on the magnitude.

3.2 Hysteresis Loops of the Swallowing Capacity


The unsteady behaviour of the turbine under pulsating conditions is typically
characterized by the hysteresis loops of its swallowing capacity. Figure 6-7 show
the swallowing capacity loops of two configurations at different operation
conditions. Figure 6 compares the loops at 48 kRPM middle load conditions for three
frequencies at 20HZ, 40HZ and 60HZ. The shape of the loops is evidently
influenced by the frequencies for both configurations: the loop at 20HZ is slim and
stretches along in pressure ratio axis. The loop follows the steady curve well except
at high load conditions. As the frequencies increases to 40HZ, the loop expands
dramatically indicating significant mass imbalance in the turbine. However, the loop
shrinks as the frequency further increases to 60HZ.

339
(a) nozzleless (b) nozzled
Figure 6. Swallowing capacity loops at 48k RPM middle load different
frequencies

The ‘unsteadiness’ of the turbine, which is reasonable to be evaluated by the area


covered by the hysteresis loops and indicates the magnitude of mass imbalance
during a pulse cycle, is not proportional to the frequency. The magnitude of the
pulse also contributes to the unsteadiness in the turbine, according to the research
of literature [14]. More importantly, comparing the loops between two
configurations, it can be seen that the shape as well as the ‘unsteadiness’ are very
similar although 15 nozzle vanes has been installed in the nozzled volute. It is
inferred that the unsteady behaviour (filling and empty/ wave action) is not directly
influenced to the volute configuration (nozzled/nozzleless).

Figure 7 shows the influence of the load on the unsteadiness in the turbine. The
swallowing capacity loops of two configurations are compared for three different
loads conditions at 40HZ. In general the loops extend outwards and move to higher
pressure ratio along the steady curves as the load increases. It is demonstrated
clearly that the load, quantified by the cycle averaged pressure ratio, enhances the
filling and emptying effect in the turbine. Comparing the loops between two volute
configurations, it is clear that the variation of the loops at the corresponding load
shows a very similar trend, for both covered area and shapes. Again the nozzle
vanes have little influence on the unsteady behaviour in the turbine for those two
configurations.

(a) nozzleless (b) nozzled


Figure 7. Swallowing capacity loops at 48k RPM 40HZ different loads

340
Different from the conclusions drawn in the literatures, the comparison discussed
above demonstrates clearly that the unsteady behaviour of the turbine is not
directly influenced by the volute configurations. Both the shape and the capsuled
area of the loops are similar for the turbine with nozzled and nozzles volute as long
as the turbine works at the similar unsteady conditions (in terms of pulse frequency
and the mean load). Instead, the ‘unsteadiness’ in the turbine under pulsating
conditions is significantly influenced by the pulse frequency and turbine load, for
which the reason can be partly explained by the research [12]. More importantly, it
should be reminded that the swallowing capacity of the two configurations is the
same during the testing. This constraint of the swallowing capacity is also one of
the key reasons for the phenomenon, which is going to be discussed following.

3.3 Influence of the swallowing capacity on the hysteresis loops


In order to investigate the influence of the characteristic of swallowing capacity on
the turbine unsteady behaviour, the hysteresis loops are going to be compared at
different swallowing capacity curves and different load conditions in this section.
However, the magnitude of the pulse is closely coupled with the frequency and
turbine load in the experimental testing and can’t be decoupled. On the other hand,
the magnitude of the pulse is directly related with the unsteadiness of the turbine.
Therefore, in order to investigate the influence of the swallowing capacity on the
unsteady behaviour, a 1-D unsteady non-viscosity model is developed for the
investigation based on the in-house code ‘ONDAS’ [13].

The geometries of the turbine system are simplified as one-dimensional pipes


connected by the same topology. The length and inlet/outlet area (thus the
volume) of the pipes are determined by the real geometries. The configuration of
the model is shown in figure 8. The volute is simplified into three components in
series: a nozzle to represent the part from inlet to the throat; a straight pipe to
present the passage of the volute; a nozzle to represent the nozzled or nozzles part
at the volute exit. The configuration of different vane angle is modelled by the
inlet/outlet area ratio of the nozzle. The turbine rotor is treated as a quasi-steady
component and the swallowing capacity curves are employed to represent the
rotor. Test results including inlet total pressure and temperature are employed as
the inlet boundary conditions. Ambient static pressure is used as outlet boundary
condition. Two-step Lax-Wendroff scheme together with TVD method is applied to
discretize the computational domains.

Figure 8. One dimensional unsteady model for the turbine

341
Figure 9. Validation of the 1D model at 40HZ middle load at 48 kRPM

The predicted hysteresis loop is compared with the test results at 40HZ middle load
at 48 KRPM in the figure 9. Both the mass flow rate range and the shape of the loop
are well predicted by the model. Specifically, the mass flow rate is well predicted in
filling process but marginally under-predicted in emptying process, resulting in a
larger hysteresis loop compared with the test result. The volume of the pipes in the
model and the approximation of the length of the volute are considered to be the
main reasons for the discrepancy. The fillets and the blockage factor of the pipes
are not modelled in the model which will result in a larger volume in the model.
Nevertheless, the 1D model is considered to be reliable enough for the investigation
according to the prediction results.

As the main purpose of the discussion is to understand the effect of different steady
flow swallowing capacity to unsteady turbine behaviour, and the experimental data
of different swallowing capacity (achieved by adjusting the nozzle angle) is not
available, two curves of pressure ratio versus mass flow rate parameter are
produced by mathematical extrapolation, representing the larger vane angle (close)
and smaller vane angle (open). By extrapolation the swallowing capacity curve is
intersected with the horizontal axis and the value is the pressure ratio produced by
the centrifugal force at the inlet of the rotor, thus this value is the same for all the
vane angles. Using the point as the anchor, the swallowing capacity curve of the
original vane angle is increased by about 20% and reduced by 20% representing
the open and close respectively, as shown in the figure 10(a).

Figure 10 shows the hysteresis loops predicted by the unsteady model. Figure 10
(a) compares the hysteresis loops with different the swallowing capacity at the
same inlet condition. The influence of the swallowing capacity on the loops is clearly
demonstrated in the figure. As the swallowing capacity drops (by closing the vane),
the loop expands significantly indicating enhancement of mass imbalance in the
turbine. The shape thus the wave action in the turbine changes as well, but not as
dramatically as the covered area. It should be noticed that the pulse magnitude and
the pressure ratio range are the same due to the same inlet condition, thus the
discrepancies of the loops are only caused by the turbine swallowing capacity. It is
further inferred that for turbines confronted by a same pulse of the inlet pressure,
the turbine with smaller swallowing capacity is more ‘unsteady’ than the one with
larger swallowing capacity.

342
As discussed in the test results, the turbine load has been shown to have significant
influence on the unsteadiness. However, since the magnitude of the pulse is closely
coupled with the turbine load in the test due to the pulse generation mechanism, it
is difficult to check the influence of the turbine load separately. On the other hand,
it is convenient to decouple the influence of the turbine load and the pulse
magnitude in the 1-D unsteady mode. Figure 10 (b) compares influence of the
mean load ( ) on the hysteresis loops under the pulses with the same
magnitude of the inlet pressure. The for the high load is 16% higher than
the original load (middle load) and the 15% lower for the low load. Dramatic
change of the loops can be seen in the figure: the loop is quite slim at low load, and
then expands outward quickly as the load increases. It is clearly demonstrated that
the turbine with higher load has stronger ‘unsteadiness’ in the turbine.

(a) three different swallowing capacity curves

(b) three different mean loads


Figure 10. Comparison of hysteresis loops for different
swallowing capacity and mean loads

343
The results of the reduced order unsteady model further confirm that the unsteady
behaviour of the turbine is significantly influenced by the swallowing capacity curve
and the turbine load. The unsteadiness can be enhanced by either reducing the
swallowing capacity or increasing the turbine load. By closely checking those two
factors, it can be seen that they can be linked to a common factor – the slope of
the swallowing capacity curve. As the vane angle reduces, the slope of the curve at
the same load reduces (subfigure (a)); or as the turbine load increases, the slope
at the mean load point on the curve reduces (subfigure (b)). Actually the effect of
the turbine rotor on the flow can be considered as throttling, thus the slope of the
swallowing capacity reflects the throttling effect of the turbine. Therefore, it is
concluded that the slope of the swallowing capacity curves of a turbine has
significant influence on the unsteadiness, either by closing the nozzle vane or
increasing the turbine load. The smaller slope or higher throttling effect can
enhance the unsteadiness of the turbine.

4 CONCLUSIONS AND DISCUSSIONS

Experimental Testing and reduced order unsteady model are applied for the
comparison of the performance of the nozzled and nozzleless mix-flow turbocharger
turbine under steady and pulsating conditions in this paper. Conclusions are drawn
as following:

(1) With the similar swallowing capacity, the nozzled turbine shows the advantage
of the efficiency over the nozzleless one at most of the conditions. The peak
efficiency is similar for two configurations, but the advantage of the nozzled
one is evident at low pressure ratio conditions. Moreover, the peak efficiency
shifts to higher velocity ratio with the inducing of the nozzle. It results from the
larger absolute flow angle at the nozzled volute exit, which is used to
compensate the blockage effect of the nozzle under the constraints of the same
swallowing capacity.

(2) Generally the cycle average efficiency has similar trend as the steady
conditions: the nozzled volute is more evident at low pressure ratio thus low
load conditions. But the magnitude of the efficiency advantage is highly
dependent on the pulse frequency and rotor speed. Specifically, at 30 kRPM,
the maximum efficiency improvement by the nozzled configuration is 6.4%
achieved at low load 60HZ; at 48 kRPM, the nozzled volute shows constant
efficiency improvement compared with the nozzleless configuration at all
operation conditions. The maximum efficiency improvement is 11.1% achieved
at low load 20HZ.

(3) The shape of the hysteresis loops of the swallowing capacity for the nozzled
and nozzleless volutes are similar, indicating that the unsteady behaviour, in
terms of the wave action/filling & emptying effect, is not obviously influenced
by the existing of the nozzle vanes. Instead, the unsteady behaviour is strongly
influenced by the pulse frequency as well as the turbine load. Specifically, as
the load increases, the loops shift along the steady curves and expand
outwards and the covered area increase significantly.

(4) The results of the 1-D unsteady model show that the swallowing capacity curve
has significant influence on the unsteadiness of the turbine under the pulsating
condition. The slope of the curve, which reflects the throttling effect of the
turbine, is one of the key factors influencing the turbine unsteadiness. The
unsteadiness can be enhanced by reducing the slope of the curve via either
reducing the swallowing capacity (closing the nozzle) or increasing the turbine
mean load.

344
REFERENCES

[1] Baines N C., and Lavy M. Flow in Vaned and Vaneless Stators of Radial Inflow
Turbocharger Turbines. Pro. Inst. Mech. Eng., Turbochargers and
Turbocharging Conference, 1969.
[2] Bhinder F S. Investigation of Flow in the Nozzleless Spiral Casing of a Radial
Inward-Flow Gas Turbine. Pro. Inst. Mech. Eng., 1969.
[3] Rodgers, C. Radial Turbines-Blade Number and Reaction Effects. ASME Paper,
GT2000-456, 2000.
[4] Spence S W T., Rosborough R S E., Artt D., and et el. A Direct Performance
Comparison of Vaned and Vaneless Stators for Radial Turbines. Journal of
Turbomachinery, Vol. 129, 2007.
[5] Simpson A T., Spence S W T., and Watterson J K. A Comparison of the Flow
Structures and Losses within Vaned and Vaneless Stators for Radial Turbines.
Journal of Turbomachinery, Vol. 131, 2009.
[6] Romagnoli A., Martinez-Botas R. F., and Rajoo S. Turbine Performance
Studies for Automotive Turbocharger Part I: Steady Analysis. 9th Int. Conf. on
Turbocharging and Turbochargers, Instn. of Mech. Engrs, London, 2009.
[7] Romagnoli A., Martinez-Botas R. F., and Rajoo S. Turbine Performance
Studies for Automotive Turbocharger Part II: Unsteady Analysis. 9th Int.
Conf. on Turbocharging and Turbochargers, Instn. of Mech. Engrs, London,
2009.
[8] Rajoo S., Martinez-Botas R F. Variable Geometry Mixed Flow Turbine for
Turbochargers: An Experimental Study. International Journal of Fluid
Machinery and Systems, Vol.1, 2008.
[9] Rajoo S., and Martinez-Botas R F. Unsteady Effect in a Nozzled Turbocharger
Turbine. Journal of Turbomachinery, Vol.132, 2010.
[10] Karamanis N., Martinez-Botas R F. Mixed-flow Turbines for Automotive
Turbochargers: Steady and Unsteady Performance. Instn. of Mech. Engrs,
London, 2002.
[11] Simpson A T., Spence S W T., and Watterson J K. Numerical and Experimental
Study of the Performance Effects of Varying Vaneless Space and Vane Solidity
in Radial Turbine Stators. Journal of Turbomachinery, Vol.135, 2013.
[12] Costall A., Martinez-Botas R. Fundamental Characterization of Turbocharger
Turbine Unsteady Flow Behaviour. ASME paper, GT2007-28317, 2007.
[13] Copeland C. D., Newton P., and Martinez-Botas R. F. A Comparison of
Timescales with a Pulsed Flow Turbocharger Turbine. 10th Int. Conf. on
Turbocharging and Turbochargers, Instn. of Mech. Engrs, London, 2012.
[14] Costall A. A One-Dimensional Study of Unsteady Wave Propagation in
Turbocharger Turbines. PhD thesis, Imperial College London, 2007.

345
Experimental and numerical investigations
on an automotive turbocharger with a
transparent bearing section
W Köhl
Technische Universität Darmstadt, Institute of Structural Dynamics, Germany
M Kreschel, D Filsinger
IHI Charging Systems International GmbH, Germany

ABSTRACT

The dynamic behaviour of an automotive turbocharger supported with full floating


ring bearings is investigated experimentally and numerically. The rotor orbit on
compressor side and, after modifying the turbocharger’s housing, the rotational
speed of the floating ring are investigated by the measurement. To determine the
ring’s rotational speed, the housing of the journal bearing section has been rebuilt
employing a transparent synthetic material. The floating ring on the compressor
side is now visually accessible. Using a high-speed camera and a subsequent image
analysis in MATLAB, the communication holes around the circumference of the
floating ring are detected and the rotational speed can be calculated afterwards.

These experimental results are compared with a numerical rotor model. The model
simulates a flexible rotor with four translational degrees of freedom and a finite-
volume-model of the journal bearings. The numerical results show the
subsynchronous limit cycles, which can also be seen in the measurements. The
results of the numerically calculated rotational speed of the floating ring and the
calculated subsynchronous frequencies in comparison to the experimental results
conclude this paper.

1 INTRODUCTION

One option to support rotors in automotive turbochargers are floating ring bearings.
The floating rings divide the oil film in an inner and outer journal bearing. The
advantage of a floating ring bearing compared to a common bearing without
floating rings is presented for example in [12], [17] and [18]: the power loss due
to friction is reduced and the negative effects of typical phenomena such as oil-
whirl and oil-whip are reduced. Numerical simulations, however, show a possible
synchronisation of oil-whirl in both oil films which can cause inadmissible high rotor
amplitudes. The expression “total instability” is introduced in [16] to describe this
synchronisation.

These oil induced vibrations are often called constant tone phenomenon. While the
rotor vibrates in this state of motion, the subsynchronous frequency changes only
barely and the rotor amplitudes increase immediately. The rotor vibrations are
transmitted from the bearing system as an undesired sound radiation. In this case,
the surrounding parts of the rotor-bearing-system (e.g. turbine and compressor
housing, exhaust system) transmit the noise and in worst case depending on the
transmission function and their resonance frequencies respectively do behave like

_______________________________________
© The author(s) and/or their employer(s), 2014
349
amplifiers. This and other subsynchronous limit cycles occur over a wide speed
range of the turbocharger, as experiments presented in [5], [7], [13] and [14]
show. The whirl frequencies are a function of the rotational speeds of the rotor and
the floating ring. Experimental investigations to determine the rotational speed of
the floating ring for validating numerical models of the journal bearings and the
rotor model are therefore essential.

During the phase of designing the journal bearings, nonlinear models to calculate
the fluid forces and friction torques are used to predict critical whirl frequencies,
see [1-4], [9] and [12]. Numerical simulations for nonlinear rotordynamics of
automotive turbochargers in [4], [6] and [15-18] show the sensitivity of the
subsynchronous limit cycles related to the geometry of the floating ring bearing.
Basic numerically calculated bifurcation sequences are shown in [16] with varying
journal bearing geometry.

In general, the rotor is approximated with a finite-element-method (FEM) including


gyroscopic effects and a high number of nodes, i.e. a large number of degrees of
freedom, see e.g. [13-14]. In some cases the energy equation [3] is solved parallel
to the REYNOLDS equation to take the distribution of the oil temperature in the oil
film into account. This detailed modelling leads to long computational times (days
and weeks).

The intension of this paper is to present the experimental results for the rotational
speed of the floating ring for validating numerical models. Furthermore it will be
shown that a very simple rotor model with very much reduced numerical effort is
sufficient to show all relevant effects related to the journal bearings. The numerical
results are very well comparable to the experimental results.

In section 2, the experimental setup for the modified automotive turbocharger with
a transparent bearing section is presented. Section 3 introduces the numerical rotor
model of full floating ring bearings. Experimental results and a comparison to the
numerical simulations are discussed in section 4.

2 EXPERIMENTAL SETUP

To determine the rotational speed of the floating ring, the cast metal housing of the
bearing section is replaced by a transparent synthetic material. After designing a
3D-CAD model, a laser lithography process was used to produce the transparent
housing. The basic design of the transparent housing is almost identical to the
original casted housing. The journal bearing’s geometry and clearances, the
connections for the oil supply and the connecting geometry for the turbine and
compressor housing are not modified. Deviating from the original casted housing,
the surfaces are designed plainly to allow an undisturbed view on the floating ring.
The test rig is operated with cold air, which allows the operation of the turbocharger
without water cooling. Thus, the water core of the bearing housing was removed.

Figure 1 shows the assembled modified transparent bearing section. The test rig is
operated with cold gas at a pressure of about 5 bar. A transmission oil (ESSOLUBE
X2 20W) is used at room temperature (27 °C). The dynamic oil viscosity is known
as a function of the temperature by a previously conducted measurement on a
rheometer. Two perpendicular aligned displacement sensors measure the rotor
amplitudes of the compressor wheel. A rotational speed sensor is aligned to the
blades of the compressor wheel. To avoid aliasing errors, all signals are filtered with
an analogue low pass filter. The control system DSpace is used for data acquisition.

350
compressor turbine
wheel wheel

floating
ring

Figure 1: Transparent bearing section after machining

The area around the floating ring is optically observed by a high-speed camera with
monochrome images. The oil flow obstructs the view on the floating ring only
slightly. A frame rate of 32,000 fps was chosen to ensure that the communication
holes pass the recording area in sufficient time to detect every hole. This setting
allows to film an area of 256 x 12 pixels. A sufficiently strong light source has to be
used because of the very short exposure time (< 32 μs). An external trigger on the
camera allows the synchronisation between the data acquisition using DSpace and
the camera shots. All components of the test rig are listed in Table 1. Figure 2
shows the installation of the turbocharger with the transparent housing, the
equipment for measuring the rotor orbit and the high-speed camera.

Table 1: Components of the cold gas turbocharger test rig


Component Type Configuration
Turbocharger IHI Charging Systems Waste gate closed
International GmbH No compressor housing
High-Speed camera Photron: 32,000 fps
Fastcam Ultima 512 512 x 12 pixels
Light source Shot: cold light 2500 LCD
Displacement sensor -Epsilon, Eddy current
DT 110-T-S1-M-C3
DSpace Sampling frequency:
8192 Hz
Oil ESSOLUBE X2 20W 27°C ≈ 110 mPas

oil in air in light source rotational


speed sensor

floating displacement
oil out high-speed
ring sensors
camera

Figure 2: Installation of the turbocharger with the transparent housing in


the cold gas test rig, left: side view on the floating ring, right: axial view
on the high-speed camera and the sensors

351
In the post-processing, the camera images are analysed in MATLAB. The
determination of the floating ring’s rotational speed is basically equivalent with the
detection of the changing RGB-colour from black to white, as the drill hole leaves
and enters the observed area, respectively. The rotational speed of the floating ring
can now be calculated from the number of the drill holes and the time difference
between two detected drill holes. For a detailed description of this method see [8].

3 NUMERICAL MODEL

The experimentally determined waterfall diagrams and ring speed ratios presented
in section 4 allow to compare the dynamic behaviour of the rotor on compressor
side with numerical results. Further on, the ring speed ratio of the floating ring and
the rotor rotational speed can now be validated. The complexity of the model is
intentionally kept low to show basic effects related to the journal bearing such as
oil-whirl. Further, one can compare the numerical calculated critical threshold
speeds with the measured ones. Increasing the modelling depth step by step,
taking gyroscopic effects into account and using a finite element description of the
rotor are scheduled for future investigations.

Figure 3 left shows the axial view on the full floating ring bearing. The position of
the rotor is given in Cartesian coordiantes and ; the position of the floating
ring is given by and . The rotor and the floating ring are rotating with Ω
and Ω , respectively.

The later used transformation


= + , = + (1)

into complex coordiantes (see [2] and [10]) with as the complex number, reduces
the number of variables.

On the right side of Figure 3 one can see the free body diagram, which is later used
to formulate the equation of motion.

Ω

Ω ℎ , , ,

Figure 3: Cut through the full floating ring bearing and free body diagram

3.1 Full floating ring bearing


For describing the full floating ring bearing, the following assumptions are
introduced:
 incompressible fluid,
 constant and equal viscosities in both oil films,
 constant pressure distribution in radial direction,
 fully enclosed cylindrical bearings, no drill holes, no defects (e.g. wear),

352
 no angular misalignment of the rotor and of the floating ring inside the bearing,
 GÜMBEL boundary conditions as cavitation model, e.g. [2].

The geometric dimensions of the full floating ring bearing, as it is used for the
experimental investigations, require a finite bearing calculation. Therefore, the
REYNOLDS equations for the inner oil film

1 ℎ ℎ
ℎ + ℎ =6 (Ω + Ω ) +2 (2a)

and the outer oil film

1 ℎ ℎ
ℎ + ℎ =6 Ω +2 . (2b)

are used, to calculate the pressures ( , ) and ( , ) on the rotor and the
floating ring, respectively (e.g. [3] and [9]). In the equations (2) and are the
coordinates in axial and circumferential direction. The gap function is given by
ℎ( , ), is the radius of the bearing and the dynamic viscosity of the fluid. The
shear stresses ( , ) are calculated by the assumption of NEWTON’s hypothesis:

ℎ ℎ
= (Ω − Ω ) + , = Ω − . (3)
ℎ 2 ℎ 2

To solve the REYNOLDS equations (2), the finite-volume-method (FVM) is used to


discretize both oil films. After discretization of the equations (2) and (3), the
resulting forces , , and , , in Cartesian coordinates and friction torques , can
be calculated with

/ /

,, =− , cos , , ,, =− , sin , ,
/ /
/ (4)

, = , , .
/

3.2 Rotor-bearing system


A simple numerical model of the rotor system is used to calculate the rotor
amplitudes and the rotational speed of the floating rings. Figure 4 shows the
approximation of the rotor-bearing system. The rotor model contains four mass
points on a massless, elastic shaft. The masses and are the masses of the
real turbine wheel and the real compressor wheel. The masses , and , in
the position of the journal bearings are chosen in a way that the centre of gravity of
the rotor is maintained. The two floating rings are approximated by , and
, . Each mass point has a translational degree of freedom, which is included in
the complex state vectors

= , , , , , , = , , , . (5)

The two floating rings have additionally one rotational degree of freedom, which
leads to the angular matrix

=[ , , , ] . (6)

353
com-
turbine pressor

,
, , ,

conical mode S-shape


shape

Figure 4: Top: approximation of the rotor-bearing system, bottom: two


mode shapes: conical mode and first bending mode “S-shape”

The natural frequencies of the rotor are known by performed experimental modal
analysis on the freely supported rotor. After setting the data for the masses in the
matrix , the symmetrical stiffness matrix can be evaluated by matching
the calculated natural frequencies from solving the problem

det − =0 (7)

With

0 0 0
0 , 0 0
= , = (8)
0 0 , 0
0 0 0

with the frequencies known from the experimental modal analysis. The unsupported
rotor model without bearings contains two rigid body modes (cylindrical and conical
mode shapes) and two bending modes (U-shape and S-shape).

A CAD-computer model of the floating rings gives the mass moment of inertia ,
which is identical for both floating rings in the two journal bearings. Assuming
, = , = , the mass matrices for the floating rings can be formulated
as:

0 0
floating = , = . (9)
0 0

The complex state vector

= , , , (10)

the mass matrix

= (11)

and the stiffness matrix

354
= (12)

are now given.

Unbalances and the influence of gravity are taken into account using the vectors
∗ ∗
unbalance and gravity , respectively. From integrating the pressures and shear
stresses resulting from equations (2) and (3), one gets the acting forces and
friction torques of the journal bearing, which are all included in the complex vector

journal bearing after transformation respecting equation (1).

The equation of motion can therefore be written as:

∗ ∗ ∗
+ = + + . (13)

The ordinary differential equation system (ode) of 2nd order (13) is solved
numerically after transforming into an ode of 1st order and using MATLAB. The
MATLAB solver is designed for numerically stiff problems and a compromise between
calculation time and accuracy.

4 COMPARISON BETWEEN NUMERICAL AND EXPERIMENTAL RESULTS

The presented experimental results are taken for run downs from Ω , in 2.5 sec
to the rotational speed 0 rpm. Therefore, the rotor system is barely affected by
disturbances caused by the air flow on the turbine wheel. The radial load is also
reduced compared to a run up because the driving air on turbine side is no longer
present. The numerical simulations are performed for run downs under the same
conditions as mentioned above.

Figure 5 shows a comparison of the waterfall diagrams between experiment and


simulation for a run down. The diagrams are taken from the measured/simulated
amplitude | ̂ | of the compressor wheel. The amplitude | ̂ | is plotted in relation to
the maximal radial clearance ℎ , of the full floating ring bearing. The
experimentally determined rotational speed of the floating ring is added afterwards.

In Figure 5 top, the real rotor system shows two dominant subsynchronous limit
cycles. The amplitudes of the synchronous motion can hardly be recognized. The
first limit cycle in the speed range (0.1–0.4) Ω , is related to the frequency
≈ 0.48 (Ω + Ω ), which agrees well with the theoretical proposed frequency
for the oil-whirl of the inner film, see e.g. [12]. In the speed range (0.7–1) Ω , a
second limit cycle occurs. The frequency of this second limit cycle 2 changes
barely from 0.41 to 0.44 Ω . This limit cycle produces vibrations with a behaviour
of the already mentioned constant tone phenomenon. This oil-whirl state is followed
by a jump in the rotational speed of the floating ring. The amplitudes of the inner-
whirl are in the range of (1–2) ℎ , ; the resulting amplitudes for the constant
tone are up to 1.4 ℎ , . At 0.25 Ω , , the floating ring stops its rotation.

The actual bearing geometry of the used turbocharger as presented in Figure 2 is


not known without uncertainties. Uncertainties are machining tolerances and
material extension due to temperature effects. Therefore, a first simulation with the
nominal bearing geometry is performed and compared to the experiment. The
synchronous motion due to unbalance is negligible as mentioned above. Therefore,
unbalance is not taken into account in the simulation. A dynamic viscosity for the
inner and outer oil film = = (27℃) ≈ 110 mPas is used.

355
Calculated waterfall diagram from measurements with the transparent
bearing housing
maybe bending
constant tone
2.5 mode?
inner oil-whirl
synchronous
2 ȳிோ Ȁȳோǡ௠௔௫
ȁ‫ݎ‬Ƹ஼ ȁ௘௫௣ Ȁ݄଴ǡ௠௔௫ 

ͲǤͶͺሺȳ ൅ ȳ ሻȀȳ ோ ிோ ோǡ௠௔௫

1.5
1 1
0.75

௠௔௫
ȳோ Ȁȳோǡ௠௔௫
1
0.5
0.5
0
0.25
5
0 0
1 0.75 0.5 0.25 0
ȳிி் Ȁȳோǡ௠௔௫
Calculated waterfall diagram from numerical model with nominal bearing
geometry
outer oil-whirl
oil wh
whirl
2.5 inner oil-whirl
constant tone
2 ͲǤͷȳிோ Ȁȳோǡ௠௔௫
ȁ‫ݎ‬Ƹ஼ ȁ௡௨௠ Ȁ݄଴ǡ௠௔௫ 

ͲǤͷሺȳோ ൅ ȳிோ ሻȀȳோǡ௠௔௫ ȳிோ Ȁȳோǡ௠௔௫

1.5 1
0.75
5

ȳோ Ȁȳோǡ௠௔௫
1
0.5
synchronous
0.5
0.25
5
0 0
1 0.75 0.5 0.25 0
ȳிி் Ȁȳோǡ௠௔௫
Calculated waterfall diagram from numerical model using a modified
geometry
first bending
constant tone mode

conical mode
2.5
inner oil-whirl
2
ȁ‫ݎ‬Ƹ஼ ȁ௡௨௠ Ȁ݄଴ǡ௠௔௫ 

ͲǤͷሺȳோ ൅ ȳிோ ሻȀȳோǡ௠௔௫ ȳிோ Ȁȳோǡ௠௔௫


1.5 1
0.75
5
ȳோ Ȁȳோǡ௠௔௫

1
0.5
0.5 synchronous
conical mode 0.25
5
0 0
1 0.75 0.5 0.25 0
ȳிி் Ȁȳோǡ௠௔௫

Figure 5: Calculated waterfall diagrams, top: experiment with transparent


bearing housing, middle: numerical model using the nominal geometry of
the bearing, bottom: numerical model using a modified geometry
(increased inner clearance , and outer length )

356
Figure 5 middle shows the waterfall diagram for a simulated run down using the
nominal bearing geometry. One can recognize the oil-whirl of the inner film and the
constant tone. The amplitudes of the inner oil-whirl with (0.5–1) ℎ , are lower
than the measured ones, but they are comparable to the experiment. The resulting
amplitudes of the constant tone are very low. The frequency of the constant tone
alters from 0.56 to 0.58 Ω , . One can recognize a third limit cycle at 0.8 Ω ,
with nearly half the frequency of the floating ring’s rotational speed ≈
0.5 Ω . This limit cycle belongs to the oil-whirl of the outer film and does not
appear in the experiment. In general, the calculated rotational speed of the floating
ring is higher than the experimentally determined floating ring speed.

To adjust the numerical results closer to the experimentally determined rotor


movement, the stiffness and the damping of the inner and outer oil film have to be
modified. Further, the floating ring speed has to be reduced. This can be reached
by changing the bearing geometry. It is known from parameter studies that
increasing the radial clearance of the inner bearing ℎ , leads to a reduced stiffness
of the inner oil film. This favours the occurrence of the constant tone phenomenon
and the rotor can enter into this state at lower rotor speeds. To reduce or even to
supress the outer whirl in the numerical results, the damping of the outer bearing
has to be increased. This can be reached by increasing the outer length of the
bearing. One can find approximate formulas for the ring speed ratio in [3] and [12].
By increasing the length , the speed ratio will also be decreased. A second
numerical model which uses an increased inner radial clearance ℎ , and an
increased outer length is introduced to adjust the subsynchronous motion states
closer to the experimental results. The numerical results are presented in Figure 5
bottom.

The outer oil-whirl does not appear anymore within the investigated speed range
after modifying the geometry of both journal bearings as mentioned above. The
inner oil-whirl is followed by amplitudes at 0.75 ℎ , , which is closer to the
experiment. The rotor enters the constant tone state at 0.7 Ω , and rotates with
amplitudes at 1.6 ℎ , , which has a sufficient correlation to the experimental
results. However, the numerical model shows a jump of the compressor’s amplitude
before leaving the constant tone in the range (0.55–0.65) Ω , . During this
speed range, the rotor vibrates in the first bending mode shape, see Figure 4
bottom right. While in the speed range of the inner oil-whirl and the constant tone
respectively, the rotor gets into a conical state of motion, see Figure 4 bottom left.
The frequency 2 of the constant tone is better estimated with 0.49 to
0.52 Ω , . The rotational speed of the floating ring is still calculated too high.

Comparison between the experimental run down and the numerically simulated run
down is carried out using the speed ratio of the floating ring speed and the rotor
speed. After determining experimentally the rotational speed of the floating ring
using the high-speed camera, the ring speed ratio results in the relative speed
range Ω , /Ω ≈ 0.1 … 0.2, see Figure 6. The calculated ring speed ratios for the
numerical simulations yield Ω , /Ω ≈ 0.34 … 0.49 and Ω , /Ω ≈
0.26 … 0.41 for the numerical model with nominal bearing geometry and the model
with a modified geometry, respectively. The ring speeds from both numerical
models are overestimated; model 2 provides a better result.

357
0.5 constant tone + outer oil-whirl numerical model
nominal bearing geometry
0.4
Ω𝐹𝑅 / Ω𝑅
0.3
constant tone
inner oil-whirl numerical model
0.2 modified bearing geometry
0.1 experiment

0
1 0.9 0.8 0.7 0.6
0.5 0.4 0.3 0.2 0.1 0
Ω𝑅 / Ω𝑚𝑎𝑥
Figure 6: Ring speed ratios for the experimental determined rotational
speed of the floating ring and the numerically calculated ring speeds

Due to a smaller inner clearance, the temperature in the inner film of the real
system increases faster. Therefore, the viscosity in the inner film gets smaller
( , < , ) which is not taken into account for the presented numerical models.
The friction torque, which drives the floating rings, is smaller in reality than
predicted numerically. Further work will improved the adaptation of the numerically
calculated floating ring speed. Experiments for stationary processes and different oil
temperatures will provide more data for numerical validation.

5 CONCLUSION

The use of a transparent bearing housing and a high-speed camera allows to


measure the rotational speed of the floating ring on compressor side. With this
method, an important dataset to validate numerical models for rotor systems with
floating ring bearings has been created. A simple numerical model of the rotor-
bearing-system is presented, that allows to detect the basic effects related to the
full floating bearing. General agreement with the measurement was achieved,
although the complexity of the real investigated system exceeds that of the
numerical rotor-bearing-model by a multiple. Investigations to determine the
natural frequencies of the unsupported rotor and the oil viscosity with a rheometer
improved the quality of the rotor model. The numerically calculated waterfall
diagrams show, that the threshold speed for entering into the constant tone
phenomenon is well estimated. Furthermore, the amplitudes of the orbit of the
compressor wheel during the oil-whirl and the constant tone are comparable to the
measured amplitudes. However, the first numerical model, that uses the nominal
geometry of the bearings, shows an oil-whirl of the outer oil film, which has not
been seen in the experiments. Adjustment of the model’s geometric parameters
allowed a better match with the measurement results.

In general, the numerically calculated ring speed ratios are overestimated


compared to the experiment. This can be improved by taking different viscosities
for both oil films into account. The usage of a constant viscosity in the inner oil film
that satisfies < will improve the model of the floating ring bearings.

6 ACKNOWLEDGEMENTS

The author thanks Prof. Dr.-Ing. Richard Markert for his guidance and contribution
to this work and to the cooperation with IHI Charging Systems International GmbH.

358
7 REFERENCE LIST

[1] ARGHIR, M.; ALSAYED, A.; NICOLAS, D.: The finite volume solution of the
Reynolds equation of lubrication with film discontinuities. In: International
Journal of Mechanical Sciences 44 (2002), 2119-2132. doi: 10.1016/S0020-
7403(02)00166-2
[2] GASCH, R.; NORDMANN, R.; PFÜTZNER, H: Rotordynamik. 2. Auflage. Springer-
Verlag Berlin – Heidelberg – New York, 2006.
[3] HORI, Y.: Hydrodynamic Lubrication. Springer-Verlag Tokyo, 2006.
[4] KAMESH, P.: Oil-Whirl instability in an Automotive Turbocharger. University of
Southampton, Faculty of Engineering and the Environment, Institute of
Sound and Vibration Research, PhD-Thesis.
[5] KIRK, R.; ALSAEED, A.; LIPTRAP, J. et al.: Experimental test results for vibration
of a high speed diesel engine turbocharger. In: Tribology Transactions 51
(2008), 422–427. doi: 10.1080/10402000801911853
[6] KIRK, R.; ALSAEED, A.; GUNTER, E.: Stability analysis of a high-speed
automotive turbocharger. In: Tribology Transactions 50 (2007), 427–434.
doi: 10.1080/10402000701476908
[7] KIRK, R.; KORNHAUSER, A.; STERLING, J.; ALSAEED, A.: Turbocharger On-Engine
Experimental Vibration Testing. In: Journal of Vibration and Control 16
(2010), 343–355. doi: 10.1177/1077546309103564
[8] KÖHL, W.; KRESCHEL, M.; FILSINGER, D.: Experimental determination of the
rotational speed of floating rings in a transparent bearing housing of a
turbocharger, In: 8. VDI-Fachtagung Schwingungen in Antrieben 2013, VDI-
Berichte 2197, 105 – 116, VDI-Verlag GmbH Düsseldorf, 2013, ISBN 978-3-
18-092197-6
[9] LANG, O; STEINHILPER, W.: Gleitlager – Konstruktionsbücher Band 31.
Springer-Verlag Berlin – Heidelberg – New York, 1978.
[10] MARKERT, R.: Skript zur Vorlesung Rotordynamik. 2. Auflage. Technische
Universität Darmstadt, Fachbereich Maschinenbau, Fachgebiet
Strukturdynamik, 2011.
[11] MUSZYNSKA, A: Rotordynamics. CRC Press, 2005.
[12] NGUYEN-SCHÄFER, H.: Rotordynamics of Automotive Turbochargers. Springer
Heidelberg New York Dordrecht London, 2012.
[13] SANANDRES, L.; RIVADENEIRA, J.; GJIKA, K. et al.: A virtual tool for prediction of
turbocharger nonlinear dynamic response: validation against test data. In:
Journal of Engineering for Gas Turbines and Power 129 (2006), 1035-1046.
doi: 10.1115/1.2436573
[14] SANANDRES, L.; RIVADENEIRA, J.; GJIKA, K. et al.: Rotordynamics of small
turbochargers supported on floating ring bearings—highlights in bearing
analysis and experimental validation. In: Journal of Tribology 129 (2007),
391-397. doi: 10.1115/1.2464134
[15] SCHWEIZER, B.: Oil whirl, oil whip and whirl/whip synchronization occurring in
rotor systems with full-floating ring bearings. In: Nonlinear Dynamics 57
(2009), 509-532. doi: 10.1007/s11071-009-9466-3
[16] SCHWEIZER, B.: Total instability of turbochargers rotors – Physical explanation
of the dynamic failure of rotors with full-floating ring bearings. In: Journal of
Sound and Vibration 328 (2009), 156-190. doi: 10.1016/j.jsv.2009.03.028
[17] SCHWEIZER, B.: SIEVERT, M.: Nonlinear oscillations of automotive turbocharger
turbines. In: Journal of Sound and Vibration 321 (2009), 955-975.
doi: 10.1016/j.jsv.2008.10.013
[18] TIAN, L.; WANG, W. J., PENG, Z. J.: Dynamic behaviours of a full floating ring
bearing supported turbocharger rotor with engine excitation. In: Journal of
Sound and Vibration 330 (2011), 4851-4874. doi: 10.1016/j.jsv.2011.04.031

359
An analytical investigation of
turbocharger rotor-bearing dynamics
with rolling element bearings and
squeeze film dampers
A Ashtekar
Cummins Turbo Technologies, USA
L Tian, C Lancaster
Cummins Turbo Technologies, UK

ABSTRACT

The objective of this investigation is to examine the dynamics of a turbocharger


supported by a deep groove or angular contact ball bearing and a squeeze film
damper. In this novel approach a six degree of freedom 3D discrete element
bearing model was interlaced with a first principle squeeze film damper model to
determine the combined stiffness and damping of the turbocharger support. The
combined model accounts for the current and the past dynamic states of the
system to provide a more accurate support behavior than the current simplified 2-D
bearing models used for rotor dynamic analysis. In addition, the Reynolds equation
is iteratively solved for the squeeze film damper model to determine the damper
behavior while accounting for side leakages. This allows for examining any shape
or size of dampers. The combined model was then used to determine the dynamics
response of the turbocharger by coupling it with a traditional quasi-static model as
well as a time dependent rotor dynamic models. The effect of bearing component
(inner race, outer race, cage and roller) defects on support stiffness and excitation
will be examined. The damper will affect not only the turbocharger dynamics but
also the bearing dynamics, affecting the bearing life.

1 INTRODUCTION

Turbochargers have commonly been equipped with journal bearings to support the
turbines and rotor assembly. However, ball bearings have become popular as a
replacement for journal bearings in turbochargers. Wang (1), in his review of
ceramic bearing technology, points out that the hybrid ceramic bearing can provide
better acceleration response, lower torque requirement, lower vibrations and lower
temperature rise than journal bearings. Hybrid ceramic ball bearings contain steel
inner and outer races, ceramic balls and usually a machined cage. Ceramic balls, as
compared to their steel counter parts, are lighter, smoother, stiffer, harder,
corrosion resistant, and electrically resistant. These fundamental characteristics
allow for a wide range of performance enhancements in bearing rotor system.
Ceramic balls are particularly well suited for use in harsh, high temperatures and/or
corrosive environment. Therefore, hybrid ceramic bearings are ideal for
turbocharger applications. Miyashita et al. (2), Keller et al. (3) and Tanimoto et al.
(4) have employed ball bearings in small, automotive turbochargers. However,

_______________________________________
© The author(s) and/or their employer(s), 2014
361
challenges still remain for high speed, high output turbochargers which demand
large bore bearings operating at DN numbers over 2 million. As the bearing size
increases, the dynamics of the bearing rotor system becomes critical for
comprehensive design and satisfactory operation of the turbocharger.

Investigators have attempted to analytically analyze the dynamics of turbocharger


rotor system. San Andrés et al. (5,6,7) has presented comprehensive models to
predict turbocharger dynamics. Inclusion of a complete fluid-film bearing model
provided an insight into the effects of bearing dynamics on the dynamics of a
turbocharger. Bou-Said et al. (8) also investigated the rotor dynamics of a
turbocharger with linear and non-linear aerodynamic bearing models. Pettinato et
al. (9) demonstrated the advantages of such turbocharger rotor dynamic models by
employing them to improve the design of bearings used in a turbocharger. Bonello
(10) implemented non-linear model to study the dynamics of turbocharger on full
floating and semi-floating ring bearings. However, most of the work in turbocharger
rotor dynamic models has been concentrated on turbochargers with journal
bearings. Therefore these models are unable to predict the rotor dynamics of
turbochargers which use rolling element bearings. Nevertheless, investigators have
attempted to develop analytical models to study the dynamics of simple rotor
systems with rolling element bearings. Gupta (11-13) was among the first to
present a three dimensional bearing dynamic model. The model developed was
capable of analyzing motion of all bearing components. Meyer et al. (14) introduced
the effects of defects on bearing and demonstrated the vibrations patterns
associated with the defects. Saheta et al. (15) and Ghaisas et al. (16) presented a
six degree of freedom, fully dynamic discrete element model. Their models consider
bearing components as sections of spheres and cylinders, which significantly
reduced the computational effort associated with bearing dynamic modeling.
Sopanen et al. (17, 18) developed a bearing model which included the effects of
inclusions. However in their analysis, cage dynamics and centrifugal loads were
ignored. Ashtekar et al. (19, 20) developed a six degree of freedom bearing model
which included the effects of bearing surface defects. In general, the previous
investigators concentrated on the bearing dynamics and ignored the complicated
interaction of the roller bearing with the shaft/rotor system. However, for a
complete understanding and examination of high speed, high output turbochargers
it is critical to combine the effects of the bearing and shaft/rotor dynamics. In high
speed applications, the rotor undergoes various mode shapes resulting in complex
motion of bearing rotor system. Lim et al. (21) and Hendrikx et al. (22) developed
a bearing model including the effects of rotor flexibility; however they neglected the
effect of bearing cage on the dynamics of the system. Tiwari (23, 24) considered
the effects of imbalance and bearing preloading on the rotor dynamics, however, a
simplified ideal bearing model was considered and rotor was assumed to be rigid.
Prenger (25) presented a bearing model capable of modeling tapered roller
bearings and angular contact bearings. Prenger’s model included the effect of
flexible shafts; however, only simple shaft models were considered and the model
was unable to handle high speed applications. BEAST software developed by Stacke
et al (26) is known to consider rotor flexibility; however, neither the model nor the
results are available in public domain.

In this investigation a model was developed to represent the turbocharger bearing


rotor system. The model combines a discrete element bearing model and a flexible
rotor model to simulate the dynamics of the bearing rotor system. The model was
then used to investigate the motion of each bearing components and determine the
forces and deflection of the rotor as a function of various operating conditions. The
results from the model were used to investigate the bearing performance at various
preloads, rotor imbalances and operating speeds.

362
2 MODEL DESCRIPTION

A ball bearing consists of an inner race, outer race, rolling elements (balls) and a
cage which separates the balls. These bearing components interact with each other
directly or indirectly, affecting the motion and forces occurring between them. The
turbocharger rotor is supported by the inner race and thus its motion and forces are
also affected by the dynamics of the turbocharger rotor. As these motions and
forces are eventually transmitted from the inner race to all other bearing
components, the turbocharger rotor affects the dynamics of all bearing
components. Similarly, any dynamic instability within the bearing is transmitted to
the turbocharger rotor. In this study, the analytical investigation includes a bearing
dynamic model which interacts with a flexible turbocharger rotor model to predict
the dynamics of rotor system.

2.1 Dynamic bearing Model


A key aspect of modeling the bearing dynamics with Discrete Element Method is
obtaining the forces and moments acting on the bearing components. In the
current model, the gravitational forces, contact forces and rotor interaction forces
are considered as a part of the analysis.

Rolling element contact forces are considered when balls are in contact with other
bearing components. Although bearing component surfaces deform to some
degree when in contact, these deformations are typically very small in comparison
to the ball’s characteristic length. Hence, in the contact model the detailed
deformations of the contacting surfaces are ignored and instead the two contacting
surfaces are allowed to overlap slightly. The degree of overlap is then used to
determine the contact forces acting on the bearing components. To simplify the
overlap calculations, bearing components are assumed to be made of simplified
geometry consisting of sections of sphere and cylinders. The overlap, , between
the elements is given by;

=( + )−| − |

Where, and are the radius of the bodies and and are the position
vectors of the respective bodies. It is possible to consider other shapes in the
simulations; however, the contact detection schemes become more computationally
intensive (Ting (27), Ting et al. (28), Matuttis et al. (29)). The normal contact
force can be determined using the overlap and Hertzian force-deflection relationship


=

Where, K is the Hertzian stiffness coefficient. This approach of calculating normal


contact force is much simpler and less computationally intensive than the method
described by Gupta (30).

The Hertzian stiffness, K for two general, non-conformal contacting solids is given
by Hamrock (31)


2
= ⁄
ℑ 9

Where R is the curvature sum given by,

= + ; = + ; = +

363
The values of and are curvatures of the body in X and Y planes respectively.
is the effective elastic modulus obtained from the elastic properties of bodies
in contact and is given by,

2
=
1− 1−
+

Where , , , are the modulus of elasticity and Poisson’s ratio for the two
bodies. Parameters , , and ℑ require iterative calculations, however, current
study uses the approximate solution provided by Hamrock (31)


= ; = ;ℑ= + ln ; =1+ ; = −1

In addition to a normal force, a tangential force exists at the point of contact


between the ball and race. This tangential force is determined using a traction
model, the relative tangential velocity at the point of contact, and the normal force
at the contact. In this investigation, the Kragelskii’s (32) model is used

( | |)
= ( + | |) +

, is related to slip velocity, , which is the difference in instantaneous velocities


of bodies in contact.

Here, values of A, B, C, and D were calculated using the method used by Gupta
(33). The tangential friction force is then given by

=−
| |

Evaluation of the tangential friction forces at the contact can be quite involved
because of the variations in local slip velocities from point to point in the contact
ellipse. However, as pointed out by Gupta (30, 34), for most bearing applications
the contact ellipse is sufficiently narrow along the direction of rolling so that the
variations in the slip velocity and hence friction force along the semi minor axis can
be neglected. Thus the total friction force can be evaluated by integrating the
friction forces along the major axis of contact ellipse (30, 34). The resulting
tangential force also creates a moment about the ball center and a moment about
the center of contact ellipse. The moment about the ball center results in rolling
motion while the moment about contact ellipse center causes the ball to spin.

The resulting contact forces and moments act equally but in the opposite directions
on both of the bodies in contact. After calculating the total force and moments
acting on the bodies, Newton’s second law is used to calculate the linear and
angular acceleration of the bodies. The accelerations are integrated with respect to
time to obtain velocities and displacements in linear and angular directions. Each
body has 6 DOF and thus each body is associated with 6 equations that are
integrated. System presented in this paper has two bearing and each bearing has
15 such bodies. The above procedure is repeated at each time step using the new
component states until some end condition, typically maximum time is reached.

2.2 Squeeze Film Damper


Hamrock’s (31) solution of Reynolds equation is used to calculate the reaction
forces due to squeeze film damper on the outer race. For a long damper

364
assumption, (l/d > 4), the side leakages can be neglected and the Reynold’s
equation reduces to

ℎ = −12

This can be solved to get the relationship

12 ( + )
=−
(1 − ) /

The damping coefficient can be expressed as,

12 ( + )
=
(1 − ) /

Where, is the absolute viscosity of the oil, is the radius of the bearing outer
race, is the outer race velocity along Z axis, c is the clearance in the squeeze
film damper and z is the position of the outer race CG.

The above relationship is suitable for any bearing with l/d ratio greater than four.
For these bearings, the side leakages can be ignored so that the infinitely long
bearing assumption holds true. However, for shorter bearings, consider the
Reynolds equation with side leakages.


ℎ + ℎ = 12

For a bearing,

=r and ℎ = (1 − )

Therefore, equation reduces to

ℎ + ℎ = −6

= / Thus,

+ = −12
4

This equation does not have an exact solution and thus needs to be solved
iteratively. For a given and c, a solution is evaluated for a bearing position
defined by h (or z, y in this case). Iterative solutions were obtained for a range of l
and c values to generate a database of SFD. Table 1 shows the range of l and c
values considered for the study.

Table 1: Range of l and c values

L (mm) 2 3 6 10 100

C (µm) 20 30 50 100 200

365
Figure 1 shows the database plots. Intermediate values were obtained using linear
interpolation. Figure 2 shows the implementation of database into DBM. Please
note that the reaction force is opposite to the direction of OR velocity. Similarly,
the reaction force along Y axis is also calculated. Both these reaction forces are
added to the total forces acting on the OR discrete element. In addition, to include
the effects of the anti-rotation pin, all outer race rotational degrees of freedom
were constrained to be fixed. For each case, DBM was run and the motion of Inner
Race, Outer Race, and Reaction Force at SFD and Damping Coefficient was
recorded. Lower IR motion and reaction forces are primary benefits of well-tuned
SFD.

From Figure 3 to Figure 5 it can be seen that the damping coefficient is sensitive to
length of the bearing and the clearance. Higher damping reactions were observed
at very small lengths and damping reactions reduced as length increased.
However, after an optimum point the damping reactions shot up as length was
increased to approach long SFD assumption. IR motion continued to reduce as the
length of bearing increased. This is primarily due to the geometrical constraints
due to larger contact surface between bearing and housing. Increased clearance
had negative effect on reaction forces as well as IR motion.

Figure 1: Database for SFD

Figure 2: Short SFD model in DBM

366
Figure 3: Effect of length on reaction forces

Figure 4: Effect of length on damping

Figure 5: Effect of clearance on reaction forces

367
2.3 Dynamic Bearing Rotor Model (DBRM)
Figure 6 depicts a schematic representation of the bearing rotor system as
represented by the DBRM.

Figure 6: Dynamic Bearing Rotor Model

In this investigation, the DBM and damper model were used to determine the
bearing response which was passed on to the flexible rotor dynamic model (FRM),
which consists of 26 nodes as shown in Figure 7. From Node 1 to Node 18 is the
turbocharger rotor modeled by Timoshenko beam elements. The turbine disk is
modeled as beam elements from Node 18 to Node 26 with four steps. The diameter
of each beam section is determined from the mass of the corresponding section of
the turbine wheel, and the differences of polar moment of inertia and transverse
moment of inertia between them are compensated by the additionally lumped
inertia properties to the center of each beam section. As for the compressor disk, it
is sectioned to five parts, and their mass and inertia properties are all lumped to
the corresponding nodes from Node 2 to Node 6. The four black arrows show the
locations of unbalance. The magnitude of the unbalance is 1.5 g.mm per plane,
though the unbalance at the compressor end is out-of-phase to the unbalance at
the turbine end, as indicated by the directions of those arrows. Due to the highly
stiff nature of the ordinary differential equations (ODEs) representing the equations
of motion of the rotor-bearing system, the implicit MATLAB® routine ode15s© is
employed to solve those ODEs with varying time steps.

Figure 7: 26 Node Dynamic Bearing Rotor Model

The two pairs of red triangles mark the locations of center of gravity of the ball
bearing inner race. Through those two interface nodes, the FRM model is connected
to the DBM, which governs the dynamics of two ball bearings. Thus the dynamic
response is passed from one model to other. Figure 6 depicts these interface point
interactions as two headed arrows indicating that the exchange of dynamic

368
response occurs from both the sides, namely, DBM and the rotor model. The two
bearings have a single piece outer race. Therefore, the outer races of the two
DBMs, each representing one of the bearing, are rigidly linked to each other. Figure
6 illustrates these linkages shown as lines. Finally the single piece outer race of the
DBMs is attached to the ground through a spring-damper arrangement representing
the squeeze film damper.

Also, to include the effects of the anti-rotation pin, all outer race rotational degrees
of freedom were constrained to be fixed. The two models, DBM and FRM, run
parallel, communicating with each other at each time step. Any motion and/or
forces due to turbocharger rotor flexibility affects the dynamics of all the bearing
components and similarly dynamic response of bearing components affect the
dynamics of the entire turbocharger bearing rotor system.

3 ADDITIONAL RESULTS AND OBSERVATIONS

3.1 Model Interaction Study


To allow for the union of an explicit Bearing model with an implicit rotor model,
three different methods were used.

In the first method the REB stiffness was evaluated using the DBM. The model was
subjected to a varying IR motion and the reaction forces from the model were
compiled to determine the stiffness matric of the bearing. This matrix can be used
as support stiffness for any rotordynamic model of choice to investigate the
turbocharger rotordynamics in presence of the REB. The table shows the matrix for
a turbocharger bearing.

Table 2: Bearing Stiffness matrix

Bearing Direction Bearing Stiffness

Fx 1.66E+06 -5.64E-07 2.86E-06 -9.62E-09 -1.23E+04

Fy -5.60E-07 1.66E+06 2.47E-06 1.23E+04 9.60E-09

Compressor Fz 2.86E-06 2.46E-06 1.43E+06 1.79E-08 -2.88E-08

Myz -9.43E-09 1.23E+04 1.79E-08 1.13E+02 1.03E-10

Mzx -1.23E+04 9.52E-09 -2.88E-08 1.03E-10 1.13E+02

Fx 1.67E+06 2.38E-06 3.71E-06 1.82E-08 -1.24E+04

Fy 2.37E-06 1.67E+06 -1.05E-05 1.24E+04 -1.82E-08

Turbine Fz 3.70E-06 -1.05E-05 1.43E+06 -7.48E-08 -3.58E-08

Myz 1.82E-08 1.24E+04 -7.48E-08 1.13E+02 -1.69E-10

Mzx -1.24E+04 -1.82E-08 -3.59E-08 -1.70E-10 1.13E+02

This is a simple and efficient approach to be incorporated in any rotordynamic


model. However, this method oversimplifies the REB and ignores the internal
dynamics and instabilities of the REBs. The results from this method are
nonetheless useful to analyze basic steady state dynamics. Comparison with other
methods shows that the basic dynamics can be evaluated to acceptable accuracy.

369
In the second method, the model was used with a quasi-static approach model.
The rotordynamic model is implicit and passes on the node state to the bearing
model. The explicit DBM model is ramped up to the state and allowed to reach a
steady state. The forces and displacements are passed back to the rotor model and
the simulation continues. This method does not completely account for the past
dynamics of the system but offers an REB solution that is analogues to journal
bearing models. This method produces stable solutions and accounts for bearing
internal dynamic response. However, the simulation resources and time required
are significant.

In the third method the DBM was run in parallel to the rotordynamic model. Each
time the rotor model passes on the states of the node, the past REB state is used
as the starting point and the DBM explicit model is ramped on from the old node
state to the new one. This allows for including the transient effects in the model.
Rotor transients have a significant effect on bearings dynamics. However, these
models do have the possibility of diverging solutions in rotordynamic models.

The results for each of these methods were compared against each other and
evaluated for a range of imbalances. The imbalance affects the rotordynamics as
well as has a significant effect on the bearing dynamics.

Figure 8: Comparison of REB Models

3.2 Preloading
Angular contact bearing are commonly preloaded, however, Hagiu (35) has
demonstrated that wrong preloading will cause considerable reduction in bearing
life. Figure 9 shows ball loads for two DBRM conditions, one which has preloading
and one without preloading. Both of these cases are operating at the speed of
50000 rpm with 10 gm-mm imbalance. Please note the increased force fluctuation
for the case of unloaded bearing. The results also demonstrate that occasionally the
ball-race load becomes zero indicting loss of ball-race contact. The loss of load
between the ball and inner race can cause ball sliding and skidding. The results
demonstrate the significant effect of wrong bearing preloading in turbocharger. It is
also to be noted that excessive preloading can lead to premature fatigue failure of
the bearing. The effects of these bearing instability is examined on the
turbocharger.

370
Figure 9: Unloaded vs Preloaded Bearing Forces

3.3 Bearing Defects


REB component configuration is of the planetary type with IR being the sun and the
rolling elements being the planets. The rolling elements roll over the IR as well as
the OR and drive the cage at the same time. Due to continuous but periodic nature
of this interaction, a defect on any of the bearing components results in periodic
excitation in the system.

Ball Defect Frequency = (Pd/(2*Bd)) * (N/60) * (1 – (Bd/Pd*cosØ)^2)

Cage defect frequency = N/120 * (1 – Bd/Pd*cosØ)

OR defect frequency = Nb/2 * (N/60) * (1 – (Bd/Pd*cosØ))

IR defect frequency = Nb/2 * (N/60) * (1 + (Bd/Pd*cosØ))

Where, Pd = Pitch Diameter, Bd = Ball Diameter, Nb = Number of Balls, N =


Speed, Ø = contact angle of the angular contact bearing.

These equations provide a good guidance; however, they ignore the 3D nature of
the bearings. i.e. there is no guarantee that the ball will pass over the defect each
time. The ball track may be wide and might miss the defect in a periodic manner.
Thus, a defect was introduced in the DBM using the defect models by Ashtekar et
al. (19,20) and their effects on the rotordynamics were observed. This allows for a
realistic simulation of the defects and their effects on turbocharger.

4 SUMMARY AND CONCLUSIONS

Investigation into replacing journal bearings of a high speed turbocharger with


hybrid ceramic ball bearings requires a detailed analytical model of bearing rotor
dynamics. For the analytical investigation rolling element bearing demands
significant speed and accuracy of contact force calculations. An analytical model
that meets these demands has been developed. A coupled dynamic model was
developed for the ball bearing rotor systems. The model combines a discrete
element bearing model and a flexible rotor model. A squeeze film iterative model
was also developed to model the squeeze film dampers required to counter the high
stiffness of the REB. The analytical model was used to investigate the different
approaches to model the REB into the system. The model differences were
highlighted under imbalance conditions to demonstrate the dynamics ignored by

371
simplified REB models. The model was also used to demonstrate the effects of
preloading on the turbocharger dynamics. The analytical model was also used to
gain knowledge of effects of the REB defects on turbocharger dynamics.

REFERENCE LIST

(1) Wang L., Snidle R., and Gu L., 2000, “Rolling contact silicon nitride bearing
technology: a review of recent research,” Wear, 246(1-2), pp. 159-173.
(2) Miyashita, K., Kurasawa, M., Matsuoka, H., Ikeya, N., and Nakamura, F.,
1987, “Development of High Efficiency Ball-Bearing Turbocharger,” SAE Paper
No. 870354.
(3) Keller, R., Scharrer, J., and Pelletti, J., 1996, “Alternative Performance
Turbocharger Bearing Design,” SAE Paper No. 962500.
(4) Tanimoto K., Kajihara K., and Yanai K., 2000, “Hybrid ceramic ball bearings
for turbochargers,” SAE Paper 2000-01-1339, pp. 1-14.
(5) San Andres L., Rivadeneria J. C., Chinta M., Gjika K., La Rue G., 2007,
“Nonlinear Rotordynamics of Automotive Turbochargers: Predictions and
Comparisons to Test Data,” J. Eng. Gas Turbines Power, 129(2), pp. 488-494.
(6) San Andres L., Rivadeneria J. C., Gjika K., Groves G., La Rue G., 2007, “A
Virtual Tool for Prediction of Turbocharger Nonlinear Dynamic Response:
Validation Against Test Data,” J. Eng. Gas Turbines Power, 129(4), pp. 1035-
1047.
(7) San Andres L., Rivadeneria J. C., Gjika K., Groves G., La Rue G., 2007,
“Rotordynamics of Small Turbochargers Supported on Floating Ring Bearings-
--Highlights in Bearing Analysis and Experimental Validation,” J. Tribol.,
129(2), pp. 391-398.
(8) Bou-Said B., Grau G., Iordanoff J., 2008, “On Nonlinear Rotor Dynamic Effects
of Aerodynamic Bearings With Simple Flexible Rotors,” J. Eng. Gas Turbines
Power, 130, 012503.
(9) Pettinato B. C., DeChoudhury P., 2003, “Rotordynamic and Bearing Upgrade
of a High-Speed Turbocharger,” J. Eng. Gas Turbines Power, 125, pp. 95-102.
(10) Bonello, P., 2009, “Transient modal analysis of the non-linear dynamics of a
turbocharger on floating ring bearings,” Proc. I. Mech. E. Part J: J.
Engineering. Tribology, 223, pp. 79-93.
(11) Gupta, P. K., 1979, “Dynamics of Rolling Element Bearings – Part IV: Ball
Bearing Results,” J. Lubrication Technology, 101, pp. 319-326.
(12) Gupta, P. K., 1979, “Dynamics of Rolling Element Bearings – Part I:
Cylindrical roller bearing analysis,” J. Lubrication Technology, 101, pp. 293-
304.
(13) Gupta, P. K., 1979, “Dynamics of Rolling Element Bearings – Part III: Ball
Bearing Analysis,” J. Lubrication Technology, 101, pp. 312-318.
(14) Meyer L., Ahlgren F., and Weichbrodt B., 1980, “An analytic model for ball
bearing vibrations to predict vibration response to distributed defects,” J.
Mechanical Design, 102, pp. 205-210.
(15) Saheta, V.,2001, “Dynamics of Rolling Element Bearings using Discrete
Element Method”, M.S. thesis, Purdue University, West Lafayette, IN, USA.
(16) Ghaisas, N., Wassgren, C., Sadeghi, F., 2004, “Cage Instabilities in Cylindrical
Roller Bearings”, J. Tribology, 126(4) pp. 681-689.
(17) Sopanen J., and Mikkola A., 2003, “Dynamic model of a deep-groove ball
bearing including localized and distributed defects. Part 1: theory,” Proc. I.
Mech. E., Part K: J. Multi-body Dynamics, 217(3), pp. 201-211.
(18) Sopanen J., and Mikkola A., 2003, “Dynamic model of a deep-groove ball
bearing including localized and distributed defects. Part 2: implementation
and results,” Proc. I. Mech. E., Part K: J. Multi-body Dynamics, 217(3), pp.
213-223.

372
(19) Ashtekar A., Sadeghi F., and Stacke L., 2010, “Surface defects effects on
bearing dynamics,” Proc. I. Mech. E., Part J: J. Engineering Tribology, 224(1),
pp. 25-35.
(20) Ashtekar A., Sadeghi F., and Stacke L., 2008, “A New Approach to Modeling
Surface Defects in Bearing Dynamics Simulations,” J. Tribology, 130(4), pp.
041103.
(21) Lim T., and Singh R., 1990, “Vibration transmission through rolling element
bearings, part I: Bearing stiffness formulation,” J. Sound and Vibration,
139(2), pp. 179–199.
(22) Hendrikx R., Van Nijen G., and Dietl P., 1999, “Vibrations in household
appliances with rolling element bearings,” Proc. International Seminar on
Modal Analysis, Katholieke Universiteit Leuven, 3, pp. 1537–1544.
(23) Tiwari M., 2000, “Effect of Radial Internal Clearance of a Ball Bearing on the
Dynamics of a Balanced Horizontal Rotor,” J. Sound and Vibration, 238(5),
pp. 723-756.
(24) Tiwari M., 2000, “Dynamic Response of an Unbalanced Rotor Supported on
Ball Bearings,” J. Sound and Vibration, 238(5), pp. 757-779.
(25) Prenger N., 2003, “Modeling the Dynamics of Rolling Element Bearings Using
ADAMS,” M.S. thesis, Purdue University, West Lafayette, IN, USA.
(26) Stacke L., Fritzson D., and Nordling P., 1999, “BEAST—a rolling bearing
simulation tool,” Proc. I. Mech. E., Part K: J. Multi-body Dynamics, 213(2), p.
63–71.
(27) Ting J., 1992, “A robust algorithm for ellipse-based discrete element
modelling of granular materials,” Computers and Geotechnics, 13(3), pp.
175-186.
(28) Ting J., Khwaja M., and Meachum L., 1993, “An ellipse-based discrete element
model for granular materials,” J. Numerical and Analytical methods in
Geomechanics, 17(9), pp. 603-623.
(29) Matuttis H. G., Luding S., and Herrmann H. J., 2000, “Discrete element
simulations of dense packings and heaps made of spherical and non-spherical
particles,” Powder Technology, 109(1-3), pp. 278-292.
(30) Gupta P. K., 1975, “Generalized dynamic simulation of skid in ball bearings,”
J. of Aircraft, 12, pp. 260–265.
(31) Hamrock B., Schmid S., and Jacobson B., 2004, Fundamentals of fluid film
lubrication, Marcel Dekker, Ohio.
(32) Kragelskii I.V., 1965, Friction and wear, Butterworth, London.
(33) Gupta P. K., 1984, Advanced Dynamics of Rolling Elements, Springer-Verlag,
New York.
(34) Gupta P. K., 1974, “Transient ball motion and skid in ball bearings,” J.
Lubrication technology, 97, pp. 261-269.
(35) Hagiu G., and Gafitanu M., 1994, “Preload-service life correlation for ball
bearings on machine tool main spindles,” Wear, 172(1), pp. 79-83.

373
Predictions for run-up procedures of
automotive turbochargers with full-
floating ring bearings including thermal
effects and different bearing setups
D Vetter, T Hagemann, H Schwarze
Institute of Tribology and Energy Conversion Machinery,
Technical University of Clausthal, Germany

ABSTRACT

Automotive turbochargers feature a variety of boundary conditions influencing the


operating behaviour during run-up procedures. As a novel aspect the influence of
the thrust bearing and load on lateral vibration phenomena is investigated in detail
for this application. The run-up predictions show that the thrust bearing and load
have significant influence on the frequencies as well as on the magnitudes of the
vibrations. In particular, the amplitudes of conical vibrational modes are damped.
Further, the consideration of the hot gas boundary conditions is identified as
decisive in order to predict the bearing parameters and consequently the vibrations
precisely.

NOTATION

A system matrix q state vector


B input matrix T temperature
B axial bearing length t time
C damping matrix U hydrodynamic circumferential
D bearing diameter speed
d fluid film damping coefficient u shaft translation degrees of
F, f force (vector) freedom
x, y, z cartesian coordinates
F0 , F1, F2 viscosity factors
 angular direction of fluid film
h film thickness
force
K stiffness matrix  lubricant dynamic viscosity
K x ,Kz turbulence factors  density
M inertia matrix Φ matrix of eigenvectors
p pressure  shaft rotational angular velocity

1. INTRODUCTION

Due to light rotor weights and high rotational speeds automotive turbochargers
commonly feature full- or semi-floating ring bearings (FRB/SFRB). The high rotating
frequency effects high unbalance forces while static forces are very low.
Consequently, the rotor operates above its linear stability limit. The design with two
separated oil films involves the advantage of additional external damping provided
by the second film if subsynchronous self-exited vibrations occur and one of the

_______________________________________
© The author(s) and/or their employer(s), 2014
375
two lubricant films becomes unstable. Instead of a total instability leading to a
destruction of the rotor, stable limit cycles are reached which enable continuous
operation. While synchronous vibrations can be analysed by linear theory [1] the
prediction of subsynchronous vibrations requires a nonlinear bearing force model.
The vibration phenomena reported for turbochargers with FRBs are
comprehensively discussed in [2].

The influence of the effective bearing geometry on the performance of high-speed


turbochargers featuring FRBs is investigated by several researchers [3]-[8].
Concordantly a significant decrease of the ratio between ring and rotor speed is
reported if the rotor speed increases. Trippett and Li [4] developed a thermo-elastic
model for the prediction of ring-speed, effective clearance, power loss and local
temperature of FRBs and validated it against measurement data. This model is
further applied and verified by measurements of Meyer [5]. An alternative model is
presented by San Andres and Kerth [6] and validated with test data. However, for
hot gas applications all parameters strongly depend on the heat disposed from a
hot shaft [7]. Porzig et al. [8] show that for hot gas applications the ring-speed and
the clearances of the inner and outer film are significantly influenced by the heat
flow from the shaft into the inner oil film.

Turbochargers commonly feature hydrodynamic thrust bearings (TB) to balance the


axial load. Aside from this function they can strongly influence lateral vibration
behaviour of high speed rotating machinery. Mittwollen et al. [9] investigated the
influence of axial clearance and load on the gyrating movement of the thrust collar
and its effect on bending vibrations. The authors demonstrate that an increase of
the axial load or a reduction of the clearance can cause a shift of critical speeds and
threshold speed of instability to higher rotating frequencies. Berger et al. [10]
studied the lateral vibration of a shaft optionally supported by a thrust bearing at
one end. Their results reveal that the thrust bearing provides additional stiffness
and damping on the flexible shaft. Further, their nonlinear computational model
predicts an excitation of a critical speed by a defect of the thrust bearing. In
agreement with these authors other researchers identified significant influences on
lateral vibrations effected by thrust bearings, e. g. [11]-[13]. However, thrust
bearings are neglected in most rotordynamic analyses and no investigations of
small turbochargers including thrust bearings can be found.

This paper presents the influence of the predicted effective bearing parameters and
the thrust bearing analysis on the theoretical vibration behaviour of a small
automotive turbocharger with FRBs. It further includes a validation of the predicted
results against measurement data from a hot gas test rig.

2. THEORY

Bearing model
The prediction of the journal bearing forces generally requires the solution of
extended and generalized Reynolds equation

  F2 p    F2 p     F1   
   U
x  K x x  z  K z z 
  h 
x  
 
F0    t
   h . (1)

Equation (1) considers the 3D viscosity distribution due to variable temperature in


all three space directions of the film by the following factors [14]:

h h h
dy y y F1 
F0  
0

, F1  
0
dy , F2     y  F
0
 dy .
0 
(2)

376
Additionally, the effects of local turbulent flow or Taylor vortices are taken into
account by the coefficients Kx and Kz according to the model presented by
Mittwollen [15]. The direct numerical solution of (1) satisfies the highest
requirements on generality and accuracy. On the other hand, it is time consuming
and provides the disadvantage of numerical instability during the iterative solution
procedure as the source term on the right hand side implicitly depends on the
solution. Thus, an approximation procedure for the nonlinear fluid film force
proposed by Glienicke et al. in [16] is applied. As rotational speeds are very high it
is assumed that the cavitation zone of the film is equal to the one determined for a
certain rotor position and exclusive rotation neglecting the time depended source
term. Consequently, (1) becomes linear and can be separated into

  F2 pr    F2 pr     F1  
   U   h    and (3)
x K
 x  x   z K
 z z  x 
  F0 

  F2 pt    F2 pt  
   
x  K x x  z  K z z   t
   h . (4)

The forces resulting from both pressure distributions can be superposed in order to
predict the fluid film force. Further, the transient forces from (4) can be
approximated using four damping coefficients dij derived from two perturbed
equations. Finally, the predicted fluid film force is determined by

 Fx   sin   r    d11 d12   x 


 F   Fr        . (5)
 y  cos   r    d21 d22   y 

Herein and are the lateral velocities between shaft and ring or ring and housing
respectively. Fr is the absolute value and γr is the direction of the bearing force
resulting from rotation, i. e. (3). Assuming validity of similitude theory Fr, γr, and dij
can be listed in non-dimensional form in a look-up table as a function of relative
eccentricity ε and attitude angle β. If periodicity is fulfilled only one pad has to be
analysed and all other information can be generated by rotational transformation.
For every angular position the eccentricity is varied by 25 steps. In order to
approximate the nonlinear force characteristic of the film the distance between two
eccentricities is reduced with increasing shaft displacement. The attitude angle is
shifted between 3° and 6° from one to another radial axis depending on the pad
angle. At every position of this grid the non-dimensional form of equation (3) is
solved based on Elrod’s algorithm [17] and the non-dimensional force due to shaft
rotation F r and its direction γr are determined. In a second step the two first order
perturbed equations are solved. Neglecting compressibility of the fluid these to
equations are derived assuming the following linearization due to small shaft
motion

H
( , z )   X  sin   Y  cos  ,

(6)
     
( , z )   r ( , z )    X    Y   Y .
 X  stat  stat

Herein Π is the film pressure, H is the film thickness and X′ and Y′ are the shaft
velocities each in non-dimensional form. The four non-dimensional damping
coefficients d of the fluid film are calculated integrating the solution variables
according to

377
 2B / D
1 1
  
 (i  1) cos   (i  2) sin   dz d
4 B / D 0 
d ik   . (7)
0  qk  2 

Figure 1 illustrates the procedure to predict the bearing force according to (5)
considering the speed dependent fluid film viscosities and bearing clearances.

Non-dimensional look-up table


Extended Reynolds Equation:
Pressure distribution due to shaft rotation

F r , γr = f(ε,β)

First order perturbed equations:


Pressure variation due to shaft movement

dik = f(ε,β)

Prediction of dimensional parameters Fr


Rotor position and
and dik considering effective radial
lateral velocities
clearance and fluid viscosities

Predicted bearing
force according to Effective radial clearance and fluid
equation (5) viscosities predicted by thermal analysis

Figure 1: Bearing force analyses

The numerical procedure used to determine the content of the look-up tables is
more comprehensively described in [18] and [19]. Optionally the friction forces can
be added to this table and used to predict the ring speed in the rotordynamic
analysis. The effective fluid viscosity parameters according to (2) as well as the
relative clearances are taken from an a priori thermal analysis according to [4] or
[8]. For both predictions the ring speed is listed as a function of rotor speed as well.
However, the ring speed is predicted by the time depended friction forces during
the run-up procedure in these investigations.

The linear stiffness and damping coefficients for the collar gyration induced by
bending vibration are derived from a perturbation of the thermo-hydrodynamic
analysis presented in [20].

Rotor model
The rotor model consists of a flexible shaft which is modelled by Finite Elements
using the Timoshenko beam theory. Compressor and turbine wheel as well as the
floating rings are assumed as rigid bodies. The mathematical formulation leads to a
system of ordinary differential equations:

378
Mu
  Cu ) .
  Ku  f(t , , u, u (8)

The vector of forces f contains the unbalance-, the gravity-, the gyroscopic- and the
nonlinear bearing forces. The matrices on the left hand side of equation (8) are
constant and time-invariant. The second order system is transformed in a first
order state space formulation using a new state vector = ( , ) and the input
vector = ( , ) .

  Aq  Bf(t , , q) .
q (9)

An eigenvalue analysis is conducted using the homogeneous system of equations.


The eigenvectors of the free system are arranged in a modal matrix Φ which is the
basic approach for a modal reduction for example described in [21]. The relation
between the physical coordinates and the new modal coordinates ∗ is given by

q*  Φ1q . (10)

The transformed and reduced system can be integrated by a Numerical Differential


Formula (NDF).

3. VERIFICATION OF THE NONLINEAR ROTORDYNAMIC MODEL

The predictions of the nonlinear


rotordynamic model are verified
with measurements presented for
an overhung rotor in [16]. The
rotor is supported by four-lobe
bearings and exhibits a distinct
nonlinear behaviour. The
discretized model of the overhung
rotor is depicted in Figure 2. The
four-lobe bearings are placed at
node two and four. The added Figure 2: Overhung rotor model
mass as well as the unbalance
are located on node six.

Figure 3 shows the predicted amplitudes as a function of the rotor speed for
different unbalance values on the left side. For low unbalances typical characteristic
of a rotor in journal bearings can be observed. There are the first resonance at
rotor speeds of about n=4000 rpm and the threshold speed of instability at
n=11000 rpm. If the unbalance increases a second resonance arises caused by an
oil-whirl which excites the first natural frequency of the rotor-bearing-system. A
comparison between measured and predicted amplitudes is depicted on the right
side in Figure 3. The amplitudes of the mean resonance AR show very good
agreement. In the experiment the subsynchronous resonance Asub occurs if an
unbalance mass of about u=130 mmg is reached. In case of the prediction this
value is shifted to u=200 mmg. Consistently, the threshold speed of instability
slightly decreases if the unbalance increases. Though, some experimental
parameters such as oil supply temperature and the mass moments of inertia of the
disk are not defined exactly in [16]. In the prediction the support structure is
assumed as ideally rigid and an interaction with the rotor-bearing system is
neglected. The influence of the drive is also not taken into account. Consequently,
an exact agreement between measurement and prediction cannot be expected.
Therefore, the good correspondence between the predicted and measured
phenomena can be described as very satisfactorily.

379
200 180
u = 100 mmg 200 300 400 Measurement A R [16]
180 160 Prediction A R
Measurement A sub [16]
160 140 Prediction A sub

140
120
Predicted amplitude [µm]

120

Amplitude [µm]
100
100
80
80
60
60

40
40

20 20

0 0
1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 0 50 100 150 200 250 300 350 400
Rotor Speed [rpm] Unbalance [mmg]

Figure 3: Measured and predicted amplitudes of the overhung rotor

4. RESULTS

Figure 4 shows a typical design of a bearing setup for small automotive


turbochargers including two FRB and a double acting thrust bearing with two
collars. It is similar to the one investigated in this paper. Further, Figure 4 includes
the discretized rotor model. The rotor consists of a flexible shaft with added
masses. Whereas the floating
ring bearings are positioned at Table 1: Characteristic parameters
node seven and eight the
thrust bearing is added on node Pa ram e te r Value
five. Compressor and turbine Total rotor length lR in mm ≈ 100
wheel are arranged on nodes Total rotor mass m R in g ≈ 100
four and eleven. It is assumed Lubricant 5W40
that unbalances exist on the Oil supply temperature T sup in °C 90
turbine and compressor wheel
FRB length B i / B a in mm 3.6 / 6.2
with a phase shift of 180°. The
FRB diameter D i / D a in mm 6.0 / 9.6
characteristic parameters of the
rotor and the bearings are TB diameter D i / D a in mm 8.5 / 11.8
listed in Table 1. Unbalance u in mmg 0.05

full-floating
thrust ring
collar S1

thrust
collar S2
outer oil
film

turbine
wheel

thrust bearing inner oil


oil film 1 film

compressor thrust bearing


wheel oil film 2

Figure 4: Schematic design and rotor model of the turbocharger

380
The measured frequency spectrum of the compressor wheel is depicted in Figure 5.
It shows a distinct low frequency vibration at high rotor speeds that can be related
to a self-excited vibration in the outer lubricant gap (Sub 3) as it is lower than the
measured ring rotating frequency. Beside Sub 3 and the unbalance synchronous
vibration (Sync) further frequencies can be observed. As these frequencies are
higher than the ring rotating frequency they are related to whirl phenomena in the
inner lubricant film. In the speed range from 500 rps up to 1000 rps the unbalance
synchronous vibrations and the self-excited vibration Sub 1 exist. Then Sub 1
bifurcates into Sub 2 and Sub 3. This effect can be related to a change of the
excited rotor mode by the whirl in the inner oil film. In the range from 1000 rps and
1500 rps two resonances appear that can be related to Sub 2. Sub 3 is interrupted
at a rotor speed range from about n=1500 rps to n=1800 rps.

3500
Sub 1
Sub 2
3000
nFR,T
Sub 3
Sync
2500

Rotor Speed [rps]


nFR,C
2000

1500

1000

500
0 200 400 600 800 1000
Frequency [Hz]

Figure 5: Measured frequency spectrum at the compressor nose of the


turbocharger at run-up according to [8]

In order to study the influence of the predicted effective bearing parameters by the
thermal a priori analysis and the thrust bearing gyrating stiffness and damping run-
up simulations are performed for three different cases. In case (A) no thrust
bearing is arranged. In case (B) and (C) a double acting thrust bearing is applied
and loads of Fth = 22 N (case B) and 87 N (case C) are investigated. The results for
both thermal analyses are presented consecutively. The different speed dependent
characteristics of the predicted effective film parameters are presented in Figure 6.

1 1.6 Porzig et al. OF CS


eff 0

Porzig et al. OF TS
eff 0

Relative effective clearance  /


Relative effective viscosity  /

0.9 Porzig et al. IF CS


1.4 Porzig et al. IF TS
0.8 Trippett and Li OF
Trippett and Li IF
0.7 1.2
Porzig et al. OF CS
0.6 Porzig et al. OF TS
Porzig et al. IF CS 1
Porzig et al. IF TS
0.5 Trippett and Li OF
Trippett and Li IF 0.8
0.4

500 1000 1500 2000 2500 3000 3500 500 1000 1500 2000 2500 3000 3500
Rotor Speed [rps] Rotor Speed [rps]

Figure 6: Predicted viscosities and clearances for the inner film (IF) and
outer film (OF) on compressor side (CS) and turbine side (TS)

For the modal reduction the eigenvectors up to the fourth bending mode of the free
system are used.

381
Results using a priori thermal analysis according to Trippett and Li [4]
The turbocharger run-up starts at a rotor speed of n=20 rps and is increased
linearly up to 3500 rps in 6 s. Beside the unbalance synchronous vibration the
predicted frequency spectrum in Figure 7 shows two subsynchronous vibrations that
can be related to the inner oil film and are called Sub 1 and Sub 2. Sub 1 starts at
rotor speeds of about n=100 rps. Its frequency rises from f=60 Hz at the beginning
to f=220 Hz at a rotor speed of n=650 rps. At this rotor speed Sub 1 bifurcates into
Sub 2 which starts with a vibration frequency of 320 Hz. With further rising rotor
speeds the frequency of Sub 2 is not changing anymore but rather stays constant.
This effect is well-known as locking effect if the oil-whirl turns into the oil-whip at a
natural frequency of the rotor-bearing system. Sub 2 reaches a frequency of
f=690 Hz at the end of the run-up procedure. The subharmonic vibrations caused
by a whirl in the outer lubricant gap will subsequently be called Sub 3. Here, Sub 3
is only slightly present if no thrust bearing is considered (case A). If thrust force
increases the following effects on Sub 1 can be detected:

 the beginning and ending are shifted to higher rotor speeds


 the amplitudes decrease

The influence of the thrust bearing on Sub 2 and the unbalance synchronous
vibrations are insignificant.

Sub 3
case (A) Sub 2 case (B)

Sync

3500 3500

1 3000 1 3000
Amplitude
Amplitude

2500 2500
0.5 0.5
2000 2000

1500 1500
0 0
0 1000 0 1000
500 500 500 500
Rotor Speed [rps] Rotor Speed [rps]
1000 1000
Sub 1 1500
0
1500
0
Frequency [Hz] Frequency [Hz]

case (C)
0.5
0
(A)
-0.5
-1
0 500 1000 1500 2000 2500 3000 3500

0.5
Amplitude

(B)
0
3500
-0.5
1 3000
-1
Amplitude

2500 0 500 1000 1500 2000 2500 3000 3500


0.5
2000
0.5
(C)
1500
0 0
0 1000
-0.5
500 500 Rotor Speed [rps]
1000 -1
0 0 500 1000 1500 2000 2500 3000 3500
1500
Frequency [Hz] Rotor Speed [rps]

Figure 7: Predicted frequency cascades and time signals at the compressor


nose using the thermal analysis according to Trippett and Li [4]

382
Results using a priori thermal analysis according to Porzig et al. [8]
The same run-up simulations are conducted using the effective bearing parameters
determined by the theoretical results of Porzig et al. [8]. Compared to the above
presented predictions Sub 3 exists over a wide range of rotor speeds and
dominates the vibrational amplitudes as depicted in Figure 8. In case (A) Sub 3
starts at a rotor speed of n=650 rps with a vibration frequency of 80 Hz and
reaches a frequency of 310 Hz at the end of the run-up simulation. At light thrust
loads of case (B) the beginning of Sub 3 is delayed to higher rotor speeds of about
n=1500 rps and the amplitudes are reduced. For maximum thrust load case (C)
the beginning of Sub 3 shifts back to lower rotor speeds of n=1000 rps.
Additionally, the following statements can be derived if a thrust bearing is arranged
and thrust load increases:

 the amplitudes of Sub 3 decrease, particularly at lower rotor speed


 the start frequency of Sub 1 remains nearly constant
 the end of Sub 1 is shifted to lower rotor speeds for light thrust loads and
back to higher rotor speeds for high thrust loads
 the frequency at the end of Sub 1 first increases and then decreases
 the maximum amplitude of Sub 1 slightly decreases

Again the influence of the thrust bearing and load on Sub 2 and the unbalance
synchronous amplitudes are insignificant.

case (A) case (B)


Sub 3

Sub 2
Sync

3500 3500

1 3000 1 3000
Amplitude

Amplitude

2500 2500
0.5 0.5
2000 2000

1500 1500
0 0
0 1000 0 1000
500 500 500 500
Rotor Speed [rps] Rotor Speed [rps]
1000 1000
Sub 1 1500
0
1500
0
Frequency [Hz] Frequency [Hz]

case (C) 0.5


0 (A)
-0.5
-1
0 500 1000 1500 2000 2500 3000 3500

0.5
Amplitude

3500
0 (B)
-0.5
1 3000
-1
Amplitude

2500 0 500 1000 1500 2000 2500 3000 3500


0.5
2000
0.5
0
1500
0 (C)
0 1000
-0.5
500 500 Rotor Speed [rps]
1000 -1
0 0 500 1000 1500 2000 2500 3000 3500
1500
Frequency [Hz] Rotor Speed [rps]

Figure 8: Predicted frequency cascades and time signals at the compressor


nose using the thermal analysis according to Porzig et al. [8]

383
Ring speed
The measured and predicted ring speeds are presented in Figure 9. Despite rising
rotor speeds a decrease of the compressor-sided floating ring speed is observed in
the measurement. This effect can be related to the very large amplitudes of the
subharmonic vibration at rotor speeds higher than n=2000 rps. The increased
eccentricities in the outer lubrication gap induce high frictional forces which
decelerate the floating ring.

The results of the prediction in case of the thermal model according to Trippett and
Li [4] are depicted in Figure 9 (left). The ring speed is underpredicted compared to
the measurement data. The ring speed predicted using the thermal model
according to Porzig et al. [8] and a dynamic torque balance is depicted in Figure 9
(right). The correspondence between measurement and prediction is significantly
improved. As the heat flow from the shaft into the inner oil film is considered the
simulated temperature distribution differs from the one predicted by the first
analysis. Consequently, the relation between the clearances of the inner and outer
lubrication gap is changed and the ring speed is modified.

700 700

600 600
Floating Ring Speed [rps]

Floating Ring Speed [rps]

500 500

400 400

300 300

200 compressor side prediction 200 compressor side prediction


turbine side prediction turbine side prediction
compressor side measurement [8] compressor side measurement [8]
turbine side measurement [8] turbine side measurement [8]
100 100
500 1000 1500 2000 2500 3000 3500 500 1000 1500 2000 2500 3000 3500
Rotor Speed [rps] Rotor Speed [rps]

Figure 9: Measured and predicted ring speeds by comparison, thermal


model according to Trippett and Li [4] (left) and Porzig et al. [8] (right)

5. CONCLUSIONS

The presented results show that the effective parameters of the lubricant gap and
the arrangement of a thrust bearing have significant influence on the predicted
results of run-up simulation. The correspondence between prediction and
measurement data was remarkably improved using the a priori thermal analysis
which considers the hot gas boundary conditions. In particular, the dominant
amplitudes of the self-excited vibration induced by the outer film are only predicted
based on this data and are experimentally identified in [8]. The predictions were
performed for an estimated unbalance configuration. Nevertheless, in additional
investigations similar tendencies were observed for phase shifts between the two
unbalances. The consideration of thrust bearings especially influences the prediction
for self-excited vibrations that are typically related to conical modes. The results
indicate that a precise modelling of the speed depended thrust load and its
interaction with the thrust bearing can be a key feature improving run-up
simulations for this application.

384
In general good agreement between prediction and the validation data was
reached. An advantage of the presented analysis is that it is highly time-effective
and run-up simulations can be performed in less than two hours on a common
desktop pc. However, there are some assumptions that have to be discussed
critically. As subsynchronous vibrations are apparent and thrust bearings for this
application often do not feature periodically arranged sliding surfaces the linear
thrust bearing model is insufficient. Further, the fluid force model is approximate,
the hydraulic interaction between the two films is neglected and the transient
pressure built-up is not taken into account for the prediction of the floating ring
speed. Consequently, the dynamic characteristic of the ring speed which is in this
case measured at compressor side cannot be simulated precisely. The effect of
misalignment between shaft and bearing due to vibrational behaviour is also
neglected.

ACKNOWLEDGEMENT

The Authors would like to thank Mr H. Raetz, Prof. J. R. Seume and Mr D. Porzig for
providing valuable and detailed information concerning their measurement and
analysis data.

REFERENCE LIST

[1] San Andrés, L., Rivadeneira, J. C., Gjika, K., Chinta, M., LaRue, G., (2007), “A
Virtual Tool for Prediction of Turbocharger Nonlinear Dynamic Response:
Validation Against Test Data”, J. of Engineering for Gas Turbines and Power,
129(4), pp. 1035-1046
[2] Schweizer, B., (2009), “Total instability of turbocharger rotors – Physical
explanation of the dynamic failure of rotors with full-floating ring bearings”, J.
of Sound and Vibration, 328(1-2), pp. 156-190
[3] Born, H., (1987), “Analytical and Experimental Investigation of the Stability of
the Rotor-Bearing System of a New Small Turbocharger,” ASME Paper No. 87-
GT-110.
[4] Trippett, R., Li, D., (1983), “High-Speed Floating-Ring Bearing Test and
Analysis”, ASLE Trans. 27(1), pp. 73–81.
[5] Meyer, A., (1987), “Äußere Lagerdämpfung für sehr hochtourige,
gleitgelagerte Rotoren”, PHD Thesis, TH Karlsruhe, Germany
[6] San Andrés, L., Kerth, J., 2004, “Thermal Effects on the Performance of
Floating Ring Bearings for Turbochargers,” Proc. Inst. Mech. Eng., Part J: J.
Eng. Tribol. 218, pp. 1–14.
[7] San Andrés, L., Barbarie, V. C., Bhattacharya, A., Gjika, K., (2012), “On the
Effect of Thermal Energy Transport to the Performance of (Semi) Floating
Ring Bearing Systems for Automotive Turbochargers”, J. of Engineering for
Gas Turbines and Power, 134(10), 102507
[8] Porzig, D., Raetz, H., Schwarze, H., Seume, J. R., (2014), “Thermal Analysis
of small high-speed floating-ring journal bearings”, Proc. of the 11th Int. Conf.
on Turbochargers and Turbocharging, IMechE, London, GB (to be published)
[9] Mittwollen, N., Hegel, T., Glienicke, J., (1991), “Effect of Hydrodynamic Thrust
Bearings on Lateral Shaft Vibrations”, Journal of Tribology, 113, pp. 811-818.
[10] Berger, S., Bonneau, O., Frene, J., (2000), “Influence of Axial Thrust Bearing
Defects on the Dynamic Behavior of an Elastic Shaft”, Tribology International,
33, pp. 153-160.
[11] Lie, Y., Bhat, R. B. (1995), “Coupled Dynamics of a Rotor-Journal Bearing
System Equipped with Thrust Bearings”, Shock and Vibration, 2(1), pp. 1-14.

385
[12] Lie, Y., Bhat, R. B. (1996), “Prediction of Hydrodynamic Thrust Bearing
Performance Using the Modified Trefftz Method”, Tribology Transactions,
39(1), pp. 112-120.
[13] Chang-Jian, C.-W., (2010), “Nonlinear Simulation of Rotor Dynamics Coupled
with Journal and Thrust Bearing Dynamics under Nonlinear Suspension”,
Tribology Transactions, 53(6), pp. 897-908
[14] Dowson, D., (1962), “A Generalized Reynolds Equation for Fluid Film
Lubrication”, Int. J. Mechanical Sciences, 4, pp. 159-170.
[15] Mittwollen, N., (1990), “Betriebsverhalten von Radialgleitlagern bei hohen
Umfangsgeschwindigkeiten und hohen thermischen Belastungen –
Theoretische Untersuchungen”, VDI series 1 no. 187, VDI-Verlag, Dusseldorf,
Germany.
[16] Glienicke, J., Eilers, M., Gerdes, R., (1993), “Nichtlineare Rotordynamik”,
FVV-Report, Frankfurt am Main, Germany
[17] Elrod, H. G., 1981, “A Cavitation Algorithm”; ASME J. of Lubrication
Technology, 103(3), pp. 350-354.
[18] Hagemann, T., Kukla, S., Schwarze, H., (2013), “Measurement and prediction
of the static operating conditions of a large turbine tilting-pad bearing under
high circumferential speeds and heavy loads”, Proceedings of the ASME Turbo
Expo, ASME Paper No. GT2013-95004, San Antonio, TX., USA
[19] Kukla, S., Hagemann, T., Schwarze, H., (2013), “Measurement and prediction
of the dynamic operating conditions of a large turbine tilting-pad bearing
under high circumferential speeds and heavy loads”, Proceedings of the ASME
Turbo Expo, ASME Paper No. GT2013-95074, San Antonio, TX., USA
[20] Lüneburg, B., Kraft, C., Dettmar, D., Mermertas U., Schwarze H., (2011),
“Enhanced Hydrodynamic Thrust Bearing Analyses of Turbo Sets for Power
Generation”, 10th EDF/Pprime Workshop, Poitiers, France
[21] Gasch, R., Knothe, K., Liebich, R., (2012), “Strukturdynamik: Diskrete
Systeme und Kontinua”, Springer Berlin Heidelberg, Germany

386
A study on unsteady aerodynamic
excitation forces on radial turbine blade
due to rotor-stator interaction
W Sato, A Yamagata, H Hattori
IHI Corporation, Japan

ABSTRACT

It is important to predict the resonant vibration level for the design of a variable nozzle
turbine. In this study, unsteady CFD analyses are conducted to investigate flow
mechanism that causes aerodynamic excitation forces. As a result, it is found that
excitation forces are affected by shock wave and clearance flow and can be modeled
with a linear function of impeller inlet dynamic pressure and Mach number. Then
resonant vibration level is calculated by the CFD and FEM, and compared with
experimental data. It is validated that CFD results can be used for qualitative
evaluation of excitation forces.

NOMENCLATURE

VGS :Variable Geometry System


VPF :Vane Passing Frequency
CFD :Computational Fluid Dynamics
S/S :Suction Surface
P/S :Pressure Surface
L/E :Leading Edge
T/E :Trailing Edge

1 INTRODUCTION

In recent years, due to the demand of CO2 reduction, further improvement of fuel
consumption is needed in automotive engines. In order to reduce the fuel
consumption, turbo-charging is widely applied with the concept of downsizing which is
to reduce engine capacity while preserving performance. To satisfy the requirement of
such engine performance for the whole operating range, a variable flow capacity
turbine which has variable nozzle vanes is mainly used in automotive diesel engines
currently. On the other hand, by applying the variable geometry system, the unsteady
interaction between the nozzle vanes and the impeller blades generates a periodical
unsteady loading which excites the impeller blade vibration. Therefore, it is needed to
predict the aerodynamic excitation forces at the design.

Several researches for unsteady rotor-stator interaction within radial turbines have
been published and these evaluated the vibratory response quantitatively by
estimating excitation force using the unsteady CFD analysis. Kawakubo (1) conducted
unsteady CFD analysis of the variable nozzle turbine where the pressure ratio is
changed with fixed nozzle opening to evaluate the flow mechanism of the impeller

_______________________________________
© The author(s) and/or their employer(s), 2014
389
excitation force. He showed the excitation force at the impeller leading edge became
higher than other region due to the leakage flow at the nozzle side clearance and the
shock wave generated near the nozzle trailing edge. Schwitzke et al. (7) also
conducted unsteady CFD analysis and showed three flow mechanism which is
responsible for the blade excitation could be identified. The first is rotor-stator
potential interaction, the second is wakes of vanes, and the third is a separation
vortex on S/S near impeller L/E. However, there are not so many papers that
systematically describe the effects of nozzle opening and turbine pressure ratio on the
turbine impeller excitation force.

Therefore, the authors conducted unsteady CFD analyses for the variable nozzle
turbine to investigate the effects of nozzle opening and pressure ratio on excitation
force, and to clarify each flow mechanism. After that, principal aerodynamic
parameters for the excitation force were identified and the excitation force was
modeled with a simple linear function. Finally, resonant vibration level was calculated
by the CFD and FEM, and compared with experimental data. It was validated that the
CFD could capture excitation forces qualitatively.

2 METHODOLOGY

2.1 Turbine Configuration


Figure 1(a) shows a cut-way view of the RHF5V turbocharger used in this study, which
was developed for the passenger vehicle's diesel engine in IHI. The turbine is
composed of stationary parts, which are inlet scroll and nozzle, and a rotating part,
impeller. Figure 1(b) shows an axial view of the nozzle vanes and impeller blades. The
number of nozzle vanes and impeller blades are 14 and 9 respectively. The impeller
diameter is about 50mm. In order to ensure the smooth operation of variable nozzle
vanes, the actual turbine has clearance gaps for both at the hub and the shroud side.
In this analysis, these gaps for shroud and hub sides are set equally and its amount is
set to 1.5% of the nozzle passage height. In this study, CFD analyses and vibratory
measurement were conducted for three nozzle openings, which were nozzle closed,
medium and fully-opened conditions and also conducted at different pressure ratio
conditions for each nozzle opening.

2.2 Numerical Procedure


In this study, unsteady 3D Reynolds-averaged compressible Navier-Stokes equations
were solved using the in-house CFD code developed in IHI Corporation. For the
convective fluxes, the MUSCL-ROE TVD scheme was used. High order accuracy can be
achieved by the MUSCL scheme and Shock wave can be captured by the TVD scheme.
For turbulence closure one-equation Spalart-Allmaras model was used. Figure 1(c)
shows meridional view and axial view of the computational mesh. An H-type structural
grid is used for each blade passage. Computational domain includes only a nozzle and
an impeller passage and an inlet scroll was not included. This is because the flow
distortion due to the inlet scroll's asymmetry is rectified and suppressed within a
nozzle passage (3). The minimum size of wall adjacent grid is set as y+ value to be
lower than 3 in order to resolve a viscous sub-layer sufficiently. It is necessary to use
the full annulus model for unsteady CFD analysis because there is no cyclic symmetry.
Since the computational region could not be reduced anymore, the total number of
grids became 16 million for the nozzle and the impeller passage. In order to reduce
the computing time, parallel computing using PC cluster was applied for this unsteady
calculation, and 3 weeks computational time is needed to obtain periodically
converged solution for each case.

As inlet boundary conditions, total pressure and swirl angle at nozzle inlet were
estimated from 1D calculation of an inlet scroll. The impeller rotating speed was set to
match the resonant frequency, which is calculated from the impeller blade's eigen

390
frequency and the number of nozzle vanes. Calculation cases investigated in this
study are summarized in Table1. When the pressure ratio was changed, the inlet
pressure was changed at the constant outlet pressure.

Figure 2 shows the procedure of calculating the excitation force from the unsteady
CFD results. The impeller interacts with the flow from the nozzle as shown in Figure
2(a) and the impeller is excited by an unsteady pressure difference between P/S and
S/S as shown in Figure 2(b). An unsteady pressure difference on each computational
grid point is analysed by Fourier transform and 1st VPF component of excitation force
is extracted to use for evaluating strength of the excitation force.

2.3 Experimental Procedure


Figure 1(d) shows the procedure of non-intrusive vibration measurement system. In
this system, optical sensors direct the light to the blade tip and detect the reflected
light to register the blade pass timing. Relative motion from the un-deformed position
is calculated by the delay or the progress of blade passing time, and vibratory
response is determined.

Nozzle Vane (VGS) Nozzle Vane (VGS)


Open condition
Medium condition
Closed condition

Scroll
Impeller
Impeller
(a) Cut-way view (b) Axial view of CFD model

(c) Computational mesh

Detecting blade
Optical sensor vibratory motion
θ2

Data acquisition &


Analysis system
θ1

Vibratory response
1/rev sensor
(d) Measurement system
Fig.1 CFD model and measurement system

391
Table1 Computational cases
Nozzle opening Pressure ratio
Case1 closed low
Case2 closed medium
Case3 closed high
Case4 medium low
Case5 medium medium
Case6 medium high
Case7 opened low
Case8 opened medium

Normalized excitation
Normalized loading FFT

force
Rotation

Normalized time Excitation order


(a) Interaction (b) Unsteady loading of (c) FFT analysis
between nozzle vanes the impeller blade
and the impeller
Fig.2 Procedure to derive an unsteady excitation force on the impeller

3 RESULTS AND DISCUSSION

3.1 Turbine performance


Figure 3 shows turbine performance characteristics predicted by unsteady CFD
calculations. Pressure ratio, efficiency and mass flow rate was normalized by the
representative values, which are the medium pressure ratio, the maximum efficiency
and the maximum mass flow rate, respectively. At the open condition, the flow is
choked and the calculation under the high pressure ratio was unstable and this result
could not be achieved.

3.2 Effect of different pressure ratios


Figure 4(a), (b) shows the distributions of 1st VPF component of excitation force on
the impeller blade calculated from the unsteady CFD result. At each nozzle opening,
the excitation force becomes higher with increasing pressure ratio but the excitation
force distribution is almost unchanged. Figure 4(b) shows the variation of
area-averaged excitation force near the impeller L/E for the different nozzle opening
and the pressure ratio to evaluate the effect of different pressure ratios quantitatively.
Figure 4(b) above chart shows that at every nozzle opening the excitation force
becomes higher with increasing pressure ratio and increasing rates depend on the
nozzle opening. In this study, the impeller excitation force under the same pressure
ratio becomes the highest at the medium nozzle opening. Figure 4(b) below chart
shows the excitation force normalized by the dynamic pressure based on the impeller
inlet stagnation density and the impeller peripheral speed. In the case of the open
condition, the normalized excitation force is almost the same and not affected by the
difference of pressure ratio. From this result, we can derive that the impeller
excitation force at nozzle opened condition is proportional to the impeller inlet total
density. On the other hand, the normalized excitation forces at the closed and medium
conditions become higher with increasing pressure ratio. These effects of different
pressure ratios at different nozzle openings are discussed in the following sections.

392
0.1

Normalized efficiency
0.0

-0.1

-0.2

-0.3

-0.4
0 0.2 0.4 0.6 0.8 1 1.2
Normalized pressure ratio

1.4

1.2

1.0

0.8
Closed Medium Open
condition condition condition
0.6
0 0.2 0.4 0.6 0.8 1 1.2
Normalized mass flow rate
Fig.3 Turbine performance characteristics predicted by unsteady CFD

Open condition
(Dimensional) Medium condition
High

MAX Closed condition


30

MIN
Excitation force
(Dimensional)

20
Pressure ratio

10

0
Normalized excitation force

0.30.6 0.8 1.0 1.2 1.4


(Nondimensional)

0.2
Low

0.1

Closed Medium Open


condition condition condition
0
0.6 0.8 1.0 1.2 1.4
Normalized pressure ratio
(a) Contour of meridional view (b) Area-averaged excitation
force near impeller L/E
Fig.4 1st VPF component of excitation force

393
At first the result at the nozzle opened condition is disscussed. Figure 5(a) shows the
instantaneous Mach number distribution at mid-span. At the opened condition, the
flow is choked near the impeller T/E and therefore Mach number distributions between
nozzle vanes and impeller blades shown in Figure 5(a) are unchanged although
pressure ratio was varied. Mach number distributions are unchanged, therefore
normalized excitation force is considered to become the same.

3.2.1 Effect of shock wave


Next, the results at the nozzle closed and medium conditions are discussed. From
Figure 5(b), the region where Mach number exceeds unity exists at the downstream of
the nozzle and spreads by the increase of pressure ratio. Figure 5(b) shows the
dilatation contours to evaluate the strength of the shock wave occurred near the
nozzle exit. Definition of the dilatation is the divergence of the velocity vector, and
zero means the flow is incompressible, and positive and negative values show the
decrease and the increase of density, respectively. At the lower pressure ratio, no
clear shock waves are observed, but at the medium pressure ratio, weak shock wave
stands near the nozzle T/E, and a clearer shock wave stands at high pressure ratio.
Strong pressure disturbance due to this shock wave at the downstream of nozzle
generates the pressure fluctuation when the impeller blades are passing the nozzle
vanes and causes the strong impeller excitation force. Therefore, not only the density
(the dynamic pressure) but also the shock wave due to high Mach number has strong
effects on the excitation force at the nozzle closed and medium conditions.

1.4

0.0

Low pressure ratio Medium pressure ratio


Absolute Mach number
(a) Open condition
1.4 +
0
0.0 -
High

Shock
wave
Pressure
ratio
Low

Closed Medium Closed Medium


condition condition condition condition
Absolute Mach number Dilatation shock
(Divergence of velocity vector)
(b) Closed and medium condition
Fig.5 Instantaneous Mach number and dilatation distribution at mid-span

394
3.2.2 Effect of nozzle clearance flow
Figure 6 shows normalized excitation force near impeller L/E and effect of nozzle
clearance flow. Figure 6(b) shows velocity vectors (showing absolute velocity vectors
in the stationary frame and relative velocity vectors in the rotating frame). In a nozzle
clearance flow (the situation the impeller blade 1 in Figure 6 is experiencing), the
absolute flow direction is more radially-inward so the flow impinges on the S/S and
causes positive unsteady pressure on the S/S. Effect of the impingement is stronger
when pressure ratio and nozzle vane loading is high. When out of the clearance flows
(the circumstance the blade 2 is going through), the flow is well guided by the nozzle
vanes, so the flow does not impinge on the S/S. These two alternating flow conditions
create a highly unsteady loading at the hub and shroud section near L/E as shown in
Figure 6(a). A similar explanation may apply to the medium condition, though the
unsteadiness at the impeller L/E is relatively lower due to the relatively lower nozzle
loading.

Absolute
Normalized excitation force

0.5
High pressure ratio
High
0.4 Medium pressure ratio
Low pressure ratio

Relative
0.3

0.2
Pressure ratio

0.1

0.0
0.0 0.2 0.4 0.6 0.8 1.0
span from hub
Closed condition
Normalized excitation force

0.5
High pressure ratio
Medium pressure ratio
0.4 Flow impinges
Low pressure ratio
on impeller
0.3
Low

0.2

1 1
0.1 2 2
Closed Medium
0.0 condition condition
0.0 0.2 0.4 0.6 0.8 1.0 shock shock
span from hub Velocity vector edge line on the impeller S/S
Medium condition under low pressure ratio condition

(a) Normalized excitation force (b) Instantaneous velocity vector


distributions near impeller L/E distribution near hub endwall
Fig.6 Normalized excitation force near impeller L/E
and effect of nozzle clearance flow

3.3 Correlation between Mach number and normalized excitation force


As described in the previous section (3.1 and 3.2), the excitation force is affected by
not only the density (the dynamic pressure) but also the shock wave and nozzle
clearance flow. On the basis of this fact, this section aims at deriving a model function
to predict excitation forces using aerodynamic parameters of preliminary turbine
design, which can simplify the preliminary design process.

395
Effect of shock wave and nozzle clearance flow become stronger when nozzle vane
loading and Mach number at the nozzle exit are high. Figure 7 shows correlation
between Mach number and normalized excitation force. The correlation of the open
condition is quite different from that of the closed and medium conditions. This is
because it is considered that time-averaged impeller incidence angle of the open
condition is negative although that of closed and medium condition is positive and
similarly the flow pattern of the open condition is quite different from that of the
closed and medium condition. At the closed and medium condition, Mach number and
normalized excitation force increase with increasing pressure ratio. It is found that
Mach number as shown in Figure 7 has a stronger correlation (a stronger linearity)
with the normalized excitation force as compared to the pressure ratio as shown in
Figure 4(b). Figure 7 shows correlations at closed and medium conditions are almost
the same and could be approximated by a linear function. This is because it is
considered that time-averaged impeller incidence angles of the closed and medium
condition are close to each other and therefore the flow patterns at the closed and
medium condition are relatively similar.

It is showed that the excitation force can be modeled with a simple linear function of
one-dimensional aerodynamic parameters (dynamic pressure and Mach number) at
the closed and medium condition, which lead to higher excitation force. Therefore
aerodynamic excitation forces can be simply predicted not only with three dimensional
unsteady CFD, but also with one dimensional steady calculation at the preliminary
design of a turbine. As a future work, this methodology will be applied on other turbine
models to investigate the validity of the derived correlation.

3.4 Comparison of predicted and measured response


Finally the CFD results are validated by the comparison of predicted and measured
response. Prediction is conducted by the unsteady CFD and FEM structural analysis.
ANSYS was applied as a FEM solver. Figure 8 shows comparison of predicted and
measured response. Both the vibration responses of prediction and measurement are
normalized by their maximum values. With regard to the sensitivity of pressure ratio
on vibration response, the prediction shows good agreement with the measurement
under the closed condition, and the prediction shows about half value of the
measurement under the medium condition (Gradient of the prediction curve is 0.6 and
that of the measurement is 1.3). This is possibly because nozzle clearances in the
measurement are a little different from that in the CFD. Figure 6(a) shows that
excitation force near the impeller L/E tip of the CFD result is small under the medium
condition. Therefore modal force at the L/E tip is considered to be low although modal
displacement at the L/E tip is large, and vibration response is low. If nozzle clearance
on the shroud side in the measurement is larger than that in the CFD, effect of nozzle
clearance flow becomes stronger and could lead to high sensitivity of pressure ratio on
vibration response. On the other hand, the sensitivity (gradient of the curve) of the
prediction increases with decreasing nozzle opening similarly to the measurement.
From these results, the CFD is considered to capture the trend of the vibration
response due to nozzle opening change and pressure ratio change.

4 CONCLUSIONS

In this study, the unsteady CFD analysis of the variable nozzle turbine was conducted
to investigate the effects on the impeller excitation force of nozzle opening and
pressure ratio. And the CFD was validated by comparison of predicted and measured
response. The conclusions were derived as follows.

At the open condition, the excitation force is proportional to the impeller inlet density
with varying pressure ratio because the flow is choked near the impeller T/E and Mach
number distributions were the same.

396
At the nozzle closed and medium condition, effect of the shock wave and the clearance
flow near the nozzle T/E becomes stronger when the pressure ratio and Mach number
is high.

The excitation force could be modeled with a linear function of one-dimensional


aerodynamic parameters (dynamic pressure and Mach number). Due to this
knowledge, excitation forces can be simply predicted at the preliminary turbine
design.

The CFD was validated by the comparison of predicted and measured response. The
CFD could capture the trend of the vibration response due to nozzle opening change
and pressure ratio change.

1.2
Open condition Closed condition
Medium condition 1.0

Closed condition 0.8


0.3 0.6
Norm'd pressure ratio=1.2
0.4
1.2 Prediction
Normalized excitation force

1.0 0.2 Measurement


Normalized vibration response

0.2 0.0
1.0 1.2 0.6 0.8 1.0 1.2 1.4
0.8 Medium condition
1.0

0.8
0.8
0.1
1.0 0.6
0.8
0.4

0.2
0.0 0.0
0.4 0.6 0.8 1.0 1.2 1.2 0.6 0.8 1.0 1.2 1.4
Open condition
Absolute Mach number at 1.0
nozzle outlet
0.8
Fig.7 Correlation between
Mach number and normalized 0.6
excitation force 0.4

0.2

0.0
0.6 0.8 1.0 1.2 1.4
Normalized pressure ratio
Fig.8 Comparison of predicted and
measured response

397
REFERENCES

(1) Kawakubo, T., 2010, “Unsteady rotor-stator interaction of a radial-inflow turbine


with variable nozzle vanes”, ASME-GT2010-23677.
(2) Ota, S., Yamagata, A., Ono H. and Unno M., 2012, “A study on unsteady
aerodynamic excitation forces on radial turbine blade due to rotor-stator
interaction”, Turbomachinery, Vol.40, pp.78-84.
(3) Ota, S. and Kawakubo, T., 2007, “A numerical study on the three-dimensional
flows in the scrolls and nozzles of a radial turbine for a variable geometry
automotive turbocharger”, Proc. Int. Gas Turbine Congress 2007 Tokyo, TS-028.
(4) Tamaki, H., Shinya, G., Masaru U., 2008, “The effect of clearance flow of variable
area nozzles on radial turbine performance”, ASME-GT2008-50461.
(5) Hattori, H., Masaru, U. and Masakazu, H., 2007, “Mistuned vibration of radial
inflow turbine impeller”, AICFM9-224.
(6) Kulkarni, A. and LaRue, G., 2008, “Vibratory response characterization of a radial
turbine wheel for automotive turbocharger application”, ASME-GT2008-51355.
(7) Schwitzke, M., Schulz, A., Bauer, H.-J., 2012, “Numerical analysis of
aerodynamic excitation of blade vibrations due to nozzle guide vanes in radial
inflow turbines”, ISROMAC-14.

398
A 3D inverse design based
multidisciplinary optimization
on the radial and mixed-inflow
turbines for turbochargers
J Zhang, M Zangeneh
Department of Mechanical Engineering, University College London, UK
P Eynon
Cummins Turbo Technologies, UK

ABSTRACT

A methodology for designing radial and mixed-inflow turbines to meet multiple


aerodynamic and mechanical requirements is presented in this paper. The method
couples a 3D inverse design code and Design of Experiment Method (DoE) and the
response surface method (RSM) to design turbines which meet various design
criteria. Initially there are 17 design parameters and 15 performance parameters.
The performance parameters include aerodynamic objective at multi operating
points such as efficiency and mass flow factor as well as stress and vibration
parameters and turbine inertia. A DOE table is created consisting of 25 different
configurations. The resulting response surface is then used to investigate the
sensitivity of different design parameters on the performance parameters. By using
the sensitivity information two different impeller geometries are designed. In
design1, the design parameters are set up in such a way as to improve the
efficiency while in design2 the aim is to reduce the turbine inertia. CFD and FEA is
run on both designs and the results confirm that the design1 has higher efficiency
than the baseline impeller but at expense of higher stress and lower impeller
stiffness.

NOMENCLATURE

ANN artificial neural network a polynomial coefficient


CFD computational fluid dynamics α1 blade leading edge angle
DOE design of experiment α2 blade trailing edge angle
DRVT blade loading d(rVθ)/dm W1 inlet passage width
EA evolutionary algorithm W2 outlet passage width
FEA finite element analysis x design parameter
LE leading edge y performance parameter
TE trailing edge ^ normalised value
MOGA multi-objective genetic algorithm
GA genetic algorithm
RBF radial basis function
RSM response surface method

_______________________________________
© The author(s) and/or their employer(s), 2014
399
1 INTRODUCTION

The design of radial turbines for turbocharger applications poses difficult multi-
disciplinary challenges. The turbine has to maximize power output from a pulsating
flow from engine with variations in total pressure and temperature at inlet, have
low inertia and yet withstand high stresses at relatively high temperatures and have
high stiffness against vibration.

This creates a difficult multidisciplinary/multipoint problem that is rather


challenging to meet by using direct design approach and trial and error iteration.
One possibility is to couple automatic optimization with direct design approach so
that the design space is explored more systematically. In such an approach the
rotor geometry is parameterized by using blade shape related parameters, such as
blade angles. Typically between 24-59 design parameters are required to represent
the impeller (see [1] - [3]). Such a large number of design parameters will then
require a large population size and hence many thousands of multipoint
computations if an evolutionary optimization method is used. Even using surrogate
models such as response surface or neural networks will require a large initial
population. Furthermore, these surrogate models will not provide accurate response
surfaces, to be used for optimization, when large number of design parameters are
used. Such approaches have also been used in compressor design. For example,
Lian and Liou [4] developed a multi-objective optimization approach using second-
order polynomial response surface model and evolutionary algorithm (EA) which
was used in the redesign of a turbopump and the NASA rotor67 blade. Verstrate &
Van den Brasembussche [5] combined a genetic algorithm (GA) and ANN to find a
compromise between the conflicting demands of high efficiency and low centrifugal
stresses for micro gas turbine blades. Pierret [6] and Pierret et al. [7] used a GA
and RBF network for shape optimization of 3D compressor blades.

An alternative approach is to use an inverse design based optimization strategy as


used for compressors in [8]. The advantage of such an approach is that a large
design space can be explored with a small number of design parameters. Such an
approach can be coupled with Design of Experiments method and appropriate
surrogate models such Response Surface Model or Kriging, see [9] for further
details. One major advantage of the inverse designed based approach is that all the
blade geometries generated in the design matrix will satisfy the specified turbine
specific work at the specified mass flow rate, something that is difficult to control
with direct design optimization. This will make it easier to achieve more accurate
response surface, which can then be used for rapid evaluation of multi-
disciplinary/multi-objective performance parameters required for optimization.

The aim of the present paper is to present a design method integrated with DOE
(design of experiment), inverse design and RSM in the multidisciplinary
optimization of turbocharger turbines.

2 OPTIMIZATION METHODOLOGY

The flowchart of the optimization method that will be developed as part of the
current project is shown in Fig. 1.

400
Figure 1 Flowchart of the optimization methodology

The design and performance parameters are selected and denoted by xi and yj
respectively. The performance parameters yj can be expressed by function of design
parameters xi:

= , = 1, 2, … , = 1, 2, … (1)

Where n and m are the total number of design and performance parameters. In this
paper, the design parameters are parameters used to define the blade geometry
which will be discussed in the blade parameterisation section. The performance
parameters are used to evaluate the blade aerodynamic and mechanical
performance which will be covered in the CFD and structural analysis section.

The first order polynomial RSM approximates the objective parameters yj with first
order polynomial:

= +∑ (2)

There are total (n+1) unknown coefficients in the eq. (2) The polynomial
coefficients ai can be determined by a standard least-square regression. A linear
DOE with at least (n+1) designs is needed to provide enough sampling points to
solve the ai. CFD and FEA calculations are performed for these designs to get the
performance parameters yj.

Once the linear RSM model is available, by analysing and comparing the coefficients
ai, we can eliminate non-essential design parameters compared to essential
parameters and reduce the number of design parameters n.

401
3 BLADE PARAMETERISATION

The commercial software TURBOdesign1 [10] is used to parametrically describe the


blade geometry. TURBOdesign1 (see [11]) is a three-dimensional inviscid inverse
design method, where the distribution of the circumferentially averaged swirl
velocity (rVθ) is prescribed on the meridional channel of the blade and the
corresponding blade shape is computed iteratively.

The circulation distribution is specified by imposing the spanwise rVθ distribution at


leading and trailing edge and the meridional derivative of the circulation d(rVθ)/dm
(blade loading) inside the blade channel. The input design parameters required by
the program are the following:

 Meridional channel shape in terms of hub, shroud, leading and trailing


edge contours.
 Normal/tangential thickness distribution.
 Fluid properties and design specifications.
 Inlet flow conditions in terms of spanwise distributions of total temperature
and velocity components.
 Exit rVθ spanwise distribution. By controlling its value, the work coefficient
is controlled.
 Blade loading distribution d(rVθ)/dm. It is imposed at two or more span
locations. The code then automatically interpolates span-wise to obtain the
two-dimensional distribution over the meridional channel.
 Stacking condition. The stacking condition must be imposed at a chord-
wise location between leading and trailing edge. Everywhere else the blade
is free to adjust itself according to the loading specifications.

The baseline geometry is a radial turbine impeller for heavy duty diesel application
with eleven blades. The blade has radial filaments as shown in Fig. 2. Due to
external constrains (dimensions of nozzle and volute), the number of blades, blade
tip diameter and blade length are constant in the optimization process.

Figure 2 Baseline Impeller Wrap Angle Contour

The blade loading is defined as d(rVθ)dm. In Fig. 3 three segment method used to
define the blade loading in TURBOdesign1 is illustrated. This method divides the
loading distribution into three separate parts by NC and ND. Parabolas are used to
define the distribution for the first and last segment and the middle segment is

402
specified by a straight line with a given slope. Using this parameterization, one can
fix the value of NC and ND and still obtain large variation in blade shape by
changing the Slope and the DRVT parameters. Changes to Slope parameter can
change the loading from fore-loaded to aft-loaded, while changes to DRVT value at
the leading edge will change the blade incidence. Typically by defining blade loading
in this way at two sections (hub and shroud) a very large variation in impeller
shape can be obtained. Hence, 4 design parameters is all that is needed to
represent the variation of in blade shape for a fixed meridional shape. However, in
this study an initial linear DOE will be performed and hence variations in NC and ND
will also be included in the optimization. In addition to blade loading the spanwise
work distribution or rVθdistribution can be modified. At leading edge the spanwise
rVθis kept fixed but trailing edge the spanwise rVθis modified. This adds additional
two parameters to the optimization.

In TURBOdesign1 a methodology is used for meridional shape parameterization,


which is outlined in Fig. 4, based on 6 parameters shown in red. These parameters
are inlet and exit width W1 and W2. Impeller leading edge and trailing edge angles
and and two control parameters Y_hub and Y_shr that affect the meridional
shape through the cubic spline method used to create the meridional shape.

Figure 3 Three segment method used to define the blade loading

Figure 4 Meridional plane parameters

Fig. 5 shows possible configurations and variation of meridional plane in our design
space.

403
Figure 5 Meridional plane variation of the design space

The blade thickness is controlled by one parameter called thickness factor which is
between 0.9 and 1.3. Base on the baseline thickness distribution, the shroud
thickness remains same and the hub thickness is multiplied by the factor. The
thickness between hub and shroud is interpolated proportionally.

So in total for the initial design table 17 design parameters were used and their
variation ranges are shown in Table 1.

Table 1 Variation ranges of design parameters

Design parameter Range


W1 7 - 11 (mm)
W2 15 – 24 (mm)
 0 - 40 (°)
 0 - 10 (°)
Y_Hub 16.5 – 21 (mm)
Y_Shr 0.2 – 0.4
RVT_TE_Hub 0 – 0.04
RVT_TE_Shr 0.06 – 0.1
NC_Hub 0.05 – 0.2
NC_Shr 0.05 – 0.4
ND_Hub 0.6 – 0.85
ND_Shr 0.6 – 0.85
Slope_Hub 1 – 2.5
Slope_Shr -1 - -5
DRVT_LE_Hub -1 - -0.1
DRVT_LE_Shr -1 - -0.1
Thickness 0.9 – 1.3

4 CFD ANALYSIS

When the blade geometry is ready the CFD calculations are performed by using
commercial software ANSYS CFX-14.5. For saving computational time only one flow
passage is modelled in the CFD which consists of a nozzle and a rotor as shown in
Fig. 6. The total number of element is about 1,000,000. The inlet boundary
condition are total pressure and total temperature. The flow direction at the inlet is
specified by xyz components. Atmospheric pressure is applied in the outlet.
Turbulence model is set as k-e. The nozzle domain is stationary and rotor domain is
rotating with a RPM ranging from 50,000 to 90,000 rpm. Periodic boundary
conditions are also applied. T-S Efficiency (Eff_50k, Eff_60k, Eff70k, Eff_80k and
Eff_90k) and turbine mass flow parameter (MF_50k, MF_60k, MF_70k, MF_80k,
MF_90k) for five different RPM are used to evaluate the turbine aerodynamics
performance and flow capacity.

404
Figure 6 Computational domain

5 STRUCTURAL ANALYSIS

ANSYS Workbench 14.5 is used to perform structural analysis and modal analysis to
get the maximum principal stress and natural frequencies. To get more accurate
results, whole wheel geometry is generated by ProEngineer (a popular CAD
software) based on one blade. It is found that the stress concentration always
occurs near the hub fillet part (Fig. 7) and increasing of the fillet radius will reduce
the local stress concentration. Therefore, the mesh in the hub fillet is refined as
shown in the Fig. 7. The first mode shape for vibration is always exducer flex (left
in the Fig. 8) and the second mode shape is inducer and exducer flex (right in the
Fig. 8). The maximum principle stress and 1st and 2nd order frequencies are chosen
as performance parameters.

Figure 7 Mesh used in the structural analysis and stress evaluation

405
Figure 8 Modal analysis

At the same time, moment of inertia and throat area which are other two
performance parameters can be easily obtained from ANSYS Workbench. Finally
there are 15 performance parameters.

6 LINEAR DOE AND SENSITIVITY STUDY

As shown in the section 3 we have in total 17 design parameters. To construct a


linear (first order polynomial) response surface at least 18 different designs are
necessary. In case some designs may fail to generate reasonable geometry, 25
designs in total are generated by DOE component in the ISIGHT5.6 (Latin
Hypercube method) based on specified design parameter ranges. 19 of 25 designs
are converged and generate blade geometry by using inverse design code
TURBOdesign1. CFD and FEA analysis are performed for these 19 designs. The
range of the specific work for these 19 designs (RPM = 70k) is 57.3 – 66.85 kJ/kg
versus baseline value of 65.1 kJ/kg. The specific work is defined as turbine power
divided by mass flow rate which can be obtained from CFD results. All the design
parameter and performance data are imported into ISIGHT5.6 to performed a linear
polynomial regression.

=∑ + (3)

Where is the ith design parameter (normalised to 0-1) and is the jth
performance parameter (normalised to 0-1).

Table 2 shows the R2 value of linear RSM model for different performance
parameters.

It can be seen that the error between the actual value and predicted value through
the RSM model is very small due to the high R2 values (> 96%).

406
Table 2 R2 of the linear RSM model

R2
Eff_50k 0.9895
Eff_60k 0.9896
Eff_70k 0.9802
Eff_80k 0.9749
Eff_90k 0.9685
MF_50k 0.9991
MF_60k 0.9989
MF_70k 0.9981
MF_80k 0.9971
MF_90k 0.9973
Stress 0.9977
1st_frequency 0.9991
2nd_frequency 0.9882
MOI 0.9984
Throat_area 0.9863

To compare the effect of different design parameters (xi) on the performance


parameter y, the coefficients is divided by the max | | and multiplied by 100 to
get Ai (as shown in Table 3).

= × 100 (4)
(| |)

Table 3 Comparison of normalised coefficients Aji

The range of Ai is from -100 to 100. For a particular performance parameter, the
greater of the absolute value of Ai is, the more significant the corresponding design
parameter is. When Ai is greater than 0, it means the performance parameter will
increase with the increasing of this design parameter. When Ai is less than 0, it
means the performance parameter will decrease with the increasing of the design
parameter.

For the MF_50k, MF_60k, MF_70k, MF_80k and MF_90k, the most significant
parameter is W1 since the values of Ai are 100. It can be easily understood by
seeing the Fig. 4 increasing of W1 will result in increase of nozzle inlet area and
more flow can enter the nozzle. Therefore, the flow capacity of the turbine will be

407
improved. For the 1st_frequency, if we increase the W2 which is the blade height at
the outlet the blade will become higher and easier to vibrate. Therefore, the blade
stiffness is reduced and 1st_frequency and 2nd_frequency will drop. Similarly for the
MOI, increasing of W2 and Y_Hub will increase the blade inertia and wheel hub
inertia. Therefore, the total MOI will increase. It is well-known that the peak
efficiency of a radial turbine () occurs when the U/C (velocity ratio or jet
speed ratio) = 0.7 corresponding to RPM = 70k in our design problem. Our CFD
calculations show that with the increasing of mixed-inflow turbineU/C with
peak efficiency moves towards higher RPM (80 or 90k). This can also be explained
by the Ai value of for Eff_80k and Eff_90k are 100.

The 7 most significant design parameters are selected through summation of all the
absolute value of Ai which are α1, NC_Hub, ND_Hub, Slope_Hub, W1, W2 and
Y_Hub. Since the variation of these 7 design parameters will have a much larger
effect on the performance parameter compared to the other design parameters.

Two designs are generated for these 7 selected design parameters based on Table 3
to validate this method of parameters reduction. The aims of design 1 and design 2
are to maximize Eff_70k and minimize MOI (moment of inertia) separately. By
comparing Ai values in row Eff_70k from Table 3 if we want to maximize Eff_70k we
need to maximize NC_Hub, W2 and minimize ND_Hub, Y_Hub in the design
space. Similarly if we want to minimize MOI we need to maximize and minimize
W2, Y_Hub in the design space. The values of other design parameters will be set
same as the baseline value or intermediate value in the Table 1. The resulting
meridional shapes of design1 and design2 are shown in Fig. 9 and compared with
the baseline impeller, shown in blue. Design1 has a mixed flow inlet with wider inlet
width and with almost the same exducer height as the baseline impeller. However,
design2 has a significantly narrower exducer height. The comparison of the total-
static efficiency and mass flow parameter (normalized by the baseline efficiency
and flow parameters at 70k rpm and 50k rpm) is shown in Fig. 10. The results
confirm that design1 results in 2-5 points improvement in efficiency as compared to
the baseline impeller and also about 5-7% increase in flow capacity. But as shown
in Table 4, design1 has 54% higher stress, 2.54% increase and 31% reduction in
modal frequencies. Design2 results in 30% reduction of inertias as compared to
baseline but almost 15 % loss of efficiency versus the baseline design.

The results confirm clearly that any optimization of the turbine would require multi-
objective optimization approach that can find the best tradeoff between these
contrasting requirements. Work is currently underway to develop such an approach
for radial and mixed flow turbines.

Figure 9 Comparison of Meridional Plane

408
Figure 10 Comparison of T-S Efficiency and Mass Flow Rate (normalised)

Table 4 Structural performance compared to the baseline value

Moment of Inertia Stress 1st frequency 2nd frequency


Design 1 2.54% 55.4% -31.4% -31.0%
Design 2 -30.2% -38.9% 33.8% 11.2%

7 CONCLUSION

A detailed methodology about blade design and optimization has been shown in this
paper, for multidisciplinary design of radial and mixed flow turbines for
turbochargers. The methodology couples a 3D inverse design method, design of
experiments and linear response surface model in order to find the most important
design parameters that affect the efficiency and mass flow parameter at multiple
operating points, maximum stress and 1st an 2nd modal frequencies and impeller
moment of inertia. A DoE table consisting of 25 configurations is generated by
using the 3D inverse design method and using 17 different design parameters that
affect the meridional shape, the blade loading, the spanwise work distribution and
the thickness distribution. Each configuration was then analysed in CFD at multiple
operating points and FEA for structural and modal analysis. The results was then
used in a sensitivity analysis using a linear response surface to find the 7 most
significant design parameters. By setting these 7 parameters appropriately two
designs were created one for higher efficiency and one for lower inertia. The
performance parameters for these two designs were obtained by CFD and FEA and
generally they confirm that the sensitivity information provides the correct direction
in terms of the design objectives. However, the results also indicated that design1
that improves efficiency also results in increase in stress and reduction in modal
frequencies. A full optimization of turbine would require a proper multi-objective
optimization.

In future work, Kriging approximation will be used to construct a more accurate


model based on the new 7 design parameters using a higher number of
configurations. A Pareto Front will be generated by applying MOGA on the Kriging
model with multiple objectives and constraints. By using this approach it should be
possible to obtain the final optimized geometry that meets the contrasting
requirements of aerodynamics, stress, vibration and inertia.

409
ACKNOWLEDGEMENTS

This research is sponsored by Cummins Turbo Technologies and EPSRC. The author
would like to thank their financial support.

REFERENCES

[1] Roclawski, H., Bohle, M. & Gugau, M. 2012. Multidisciplinary Design


Optimization of a Mixed Flow Turbine Wheel. Proceedings of ASME Turbo
Expo 2012. GT2012-68233.
[2] Chahine, C., Seume, J. R. & Verstraete, T. 2012. The Influence of
Metamodeling Techniques on the Multidisciplinary Design Optimization of a
Radial Compressor Impeller. ASME Turbo Expo, GT2012-68358.
[3] Muller, L., Alsalihi, Z. & Verstraete, T. 2013. Multidisciplinary Optimization
of a Turbocharger Radial Turbine. Journal of Turbomachinery, Vol 135.
[4] Lian, Y. & Liou, M. 2005. Multiobjective Optimization Using Coupled
Response Surface Model and Evolutionary Algorithm. AIAA Journal, 43(6):
1316-1325.
[5] Verstraete, T., ALsalihi, Z. & Van den Braembussche, R. A. 2007.
Multidisciplinary Optimization of a Radial Compressor for Micro Gas Turbine
Applications. ASME Turbo Expo 2007.
[6] Pierret, S. 2005. Multi-objective Optimization of Three Dimensional
Turbomachinery Blades. European Conference for Aerospace Sciences.
[7] Pierret, S., Filomeno Coelho, R. & Kato, H. 2007. Multidisciplinary and
Multiple Operating Points Shape Optimization of Three-dimensional
Compressor Blades. Struct. Multidisc. Optim. 33: 61-70.
[8] Bonaiuti, D. & Zangeneh, M. 2009. On the Coupling of Inverse Design and
Optimization Techniques for the Multiobjective, Multipoint Design of
Turbomachinery Blades. Journal of Turbomachinery, Vol 131.
[9] Queipo, N. V., Haftka, R. T., Shyy, W., Goel, T., Vaidyanathan, R. & Tucker, P.
K. 2005. Surrogate-based analysis and optimization. Progress in Aerospace
Sciences, 41: 1-28.
[10] TURBOdesign1 - Advanced Design Technology Ltd, 2013.
[11] Zangeneh, M. 1991. A 3D Design Method for Radial and Mixed-Flow
Turbomachinery Blades. Journal of Numerical Methods in Fluids, 13: 599-
624.

410
Numerical investigation on the high-cycle
pressure fluctuation mechanism of VNT
rotor
X Shi, C Yang, Y H Liu, S T Liu, B Zhao
School of Mechanical Engineering, Beijing Institute of Technology, China

ABSTRACT

Due to the rotor-stator interaction of Variable Nozzle Turbine (VNT), the strong
excitations at the leading edge of turbine rotor can be created during engine braking
down or transient acceleration. This may be a concern of high-cycle fatigue failure of
VNT rotor blades. In this paper the 3-D unsteady numerical simulation of a VNT was
performed to investigate the effects of nozzle vanes clearance leakage, wake and
shock wave on the rotor blade pressure fluctuations, which can specify where and how
these three factors influence the excitations of the turbine rotor blade. The results
show that the shock wave and nozzle vanes clearance leakage are the main reasons
for rotor blade surface pressure fluctuation. The shock wave only affects the suction
surface at the leading edge of rotor blade because of the shape of rotor blade. When
the nozzle vanes opening was closed small, the shock wave and the amplitude of the
pressure fluctuations on the rotor blade surface were greatly increased compared to
that for full open condition. In the regions near the hub and shroud side on the rotor
blade suction surface, the magnitude of the pressure fluctuation is increased
significantly by the influence of the nozzle vanes clearance leakage. Through FFT
analysis of the pressure fluctuation at the rotor blade leading edge, it can be seen that
the fluctuation amplitude increased significantly on the rotor passing frequency and
this also implies the possibility of VNT high-cycle fatigue failure with the
high-frequency excitations.

1 INTRODUCTION

Turbocharger has been widely used in diesel engines due to it has opportunities to
improve engine power and fuel economy as well as reduce emissions[1,2]. In the last
two decades, stronger emphasis from emissions legislation was forcing automotive
manufactures to improve part load performance of the engine. As a result, a better
method to meet this requirement is the use of a variable geometry turbine (VGT)
which can increase pressure ratio and adjust flow angle at low mass flow condition by
varying throat area of nozzle. But the strong excitations at the leading edge of turbine
rotor can be created during engine braking down or transient acceleration. This may
be a concern of high-cycle fatigue failure (HCF) of VNT rotor blades.

Leakage and wake from nozzle vanes have significant impact on flow distribution at
inlet of downstream rotor, turbine performance and unsteady load of rotor blade[3,4].
When the nozzle vanes was closed to small openings and turbine operates at high
pressure ratio, shock wave appears on the suction side of nozzle vane near trailing
edge and induces high pressure response on the surface of blade by impinging at the
leading edge[5,6]. Thus leakage, together with wake and shock wave, contribute to

_______________________________________
© The author(s) and/or their employer(s), 2014
411
HCF problem of downstream rotor. As their features are different each other, it is
necessary to know in detail the progress that each of them induces highly
aerodynamic loading on turbine wheels.

In this paper a 3-D unsteady numerical simulation of a VNT was performed to


investigate the effect of nozzle vanes clearance leakage, wake and shock wave on the
rotor blade pressure fluctuations, and then specify where and how these three factors
influence the excitations of the turbine rotor blade.

2 DESCRIPTION OF TURBINE

The variable geometry turbine (VGT) used in this investigation is originated from a
turbocharger with mixed rotor. The volute was not included in the computational
domain because of current project researches the interaction between nozzle vanes
and downstream rotor. Thus there are 9 nozzle vanes and 13 rotor blades in this
researching turbine stage. Each of nine nozzle vanes equips with two clearances with
2% of spanwise height of vane. They locate respectively on the hub and shroud sides.
Rotor blade’s tip gap locates only on the shroud side and is about 0.35 mm. Throat
area of the nozzle vane can be changed by rotating the adjusted shaft, this means that
the nozzle vanes’ opening condition is various.

3 NUMERICAL METHODOLOGY

3.1 Computational grid


Domain scaling method is used for geometry configuration and computational mesh to
reduce requirement of CPU time and space of memory. A method to applied domain
scaling method is to change number of rotor blades from 13 to 12 to meet the required
condition that both sides of rotor-stator interface are same periodicity. After scaling
the computational domain, only 1/3 of blade rows consisted of three nozzle vanes and
four rotor blades was modeled in this paper. Steady simulations showed that the
changed blade number causes a small deviation in turbine performance. The
computational grid was generated automatically by IGG/Autogrid5. Totally 2.39
million cells are used for the numerical simulation.

3.2 Numerical methods


The EURANUS flow solver integrated in NUMECA Fine/Turbo is employed in this work.
Three dimensional time-averaged Navier-Stokes (RANS) equations were solved. Cell
centered control volume approach is used for central space. Differential scheme and
Jameson type of artificial dissipation items are used for scheme stability. For the
artificial coefficients, 1.0 and 0.01 are taken separately for 2nd and 4th order items.
A 4-stage Runge-Kutta integration scheme is used for time discretization in
steady-state simulation. CFL number of 3.0 is used in all the computations. In current
case, flow is supposed to be fully turbulent and one equation Spalart-Allmaras model
is used for closure of RANS equations. For the unsteady computations, Jameson’s dual
time step method is used. 120 physical time steps are set for the computational
domain. The rotating speed of rotor is 120 krpm.

3.3 Boundary conditions and convergence


Total pressure, total temperature as well as flow angle calculated according to volute
are imposed at the nozzle inlet. Static pressure was specified at the rotor outlet
boundary with radial equilibrium. Adiabatic condition is set for all the solid walls. Only
downstream rotor blade surface and hub rotate, and other surfaces are stationary. For
steady computation, the conservative coupling by pitchwise rows was used at the
rotor/stator interface, which guaranteed an exact conservation of mass flow,
momentum and energy through the interface. This approach adopts the same

412
coupling procedure for all the nodes along the circumferential direction, even if the
local flow direction is different from the average one. For the unsteady computation,
a direct interpolation on sliding meshes was employed at the rotor-stator interface.

3.4 Verification on numerical results


Only steady CFD method was used for validation by comparing the numerical results
with experimental data. Figure 1 show the comparisons of turbine performance which
is from CFD and experiments respectively. The predicted mass flow of turbine is about
3.5 present higher than the test pressure ratio. This is mainly due to the difference
between computational domain and test hardware. The volute was not included in the
simulation model. As this, the flow loss in volute was not calculated, resulting in that
predicted efficiency is about 2.5 percent higher than tested data. Despite of this
discrepancy, current CFD method which is able to predict well turbine performance
can be applied to assist turbine analysis.

0.70 5.0
Cal Cal
Exp 4.5 Exp
0.65
4.0

Pressure ratio
0.60
Efficiency

3.5
0.55
3.0

0.50 2.5

2.0
0.45
1.5
0.40
0.4 0.5 0.6 0.7 0.8 0.01 0.02 0.03 0.04 0.05 0.06
U/C Dimensionless flow rate
Figure 1 Comparison of predicted turbine performance with test result

4 ANALYSIS AND DISCUSSION

4.1 Comparison of shock wave for two kinds of opening


Figure 2 shows the distribution of static pressure with isoclines at mid-span for two
kinds of open conditions. From these two pictures it can be seen: for large opening
only one shock wave stands on the suction side of nozzle vane while two shock waves
appear in small opening, one locates at throat of nozzle and another one stands on
suction surface of nozzle vanes near trailing edge. Notice that shock wave can impinge
at the leading edge of downstream rotor blade for any opening. Compared with large
opening, however, the shock wave stood on the suction side of nozzle vane is much
stronger in small open condition. Thus stronger interaction between shock wave and
blade occurs when nozzle vanes are adjusted to small opening.

Figure 2 Distribution of static pressure at mid-span

413
When the turbine operates in small opening, the flow between nozzle vanes and
downstream rotor is highly unsteady. As downstream rotor blade passes, the intensity
of shock wave changes, as indicated in Figure 3. The shock wave generates when the
smallest width between nozzle vane and rotor forms, as indicated at time of t1 (t1~t5
correspond to different instantaneous time). At this time, the strength of shock wave
increases nearly to maximum and the rotor blade suction surface near leading edge
experiences serious interaction from shock wave, possibly leading to strong pressure
response in this region. As the leading edge of rotor blade leaves away, the shock
wave begins to weaken. This progress is clearly shown in Figure 3. Another thing
should be noticed is different with axial turbine, the inlet of rotor blade is radial and
the shock wave can only impinge a small area near the leading edge of suction surface.
It has little effect on pressure side of downstream rotor blade. As known, the strong
interaction between shock wave and leading edge of rotor is easy to induce high cycle
fatigue (HCF) of downstream rotor because of highly unsteady pressure response
must have been induced when shock wave impinges at the blade.

t1 t2 t3 t4 t5
Figure 3 Interaction between shock wave and rotor, at small opening

4.2 Nozzle vanes’ leakage


The nozzle vane’s leakage flow increases as the nozzle opening is closed small. This
fact has been indicated by many researchers[3,7-9]. So only small opening in this
paper is used for the analysis of the interaction between nozzle vane’s leakage and
downstream rotor. The air flows from vane’s pressure side (convex surface) to suction
side (concave surface) and then forms leakage in the downstream region of nozzle
vane. The leakage rolls up and then forms vortex, resulting in higher entropy appears
in the center of leakage flow. Moreover, during downstream transportation, the
leakage continuously mixes with main flow and leads more flow loss, so the width of
region with high entropy gradually increases. These features have been shown well by
distribution of entropy in Figure 4.

In addition to mixture with surrounding flow, the leakage can also influence
downstream rotor like shock wave by the mean of hitting at leading edge. When the
leakage hits at the leading edge of downstream rotor, the leakage was cut into two
parts. Downstream part transports fast along suction surface of rotor blade,
meanwhile, more flow loss was induced on the suction side of rotor blade, such as the
region marked with A. Another part interacts with boundary layer on the pressure side
of rotor blade. This progress has been shown well by several patterns signed with B,
C and D in Figure 4. It should be noticed that on the pressure side of blade
transportation speed of leakage is not as high as that on suction side. As a result, the
time nozzle vane’s leakage interacts with pressure side of blade is longer. Besides, the
leakage flow from nozzle vane not only changes the flow angle entering rotor, but
influences the development of boundary layer of blade as well. Meanwhile, the
pressure distribution on the blade surface must be changed. This is why nozzle vane’s
leakage influences HCF of downstream rotor.

Figure 5 shows the distribution of entropy at midspan for small opening. Wake
generated by nozzle vanes flows circumferentially and experiences long distance
before entering into downstream rotor channel. This long track wake transports

414
provides more time to mixture between wake and surrounding air. Thus wake is weak
when it hits at leading edge of blade, and its influence on the pressure response on the
surface of blade is not as serious as shock wave and leakage flow. Moreover, the
longer distance wake transports, the more flow loss mixture leads. According to this
phenomenon, the small space between nozzle vanes and downstream rotor should be
available to improve the turbine performance. But, on other hand, the small space
must enhance the aerodynamic interaction and then increase risk of HCF of rotor. Thus
this contradiction should be taken into account in design progress.

Figure 4 Interaction of nozzle vane’s leakage and rotor, at 90% span

Figure 5 Distribution of entropy, at midspan

4.3 Aerodynamic responses on downstream rotor


From above analysis, it has been known that the shock wave, together with nozzle
vane’s leakage, interact with downstream rotor when turbine operates at small
opening. Even though blade force is not the only parameter affecting the force
response [10], those strongly aerodynamic interactions still influence seriously HCF of

415
blade by mean of inducing pressure response on blade surface. Figure 6 shows the
root mean square (RMS) of pressure fluctuation on both pressure and suction sides of
the rotor blade for small opening. The larger RMS value is, the stronger pressure
fluctuation is. From these two pictures can be seen that the regions with high RMS
mainly concentrate in the inlet of blade. In other word, the pressure in the regions
near leading edge is more sensitive to nozzle vane’s influences. This is due to the fact
that shock wave and leakage mainly contact the leading part of blade. Compared with
pressure side, stronger pressure fluctuation appears on the suction side near leading
edge.

For the suction side RMS pressure distribution, due to nozzle vane equips with two
gaps which locate respectively on hub and shroud sides, the leakage flow from both
gaps leads to two regions with high RMS near hub and shroud. As analysis above,
strong shock wave appears and impinges at the leading edge of blade. It results in
large region with high RMS appears near midspan location. This must increases the
risk of blade fatigue.

(a) pressure surface (b) suction surface


Figure 6 RMS pressure distribution on the blade surface, small opening

4.4 Distribution of pressure fluctuation


RMS of pressure is able to reflect information about pressure fluctuation on the blade
surface, but it includes all information that pressure fluctuates at many frequencies.
Based on this fact the Fast Fourier Transformation (FFT) method has an advantage to
reflect pressure response at any one frequency. Harmonic analysis was carried out to
calculate the pressure response at several monitoring points on the blade surface.

Pressure spectrum result on the leading edge of rotor blade at midspan is shown in
Figure 7. From this result it can be known that dominant frequency is found at
frequency of 18 kHz, which is the same as nozzle vane passing frequency (VPF), with
pressure pulsation amplitude being about 570 kPa. Subordinate signal is found at
frequency of 36 kHz, which is twice of the VPF, with pressure pulsation amplitude only
being about 140 kPa. Other signal with smaller pulsation amplitude can also be found
at frequency of 54 kHz, which is three times of the VPF. This fact implies that VPF is
most dominant frequency among several excited frequencies if VNT suffers from
high-cycle fatigue failure due to the high-frequency excitations. It is also noted that
the VPF is associated with vane number. Bauer et al.[11] had applied numerical
methods to examine the aerodynamic excitation of the blades in radial turbines with
guide vanes, and the results revealed that unsteady loads appear significantly higher
for 12 compared to 22 vanes. For the thin nozzle vane, its number can be adjusted in
a large scale while few selections about the vane number can be provided for thick
vane. This is the reason that only one kind of nozzle vane number was modeled and
commutated in current article.

416
Figure 7 Pressure spectrum at rotor leading edge, mid span

Through FFT analysis of the pressure fluctuation at six predicted points on the both
surfaces of rotor blade at midspan, the amplitudes of pressure pulsation at frequency
of VPF were obtained and shown in Figure 8. A fact along meridian direction the
magnitude decreases gradually can be seen. This means that nozzle’s influence on the
downstream rotor weakens gradually along the flow path. In addition, the amplitudes
of pressure pulsation at VPF on the suction side are far larger than that on pressure
side. This fact also indicates that nozzle vanes have more influence on the suction
surface of rotor blade.

Illustration of monitoring points

600 600

500 500
Amplitude /kPa

Amplitude /kPa

400 400

300 300

200 200

100 100

0 0
1 2 3 4 5 6 1 2 3 4 5 6
Location of monitoring points location of monitoring points
Pressure side Suction side
Figure 8 Pressure pulsation magnitudes at nozzle VPF at monitoring points

5 CONCLUSIONS

Shock wave and nozzle vanes clearances’ leakage are the dominant reasons for
pressure response on rotor blade surface, and the influence from nozzle vane’s wake
is not as strong as shock wave and leakage because of it mixes fully with surrounding

417
flow during transportation in the space between two rows. When the nozzle vanes
locate at small open condition, the shock wave increases significantly compared with
large open condition. Moreover, shock wave mainly influences the pressure response
on suction surface of blade near leading edge. In the regions near the hub and shroud
side, the pressure fluctuation at leading edge responded to leakage from nozzle
vane’s clearance is strong at small opening. Through FFT analysis of the pressure
fluctuation on the rotor blade surface, it can be known that the dominant frequency of
pressure pulsation is of same as the nozzle vane passing frequency.

This article is the first part of our researching project. The CFD/FEA coupled forced
response analysis had been done, but it was not included in this paper due to the
limited length of paper.

ACKNOWLEDGMENT

This research was supported by Ford Motor Company and National Natural Science
Foundation of China (NSFC, 51276018). The authors would like to acknowledge their
financial support and Dr. Harold Sun of Ford for his strong technical support.

REFERENCES

(1) Watson, N., Janota, M.S., 1982, “Turbocharging the internal combustion engine”,
Wiley Interscience, New York, 1982.
(2) Baines, N.C., 2002, “Radial and mixed flow turbines options for high boost
turbocharger” 7th International Conference on Turbocharger and Turbocharging.
(3) Hu, L., Sun, H., Yi, J., et al, 2013, “Investigation of Nozzle Clearance Effects on a
Radial Turbine: Aerodynamic Performance and Forced Response”, 2013 SAE
International, 2013-01-0918.
(4) Hu, L., Yang, C., Sun, H., et al, 2011, “Numerical Analysis of Nozzle Clearance’s
Effect on Turbine Performance”, Chinese Journal of Mechanical Engineering,
2011; vol. 24, 4: pp. 618-625.
(5) Chen, H., 2006, “Turbine wheel design for Garrett advanced variable geometry
turbines for commercial vehicle applications”, 8th international conference on
turbochargers and turbocharging, Institute of Mechanical Engineers, 2006,
London UK.
(6) Kawakubo, T., 2010, “Unsteady rotor-stator interaction of a radial inflow turbine
with variable nozzle vanes”, Proc. Of ASME Turbo Expo, GT2010-23677.
(7) Walkingshaw, J. R., Spence, S.W., Thornhill, D., et al, 2012, “An experimental
assessment of the effects of stator vane clearance location on an automotive
turbocharger turbine”, 10th international conference on turbochargers and
turbocharging, Institute of Mechanical Engineers, 2012, London UK.
(8) Spence, S.W., Neill, J. W., Cunningham, G., et al, 2006, “An investigation of the
flow field through a variable geometry turbine stator with vane endwall
clearance”, Proceedings of the Institution of Mechanical Engineers Part A:
Journal of Power and Energy, 2006, vol. 220: pp. 899-910.
(9) Tamaki, H., Goto, S., Unno, M., “The effect of clearance flow of variable nozzle on
radial turbine performance”, Proc. Of ASME Turbo Expo, GT2008-50461.
(10) Schwitzke M., Schulz A., Bauer H., “Prediction of high-frequency blade vibration
amplitudes in a radial inflow turbine with nozzle guide vanes”, Proc. Of ASME
Turbo Expo, GT2013-94761.
(11) Bauer H., Schulz A., Schwitzke M., “Aerodynamic excitation of blade vibrations in
radial turbines”, MTZ worldwide, 2013, Vol. 74, 6: pp. 48-54.

418
Thermal analysis of small high-speed
floating-ring journal bearings
D Porzig1, H Raetz2, H Schwarze1, J R Seume2
1
Institute of Tribology and Energy Conversion Machinery,
Technical University of Clausthal, Germany
2
Institute of Turbomachinery and Fluid Dynamics,
Leibniz Universität Hannover, Germany

ABSTRACT

In the present paper, the impact of thermal boundary conditions on characteristic


parameters of full floating-ring journal bearings (FRB) typically used in small
automotive turbochargers (TC) is discussed. A complex thermo-hydrodynamic
simulation model is introduced which is validated by experimental data under real
TC operating conditions on a hot-gas test stand. The conducted study clarifies the
significant impact of the shaft temperature on overall bearing operating parameters
which cannot be neglected. Compared to using an adiabatic shaft model, the
accuracy of ring-speed prediction is considerably improved by a non-adiabatic
approach.

NOTATION

A sliding surface area Q lubricant flow rate


Bi, Bo ring inner / outer width Ri, Ro ring inner / outer radius
Di, Do ring inner / outer diameter Tsup TC lubricant supply temperature
F film fill factor Tshaft shaft temperature boundary condition
h film thickness η mean effective lubricant viscosity
nshaft TC shaft speed ρoil lubricant density
psup TC lubricant supply pressure τD drag torque
∆p pressure loss φ circumferential coordinate (angle)
pl,i inner film supply pressure ωR, ωS ring and shaft angular speed
P film pressure field Ω ring speed ratio (Ω = ωR/ωS)

1 INTRODUCTION

Today, turbocharging the internal combustion engines of passenger cars is one of


the most promising technologies to increase the power output, improve the engine
efficiency and keep the emissions below the legal regulatory limits. The rotors of
turbochargers used in passenger vehicles are typically supported by two journal
bearings of the floating ring type which consist of a short cylindrical bushing placed
between the shaft and the turbocharger housing forming two thin fluid films in
series (Figure 1). This bearing system lubricated with internal combustion engine oil
has proven to be robust, reliable and very cost-effective due to its simplicity and

_______________________________________
© The author(s) and/or their employer(s), 2014
421
ease of manufacturing [1]. In comparison with a semi-floating ring bearing (SFRB),
where the rotation of the ring is inhibited by an axial pin, a full-floating bearing has
the advantage of reducing the rate of shear in the fluid films, thus reducing the
total frictional power loss [2]. Integrated with an internal combustion engine, the
turbocharger’s bearing friction plays an important role for overall system efficiency
and behavior because friction leads to an inertial lag upon acceleration and
produces backpressure on the engine caused by the turbine wheel [1]. In
automotive turbochargers, the bearing system is exposed to a difficult thermal
environment [1]. Besides the shear-driven frictional losses within the fluid films
acting as heat sources, the turbine wheel is also exposed to hot exhaust gas from
the engine which reaches temperatures of 600°C in modern Diesel engines and
higher temperatures in modern Otto-cycle engines. This results in non-uniform
temperature distributions in the shaft, the TC casing, and the bearing system itself
which is complicated to analyze because of the multiplicity of different heat flow
paths [3, 4].

Current automotive TC performance prediction tools include diabatic TC models [5,


6] using lumped-parameter 1D approaches for the thermal energy flows to improve
TC operating predictions mainly in part load and transient simulations. Bohn et al.
[7, 8, 9] used a 3D Conjugate Heat Transfer (CHT) model with experimental
validation to advance the understanding of the heat balance in automotive
turbochargers but without direct coupling to a bearing friction model including heat
flux.

Figure 1 - Sectional view of a typical turbocharger for passenger vehicles


and schematic illustration of the full-floating ring bearing system

Early investigations of (semi) floating ring bearings (summarized in [10, 11]) were
limited to experiments at low shaft speeds used for verification of isothermal
analyses only. Ring speeds predicted by isothermal flow models exhibit poor
correlation with respect to test data especially at high shaft speeds [12]. Trippett
and Li [13] presented a thermal model indicating that thermal effects significantly
affect the performance of high-speed floating-ring bearings and cannot be
neglected.

The models used to characterize the thermal state of the TC were continuously
refined in later publications. San Andrés et al. [10, 11] developed a model
incorporating a lumped-parameter thermal energy balance for the estimation of the
lubricant viscosity and change of operating clearances in the fluid films. The model

422
predicts the bearing performance based on an effective lubricant viscosity and
assumes constant film and ring temperature throughout the flow domains. The
comparison of predicted ring speed ratios with measured values from a cold gas
test rig showed a fair correlation but underestimated the decrease of the ring speed
ratio Ω for high shaft speeds. The authors pointed out that in an actual TC driven
by hot gas, the shaft is a source of thermal energy which was assumed to be
adiabatic due to missing reliable test data.

The most recent thermohydrodynamic model for (semi) floating ring bearings
presented by San Andres et al. [3] models the thermal energy transport in a (semi)
floating ring bearing system and predicts pressure and temperature distributions in
both fluid films, the radial ring temperatures and other performance parameters
like operating film clearances, drag power losses, and heat flow distributions. The
model is used to predict the performance of a typical semi floating ring bearing for
a TC application with shaft speeds of up to 240 krpm, revealing a large radial
temperature gradient across the floating ring. The thermal model presented does
not include non-uniform temperatures in the shaft and the TC casing but it assumes
uniform casing and shaft surface temperatures over the whole operating range. The
temperature distribution in the floating ring is taken as uniform in the
circumferential and axial directions and considers heat flow in radial direction only.
While the model points out the most relevant thermal effects on FRBs and gives a
good overview of the heat flow paths, the predictions are made for a semi floating
ring bearing system only. Also, no comparison with measurements from a hot gas
test was shown.

For a systematic investigation of thermal boundary conditions and their impact on


FRB performance parameters, a complex thermo-hydrodynamic, steady-state
model is introduced in the present paper. The model predicts pressure, temperature
and viscosity distributions in both fluid films and analyzes heat flow paths according
to thermal boundary conditions. The temperature distribution in the fluid films,
floating ring, shaft and an annular TC casing segment is calculated three-
dimensionally which provides for an in-depth investigation of thermal effects like
varied boundary conditions and their influences on bearing performance
parameters. As a second important aspect, a sophisticated experimental and
numerical approach for the determination of thermal boundary conditions is
described. The predicted performance parameters for varied thermal boundary
conditions and heat flow concepts are later compared with actual measured test
data from a hot gas test. The importance of precise thermal boundary conditions for
accurate bearing performance predictions is shown in detail below.

2 THEORETICAL ANALYSIS

The floating ring bearing system performance parameters are calculated by a


numerical simulation program. The pressure distributions in the outer and inner
fluid film are determined by solving a two-dimensional (2D) extended and
generalized Reynolds equation [14, 15] considering the effective hydrodynamic
angular velocity for each individual film. A mass conserving cavitation model based
on the algorithm by Elrod [16] locates the cavitation boundaries and gives
information about the local oil film fill factor throughout cavitating lubricant films.
Boundary conditions for the solution of the Reynolds equation are ambient pressure
at the bearing discharge planes and supply pressure at the individual oil supply
holes in the outer and inner films. The resulting inner and outer film pressure
distributions are shown in Fig. 2 (left) for reference. Integrating the individual
pressure field over the corresponding bearing surface returns the fluid film reaction
forces resulting from the hydrodynamic pressure and hydrostatic oil supply [3]. An
iterative algorithm finds the balance between the external radial static load on the

423
FRB and the fluid film reaction forces by iteratively modifying the displacements of
journal and floating ring, respectively.

The temperature distribution in the fluid films is derived by solving a full 3D energy
equation for a steady, incompressible and laminar oil film, explained in detail in
reference [14]. Unlike most previous thermal FRB models [3, 10], the present
computational program also solves 3D energy transport equations for an annular
segment of the casing and the ring (model scope shown in Fig. 1 (right) for
reference), taking the individual thermal material properties into account. Semi
floating rings in particular show a significant temperature gradient in
circumferential and axial directions due to non-uniform temperature distributions in
the adjacent fluid films. In full-floating rings, the influence of the oil feed holes to
the inner film will produce local spots of lower temperature along the ring
circumference as well. This makes the calculation of a 3D temperature distribution
in both semi-floating and full-floating rings necessary for precisely modeling the
heat transfer through the ring. The TC casing will usually develop a temperature
variation across the circumference as the wide oil supply inlet acts as a local heat
sink while conductive heat transfer through the TC casing from the turbine heats up
the surroundings of the bearing unevenly as is confirmed by test data below. As the
shaft temperature does not vary in the circumferential direction due to the high
rotational speed, a two-dimensional energy transport equation is sufficient for
proper modeling of the heat flow in the shaft segment [14]. At the interfaces
between the single sections of the bearing, a continuous heat flux is postulated.
Continuity of the temperature distribution is preserved at the interface by setting
the local lubricant temperature equal to the temperature of the solid surface. Due
to high rotational speeds, the temperature is set to be circumferentially uniform at
the outer ring and shaft surfaces. The local film fill factors are used to scale the
heat transfer at the interfaces to account for the lowered heat transfer capability in
partly filled film regions [16]. The outlined approach presents an efficient and
precise method to calculate the temperature distribution in the FRB, exactly
modeling heat transfer between lubricant and solid bearing surfaces and neither
requiring calculation of local heat convection coefficients nor iterative balancing of
heat flows between bearing sections. Fig. 2 (right) depicts the resulting
temperature distributions at the inner and outer film centers for a symmetrical FRB
(cut at the axial mirror plane). In those regions of the film where fresh lubricant
enters, the temperature is significantly lowered. The maximum film temperature
develops towards sides of the bearing in both films. The 3D temperature
distributions in the lubricant films and the associated Reynolds equations are
coupled through temperature-dependent, local effective lubricant viscosities that
vary in all three dimensions. Since the Reynolds equation is solved two-
dimensionally only, the radial variance of the effective lubricant viscosity is therein
considered using correction factors [14].

Operating film clearances in the inner and outer film contact change during
operation due to thermal expansion of the individual bearing solids (shaft, ring and
TC casing). Estimation of the effective clearances is performed by considering this
thermal expansion of shaft, ring and casing which depends on the specific material
characteristics, the solid surface temperature and the clearances at room
temperature. The clearances are calculated using a simple model also shown in [3]
and are updated on every completed solution of the temperature distribution in the
bearing system.

424
Figure 2 - Pressure distribution (left) and temperature fields at film center
(right, only half bearing shown) for both lubricant films (operating
conditions: Tsup = 90 °C, psup = 3 bar(abs), nshaft = 80 krpm)

Determination of the floating ring rotational speed is based on a balance of the


shear-driven torques τD,o and τD,i on the inner and outer ring sliding surfaces [10].
The drag torques are a function of the local film parameters pressure (Po, Pi), mean
lubricant viscosity (ηo, ηi), film thickness (ho, hi) and fill factor (Fo, Fi). The local fill
factor equals 1 for fully filled film regions and has a value between 0 and 1 in areas
where the film cavitates, scaling the local shear to take the lowered friction in partly
filled film regions into account [16]. Simplified equations for drag torque calculation
for both film flow regions are given in (1) and (2).

 o ho Po 
 D,o  Ro   F
Ao
o
ho
R Ro  Fo   dA
2   o
(1)

 i hi Pi 
 D,i  Ri   Fi S  R  Ri  Fi   dAi (2)
Ai  hi 2  

The residual of the drag torque balance is reduced with an iterative Newton-
Raphson algorithm by adjusting the floating ring rotational speed.

The fluid films are hydraulically interfaced through the radial lubricant feed holes in
the floating ring circumference. With full floating rings, a pressure drop across the
feed holes ∆pc, proportional to the rotational speed of the ring results from
centrifugal forces [17]:

 oil
pc 
8

 R2  Do2  Di2  (3)

As the half-moon grooved supply plenum spans the TC casing surface only partially,
not all of the feed holes will face the outer supply groove during ring rotation which
leads to a pressure deviation between the feed holes. The movement of the holes
presents a disturbance of the hydrodynamic pressure in the outer lubricant film as
the oil flow into the inner film leads to local collapse of the outer film pressure. A
hydraulic film interfacing model for FRBs is used to transform the transient
processes involved into a steady-state model. In the simulation program, the
circumferential area where the radial feed holes traverse is set to a uniform
pressure pl,o which is taken as a boundary condition in the outer films’ Reynolds
equation. The field of constant pressure creates a drainage area allowing a
volumetric oil flow Qo,drain to leave the outer film. The feed pressure in the inner film
pl,i

425
pl ,i  pl ,o  pc (4)

is adjusted by an iterative algorithm until the amount of oil leaving the outer film
region through the specific drainage area Qo,drain equals the total lubricant flow rate
leaving the inner film region Qi,drain. The fluid entering the holes on the ring’s outer
circumference is preheated due to heat flows into the outer film (from TC casing
surface and through the floating ring) and shear-driven power loss. This is taken
into account by the thermal model by calculating the mean temperature of the
lubricant leaving the outer film through the drainage area and feeding the inner film
with the same temperature.

The numerical procedure starts with estimated values for temperature, pressure
and viscosity distribution and an assumed floating ring speed. A conservative finite
difference scheme (FVM) is then used for solution of Reynolds and energy
differential equations [14]. The hybrid scheme is utilized to stabilize the conjugate
heat problem of the energy equation [18]. The energy equations of the films and
the heat conduction equations for the solid bearing sections are solved in a single
linear system of equations. In the first step, the equilibrium between the outer
static load and the film reaction forces is calculated. The thermal state of the FRB is
solved with subsequent adjustment of the operating clearances. In the outer loop of
the computational program the floating ring speed is altered which requires the
beforehand listed steps to be repeated until the thermal state and the drag torques
on the floating ring surfaces are balanced.

3 EXPERIMENT

For model validation, experimental investigations under varied lubricant supply


conditions are conducted in a turbocharger test facility of the Institute of
Turbomachinery and Fluid Dynamic of Leibniz University. A passenger car
turbocharger with full-floating ring bearing is used (Fig. 3). The lubricant is fed to
the outer film through a half-moon shaped groove spanning one third of the casing
surface circumference. From the outer film, the lubricant is supplied to the inner
film through six equally spaced radial feed holes in the full floating ring. Table 1
summarizes the relevant rotor and oil supply information.

Table 1 - Rotor and bearing system data

Rotor length l = 100 mm


Rotor mass m = 100 g
Lubricant dynamic viscosity at 40 °C η = 71.98 mPas
Lubricant supply pressure psup = 3 bar(abs)
Lubricant supply temperature Tsup = 70, 90, 110°C

As described in the preceding section, thermal boundary conditions have a


substantial impact on bearing performance. To obtain realistic experimental data as
model input and for validation purpose, the turbocharger is instrumented with
about forty temperature and thirty pressure probes. The set-up of the hot-gas
turbocharger test rig and the instrumented turbocharger is shown in Figure 3. For
ring speed measurement, an eddy current turbocharger speed measurement
system is used for both FRBs. The sensors are instrumented in the bearing casing
tangentially to the end of the individual oil supply groove detecting the passage of
the radial feed holes in the floating rings. Additionally, the compressor ring is
instrumented with a high-frequency “Kulite” pressure sensor detecting the time-
resolved pressure in the oil supply groove.

426
Figure 3 - Set-up of hot-gas turbocharger test rig (left), instrumented
turbocharger (right)

The bearing brackets are instrumented with four thermocouples at the turbine side
and two at the compressor side at three axial positions. Figure 4 gives a schematic
sketch of the floating ring instrumentation. In addition, three temperature probes
have been placed in the upper section of the bearing housing to detect the global
temperature gradient from the turbine to the compressor side.

Sensor

Figure 4 - Ring speed measurement (left), schematic of the floating


ring temperature instrumentation (right)

The experimental investigations are conducted with the lubricant type SAE 5W-40.
The test matrix includes three lubricant inlet temperatures and the shaft speeds
tested ranging from 40.000 rpm to 200.000 rpm in steps of 10.000 rpm. The
turbine inlet temperature is constant at 600°C for all test runs. The compressor
operates at the optimum efficiency of each speed line shown as red line (Fig. 5 (a)).

Figure 5 (b) displays the measured operating points and floating-ring speed ratios
Ω (rotational ring speed divided by shaft speed) of the compressor and the turbine
ring for the three different oil supply temperatures at constant oil inlet pressure of
3 bar(abs). The maximum speed ratios are measured at the highest lubricant feed
temperature decreasing with higher shaft speed. This qualitative behavior was also
shown by several authors for measurements with a cold gas driven turbine and an
oil supply temperature of 37.7°C [19]. In the present case, the outer temperature
boundary conditions are nearly constant at low rotational rotor speeds mainly
affected by the supply conditions. The shaft temperature affecting the inner oil film
temperature is a function of viscous dissipation leading to a stronger decrease of
ring speed at higher rotor speeds. Measurements with hot gas driving the turbine
show a difference in speed for the compressor and the turbine radial bearings at TC
part-load condition due to heat flows. Figure 5 (c) shows the compressor outlet
temperature and the bearing housing temperatures at three different positions. The

427
turbine side shows a nearly linear characteristic whereas the compressor side
follows the compressor outlet temperature. The corresponding bracket
temperatures are shown in Figure 5 (d).

(a) (b)

(c) (d)

Figure 5 - Measurement results of operating points (a), ring speed


ratios (b), bearing housing temperatures (c), averaged
bearing bracket temperatures (d)

At part-load and depending on the oil supply temperature, the inner casing surface
acts as a heat source or sink and the heat flow from the casing affects the mean
temperature of the outer film, resulting in a change of the effective viscosity and
ring speed ratio. In addition, the shaft temperature rises to higher compressor
pressure ratios and dissipation in the inner fluid film which leads to a further
increase of the shaft temperature. For high rotational speed and high dissipation in
the bearing, the temperatures and thus the rings speed ratios converge for both
rings.

Figure 6 (a) shows a waterfall plot and the orbit for the shaft displacement for rotor
speeds from 40 to 200 krpm. The test data shows three significant subsynchronous
motions up to shaft speeds of 180.000 krpm related to the subsynchronous 1-3.
The corresponding dynamic pressure in the compressor ring oil feed sickle is shown
in Figure 6 (b). Figure 6 (c) depicts the orbit measurement for different rotor
speeds. In the rotor speed range from of 500 to 1000 Hz, the rotor vibrates in the
gyroscopic conical forward mode (subsynchronous 1). At 1000 Hz, the system
jumps into the gyroscopic translational forward mode (subsynchronous 2). For a
rotor speed of about 1800 Hz, the inner fluid film becomes stable and the oil
whirl/whip in the stable outer fluid film becoming unstable excites the gyroscopic
conical forward mode (Subsynchronous 3). Above rotor speeds of 2800 Hz, the

428
subsynchronous 3 disappears and the rotor vibrates around a stable equilibrium
position [20].

(a) (c)
0.01
nrotor
0.008
amplitude in mm

nFR,T

0.006 nFR,C

0.004
0.5* sync nFR
1*sync nFR
Sub 2 1*sync nrotor
0.002

Sub 3
0
200 Sub 1
ro t
o rs 150
1500
pe
ed 1000
in 100
kr p 500 in Hz
m ency
50
0 frequ

(b)

0.1
nrotor
0.08
nFR,T
amplitude in bar

0.06 nFR,C 1*sync noil feed holes

0.04 2*sync noil feed holes

sync
0.02 4*sync noil feed holes

0
200
rot
o rs 150 10000
pe 8000
ed 6000
in 100
kr p in Hz
4000
m 2000 ency
frequ
50
0

Figure 6 - Waterfall plot of shaft displacement (a), waterfall plot of


dynamic pressure (b), orbit measurements for different rotor speeds (c)

These results matches quite well the rotordynamic behavior presented in previously
published literature[20, 21, 22]. The subsynchronous 3 corresponding to ≈50% of
the ring speed denoting instability of the outer FRB oil film. For the high amplitudes
in the shaft speed range from 100 to 190 krpm, subsynchronous 3 increases

429
leading to higher eccentricities of the ring and thus to a decreasing ring speed [23,
20]. The same characteristic of decreasing compressor ring speed is shown in [24].
A deeper rotordynamic analysis based on the measurements presented in this
paper is given by Vetter [23].

4 COMPUTATIONAL PREDICTIONS

A heat flow analysis for two different thermal boundary conditions of the shaft is
presented below. The predicted results for an adiabatic shaft segment (no heat flow
in or out of the shaft) are similar to those results expected for a turbocharger
driven by cold gas. The results calculated with a uniform temperature prescribed at
the cut faces of the shaft give a fair approximation for the operating characteristic
parameters of a turbocharger under realistic operating conditions driven with hot
exhaust gas. Due to the fact that a shaft temperature measurement is not practical,
the shaft end temperatures are assumed to be 200°C. For the actual predictions
this temperature is used as a boundary condition on the axial cut faces of the shaft
segment considered by the model. In both cases, the casing temperatures
measured on the hot gas test rig are used as outer thermal boundary conditions for
the casing segment to gain a better comparability. For simplicity, only the results of
the turbine-side FRB are shown. The oil supply temperature is 90°C and the
lubricant is fed to the outer film with an inlet pressure of 3 bar(abs).

The most relevant predicted temperatures are shown in Figure 7 which differ
significantly for the two cases. As for the adiabatic shaft, the mean temperatures of
the bearing components surfaces and of the fluid films are close to the system
supply temperature at low shaft speeds. A variation of the predicted mean shaft
temperature surface across the operating range is found even for the model with
constant temperature at the shaft axial cut faces. The shaft surface is actively
cooled by the inner film lubricant flow resulting in an inhomogeneous temperature
distribution in the shaft segment.

Figure 7 - Predicted mean temperatures of lubricant films and solid


surfaces

In both plots of Figure 7, the average film and ring surface temperatures stays at
levels slightly above supply temperature even at high shaft speeds. The
characteristic of the inner ring surface temperature appears linked to the inner film
temperature. For the adiabatic shaft model (left), the temperature gradient
between the inner and outer ring surfaces increases significantly with higher shaft
speeds. For the externally heated shaft (right) this characteristic is found less
prominent. In both cases, the inner fluid film exhibits a high temperature gradient
which exists in radial direction at high shaft speeds.

430
Figure 8 depicts the dimensionless heat flows within the FRB plotted over the
operating range, normalized with the maximum heat flow from the shaft. In case of
an adiabatic shaft (top row), the only heat source in the inner film is dissipation,
i.e. the shear-driven drag power loss. As the lubricant flow rate through the inner
film is very small compared to the outer film flow rate, there is only a limited
amount of heat that can be carried away convectively. The inner film mean
temperature starts at the outer film’s supply temperature at low shaft speeds but
heats up with increasing speed. The outer film, by contrast stays near the supply
temperature even at high shaft and ring speeds due to the large amount of oil flow.
This leads to an increasing temperature gradient over the ring in the radial
direction. Due to the excellent thermal conductivity of the floating ring material
(brass) and the high radial temperature gradient, the larger portion of heat is
conducted through the floating ring into the outer film. Here the heat flow mixes
with the heat dissipated due to surface shear and the pressure driven flow
generated by the ring rotation. As the measured casing temperature lies above the
lubricant inlet temperature throughout the whole operating range and slightly
increases towards the maximum shaft speed, a small heat flow from the casing into
the outer film is also found. In the outer film, the total amount of heat is carried
away convectively with the oil flow.

Figure 8 - Top row: Dimensionless heat flow characteristics with adiabatic


shaft, bottom row: with shaft cut faces at a prescribed temperature of
200°C

For the non-adiabatic approach (bottom row) especially for the inner film, a
significantly different heat flow characteristic is found. Prescribing a uniform
temperature at the shaft outer faces, a second heat source is introduced. Even at
low shaft speeds, the inner film mean temperature reaches high levels above the

431
outer supply temperature. The resulting lower mean effective viscosity leads to a
decreased drag torque on the floating ring inner surface which consequently lowers
the ring speed ratio compared to the adiabatic shaft case.

At low shaft speeds, the convective heat flow (lubricant carries heat away) in the
inner film region is dominant over the conductive heat flow through the floating
ring. The fluid, entering the inner film through the feed holes is – although it has
been preheated in the outer film – considerably colder than the inner film fluid and
thus carries heat out of the inner film. As the ring speed increases, the outer film
mean temperature rises and the pressure drop in the ring feed holes grows
resulting in a reduced flow that enters the inner film trough the feed holes at higher
temperature. While the heat is carried away by lubricant convectively, the radial
temperature gradient through the ring increases and leads to an enhanced
conductive heat transport. In the outer film region, the heat flows similar to the
adiabatic shaft case, except for the heat flow entering through the ring outer
surface which is at a high level even for low shaft speeds already. Figure 9 depicts
the influence of the shaft thermal boundary conditions on the mean effective
lubricant viscosity in the inner and outer film regions shown normalized with the
lubricant viscosity at the oil inlet temperature (90°C). For the outer film, a similar
characteristic is found in both cases. The shaft temperature has only limited
influence on the outer film properties due to the high local lubricant flow. A
different characteristic of the mean lubricant viscosity is evident for the inner film.
For the adiabatic case, the mean viscosity starts at a high level close to the outer
film viscosity but drops rapidly with increasing shaft speed. With the shaft acting as
a heat source, the inner film viscosity already starts at a low level for small shaft
speeds and reduces further with increasing shaft speeds but at a significantly lower
rate compared to the outer film’s viscosity characteristic. As the ring speed is
primarily determined by the ratio of the mean viscosity between the inner and
outer film, the characteristics of the predicted ring speed ratio for both thermal
boundary conditions differ significantly. Due to the mean viscosity gradient between
the inner and outer films developing with increasing shaft speed, the adiabatic shaft
model leads to a decreasing ring speed ratio while for the heated shaft, the ring
speed ratio shows only a slight decline over the operating range.

Figure 9 - Dimensionless mean effective lubricant viscosities (left) and ring


speed ratio for varied thermal boundary conditions (right)

The predictions show a significant influence of the shaft thermal boundary


conditions on relevant bearing parameters like temperature and lubricant viscosity
distribution, mechanical power dissipation, and ring speed ratio. The predicted heat
flows, especially for the inner film region, differ considerably.

432
The casing thermal boundary conditions influence the heat flow and the bearing
operating conditions as well. If the inner casing surface acts as a heat source (as it
does for the TC studied), the heat flow from the casing increases the mean
temperature of the outer film resulting in a reduced effective viscosity and a (slight)
decrease in ring speed ratio. However, since the oil flow in the outer film is
significantly larger than in the inner film, for the studied TC under the stated
operating conditions, the extra heat from the casing surface is carried away
efficiently and thus only leads to minor variation of the bearing parameters. These
results may vary for different operating conditions like modified oil feed pressures
and temperatures or exhaust gas entering the turbine with higher temperature or
active cooling of the TC casing.

5 COMPARISON WITH HOT GAS TEST DATA

For model verification, predicted ring speed ratios are compared with measurement
results from the hot gas test rig (Fig. 10, left). The measured TC casing
temperatures are used as thermal boundary conditions for the casing model. The
shaft temperature is again set to 200°C for all calculations. For brevity, only results
from the turbine-side FRB are shown.

For oil supply temperature set to 70°C and 90°C the predicted ring speed ratios
show an excellent correlation with test data at medium to high shaft speeds. At
shaft speeds below 60 krpm, the prescribed shaft temperature leads to an
overestimation of the viscosity reduction in the inner film because the heat flow
from the shaft results in too low a prediction of the ring speed ratio. This indicates
that the shaft temperature is lower than 200°C in reality. Predicted ring speed
ratios for an oil supply temperature of 110°C are higher than measured ring speed
ratios throughout the whole operation range, even though the trend is matched
well. This leads to the conclusion that shaft temperatures during the operation are
higher than 200°C such that a reduced viscosity and thus drag torque is present in
the inner film.

Figure 10 - Predicted ring speed ratios compared to test data for varied
lubricant supply temperature (left), varied shaft thermal boundary
conditions compared to test data for supply temperature at 70°C (right)

To get an indication of real shaft temperatures during test rig operation, sensitivity
analyses were performed for a lubricant supply temperature of 70°, three discrete
shaft speeds and shaft temperatures ranging from 100°C to 250°C in steps of
25°C. The results are shown in Figure 10. For comparison, the results predicted for
an adiabatic shaft segment are shown as well (gray), with the adiabatic analysis
misjudging the real ring-speed ratios significantly. The expected actual shaft

433
temperature can be obtained by searching for the intersections of the measured
ring speed ratio curve with the ring speed ratios calculated for different shaft
temperatures at the particular shaft speed. Specifically, for the lowest shaft speed
(40 krpm), a significantly lower shaft temperature of 150°C than assumed is found
while at high shaft speeds the assumed 200°C seem to fit the experimental results
well.

6 CONCLUSIONS

In the present paper, a complex thermo-hydrodynamic computational model for the


prediction of relevant FRB operating parameters was presented. For making
realistic assumptions concerning the thermal boundary conditions for the FRB and
for model validation purposes, an extensive experimental study on a hot-gas
turbocharger test bench was conducted.

The computational model was used to study the thermal boundary condition
influences on bearing operating parameters and heat flow characteristics.
Comparing predicted ring-speed ratios assuming an adiabatic shaft with test data
from a hot gas test bench showed a significant discrepancy. In contrast, predictions
produced with a fixed shaft temperature showed a significant improvement in
agreement with the test data over a wide operating range. A sensitivity analysis of
the shaft temperature revealed the serious impact of the shaft thermal boundary
condition on the bearing operating parameters.

The present work shows that the knowledge of the thermal boundary conditions is
mandatory for a precise prediction of the bearing performance parameters.
Neglecting non-adiabatic influences leads to an inaccurate prediction of TC
performance in rotor dynamic and combustion engine process simulations,
particularly in TC part load conditions. For further improvement of the prediction
quality, the bearing simulation model should be integrated into a TC performance
prediction model including non-adiabatic approaches to determine thermal
boundary conditions for the numerical bearing simulation. Furthermore, the thermal
influence of the thrust bearing as well as water cooled bearing housing should be
considered.

ACKNOWLEDGEMENTS

The work was supported by the Forschungsvereinigung Verbrennungs-


kraftmaschinen e. V. (FVV) whose funding is gratefully acknowledged. The authors
would also like to thank Borg Warner Turbosystems GmbH for substantial material
support during the study. In addition the authors would like to thank the technician,
Mr. Bielenberg, for his crucial contribution to instrumentation and experiments.

REFERENCE LIST

[1] Brouwer, M. D., Sadeghi, F., Lancaster, C., Archer, J., Donaldson, J., (2013),
“Whirl and Friction Characteristics of High Speed Floating Ring and Ball
Bearing Turbochargers”, ASME J. of Tribology, 135(10), 041102
[2] Shi, F., Deng, D., (2011), “An Analysis for Floating Bearings in a
Turbocharger”, SAE Technical Paper 2011-01-0375
[3] San Andrés, L., Barbarie, V. C., Bhattacharya, A., Gjika, K., (2012), “On the
Effect of Thermal Energy Transport to the Performance of (Semi) Floating
Ring Bearing Systems for Automotive Turbochargers”, J. of Engineering for
Gas Turbines and Power, 134(10), 102507

434
[4] Baines, N., Wygant, K. D., Dris, A., (2012), “The Analysis of Heat Transfer in
Automotive Turbochargers”, J. of Engineering for Gas Turbines and Power,
132(4), 042301
[5] Malobabic, M. and Rautenberg, M. 1987, “Adiabatic and Non-Adiabatic
Efficiencies of small turbochargers”, International Gas Turbine Congress,
Tokyo, Japan, Oct. 26-31, pp. 57-64, Paper No. 87-Tokyo-IGTC-105
[6] Shaaban, S., and J. Seume. “Impact of Turbocharger Non-Adiabatic Operation
on Engine Volumetric Efficiency and Turbo Lag.” International Journal of
Rotating Machinery 2012 (2012)
[7] Bohn, D., Heuer, T., and Kusterer, K., 2003, “Conjugate Flow and Heat
Transfer Investigation of a Turbocharger—Part I: Numerical Results,” ASME
Paper No. GT2003-38445
[8] Bohn, D., Heuer, T., and Kusterer, K., 2003, “Conjugate Flow and Heat
Transfer Investigation of a Turbocharger—Part II: Experimental Results,”
ASME Paper No. GT2003-38449
[9] Heuer, T., Engels, B., and Wollscheid, P., 2005, “Thermomechanical Analysisof
a Turbocharger Based on Conjugate Heat Transfer,” ASME Paper No.GT2005-
68059
[10] San Andrés, L., and Kerth, J., 2004, “Thermal Effects on the Performance of
Floating Ring Bearings for Turbochargers,” Proc. Inst. Mech. Eng., Part J: J.
Eng. Tribol. 218, pp. 1–14
[11] San Andrés, L., Rivadeneira, J. C., Chinta, M., Gjika, K., LaRue, G., (2007),
“Nonlinear Rotordynamics of Automotive Turbochargers: Predictions and
Comparisons to Test Data”, J. of Engineering for Gas Turbines and Power,
129(4), pp. 488-493
[12] San Andrés, L., Rivadeneira, J. C., Gjika, K., Groves, C., LaRue, G., (2007),
“Rotordynamics of Small Turbochargers Supported on Floating Ring Bearings –
Highlights in Bearing Analysis and Experimental Validation”, ASME J. of
Tribology, 129(4), pp. 391-397
[13] Trippett, R., and Li, D., (1983), “High-Speed Floating-Ring Bearing Test and
Analysis”, ASLE Trans. 27(1), pp. 73–81
[14] Hagemann, T., Kukla, S., Schwarze, H., (2013), “Measurement and prediction
of the static operating conditions of a large turbine tilting-pad bearing under
high circumferential speeds and heavy loads”, Proceedings of the ASME Turbo
Expo, ASME Paper No. GT2013-95004, San Antonio, TX., USA
[15] Dowson, D., (1962), “A Generalized Reynolds Equation for Fluid Film
Lubrication”, Int. J. Mechanical Sciences, 4, pp. 159-170
[16] Mittwollen, N., (1990), “Betriebsverhalten von Radialgleitlagern bei hohen
Umfangsgeschwindigkeiten und hohen thermischen Belastungen –
Theoretische Untersuchungen”, VDI series 1 no. 187, VDI-Verlag, Dusseldorf,
Germany
[17] Meyer, A., (1987), “Äußere Lagerdämpfung für sehr hochtourige,
gleitgelagerte Rotoren”, PHD Thesis, TH Karlsruhe, Germany
[18] Patankar, S. V. (1980), “Numerical Heat Transfer and Fluid Flow”, McGraw-
Hill, New York
[19] San Andrés, L., Rivadeneira, J. C., Gjika, K., Chinta, M., and LaRue, G.,
(2007), “A Virtual Tool for Prediction of Turbocharger Nonlinear Dynamic
Response: Validation Against Test Data”, J. of Engineering for Gas Turbines
and Power, 129(4), pp. 1035-1046
[20] Schweizer, B.: “Dynamics and Stability of Automotive Turbochargers”,
Archive of Applied Mechanics, Volume 80, Issue 9, pp. 1017-43, 2010, Doi:
10.1007/s00419-009-0331-0

435
[21] Schweizer, B.; Sievert, M.: “Nonlinear Oscillations of Automotive
Turbocharger Turbines”, Journal of Sound and Vibration, Vol. 321, pp. 955-
975, 2009, Doi: 10.1016/j.jsv.2008.10.013
[22] Tomm, U.; Boyaci, A.; Proppe, C.; Seemann, W.; Busch, M.; Esmaeil, L.;
Schweizer, B.: “Rotor Dynamic Analysis of a Passenger Car Turbocharger
using Run-Up Simulation and Bifurcation Theory”, IMechE, 9th International
Conference on Turbochargers and Turbocharging, pp. 335-347, London, May
2010
[23] Vetter, D., Hagemann, T., Schwarze, H., (2014), “Predictions for run-up
procedures of automotive turbochargers with floating-ring bearings including
thermal effects and different bearing setups”, Proc. of the 11th Int. Conf. on
Turbochargers and Turbocharging, IMechE, London, GB (to be published)
[24] Chebli, E., Müller, M., Leweux, J. Gorbach, G., (2013), “Entwicklung Eines
Abgasturboladers für die Schweren Nfz-Motoren von Daimler”, MTZ -
Motortechnische Zeitschrift, Volume 74, Issue 2, pp 124-129

436
Stability analysis of pressurized Gas Foil
Bearings for high speed applications
R Hoffmann, T Pronobis, R Liebich
Chair of Engineering Design and Product Reliability, Berlin Institute of Technology,
Germany

ABSTRACT

One of the most efficient cooling methods for Gas foil bearings is a side feed
pressurization. Besides cooling effects, experimental investigations have shown a
reduction of sub synchronous vibrations, which have a major impact on the rotor
dynamic performance. This current paper examines the effects of side feed
pressurization on the dynamic behaviour. A multi physics model is used to calculate
dynamic properties (stiffness and damping). In addition the effect of a pre-swirl
side feed pressurization is considered.

1 INTRODUCTION

Over the past three decades Gas foil bearings (GFBs) have been successfully
introduced in the field of high speed turbo machineries. Low drag friction, high
speed operation and the omission of an oil system are some advantages of
compliant foil bearings (1). In general, GFBs are based on the hydrodynamic
pressure. This pressure is induced by a generated slip stream between the turning
bearing journal and the bearing foil (see Figure 1b). An elastic structure comprises
one or more thin top foils supported by corrugated bumps (see Figure 1a) which
are the main difference to common gas bearings with a rigid bearing housing.
Therefore an optimal film thickness is achieved and higher loadings compared to
rigid gas bearings are possible (2). Due to friction contacts inside the corrugated
structure a structural damping is induced and can be estimated experimentally
(3)(4). A tuning of the compliant structure by staggering of bumps, applying
coatings, using multiple bump layers results in higher load capacities and can
reduce high sub synchronous whirl amplitudes (2)(5)(6).

Nevertheless, recent developments in GFB applications have heightened the need of


an appropriate thermal management for high temperature environments due to
limited temperature durability (7)(8). Hence, special coatings and cooling
techniques have been established to prevent thermal instabilities and thermal
seizure (8)(9). One of the most efficient cooling method is a side feed
pressurization of GFBs (7)(9)(10). A forced cooling flow is streaming underneath
the corrugated bearing structure and the clearance between bearing journal and
top foil (see Figure 2a). Heat is transported by convection and conduction effects.
In addition Kim et al. (10) have shown that the effect of this method reduces the
sub synchronous amplitudes for low loaded bearing conditions (W≈ 5 N). The sub
synchronous whirl vibrations have a major impact on the dynamic behaviour. To
reduce sub synchronous vibration several methods and devices have been
introduced (11)(12)(13)(14).

_______________________________________
© The author(s) and/or their employer(s), 2014
437
However, there has not been a discussion about the impact of side pressurization in
detail and especially the impact of a pre-swirl pressurized GFB on rotor dynamic
stability has not been reported. A pre-swirl has a direct influence on the velocity
field inside the film thickness of a GFB. First, a 2D FE-model of the corrugated
structure is coupled with the non-linear compressible Reynolds equation, which is
discretized by a hybrid finite difference scheme. A perturbation method is applied to
estimate linearized bearing parameters, they are inputs of a linear stability
analyses. It is limited to small orbits around a static equilibrium position. However,
to estimate the onset speed of whirling motions it is an appropriate method, as
shown in (11). Finally, linearized bearing parameters will be analysed in detail.

2 THEORETICAL MODEL

The elastic structure of a simple GFB in a coordinate system ( , and ) is shown in


Figure 1b). Due to the presence of a turning journal with an angular speed and
centre displacements εx and εy a forced slip stream with a film thickness of ℎ( , ) is

a) b)

Figure 1. Schematic GFB structure a) and


GFB with a dynamic pressure field b)

induced and results in a dynamic pressure field ( , ), where and are related to
the axial and circumferential coordinate. This pressure generates a reacting force
fB. The equilibrium condition is reached if the sum of the loading vector =
( , ) and the reacting force vector fB is zero. An integration of the pressure
field along the axial and circumferential direction yields the reacting force vector fB
(Eq. (1)), where ( ) is the pressure underneath the foil and along the axial
direction z. It acts under the attitude angle = ( , ).

( , )− ( ) cos
=− d d (1)
sin
Eq. (2) describes a perfectly aligned journal, expansion effects due to temperature
gradients and centrifugal forces are neglected. The film thickness is composed of
two parts: A rigid term ℎ ( ) and a compliant term ℎ ( , ). The first part includes
the nominal bearing clearance and the journal centre displacement. The compliant
term ℎ ( , ) describes the deformation of the corrugated structure due to the
pressure field and is calculated by a structural model as shown below.

438
ℎ( , ) = + cos θ + sin θ + ℎ ( , θ) (2)
( )

The pressure field ( , ) is calculated by solving the Reynolds equation for a


compressible, isothermal and isoviscous fluid, Eq. (3). It links the pressure field
with the film thickness under the presence of journal rotation speed ( ), where R
is the bearing radius and μ the absolute viscosity. Note the rotation speed is a
function of due to side feed pressurization.

( )
+ = ( ℎ) + 12 ( ℎ) (3)

2.1 Side feed pressurization

a) b)

Figure 2. Side feed pressurized GFB a) and swirl generator b)

Figure 2a) shows a schematic side feed pressurized GFB. A pressure drop
underneath the top foil along the bearing length is generated due to the pressure
difference between the feed pressure and ambient pressure , Eq. (4).

( )= 1− + (4)

Following boundary conditions are applied for solving the Reynolds equation:
( , 0) = and ( , ) = . The pressure of the leading and trailing edge of top
foil at = 0° is given by ( = 0, ) = ( = 2 , ) = ( ). Due to a side feed
pressurization the circumferential speed becomes a function of and can be
described by Eq. (5). It has been used by Black et al. (15) to describe pre-swirled
gas seals. However, (10) have applied this equation to estimate side feed
pressurization of GFBs, while neglecting swirling ( = 0).

( )= 1− + (5)

Where


= , = (6)

(Temperature and ideal gas constant )

439
Considering a pre-swirl, the swirl number is applied. It is the ratio between the
nominal circumferential velocity = and the circumferential part of the
absolute velocity of a pre-swirl guide vane, as shown in Figure 2b). These guide
vanes are placed at one side of the gas foil bearing.

,2
= (7)

Due to neglecting of boundary layer effects in the swirl generator an isentropic flow
can be assumed and yields to Eq. (8), where =1.4. If = = ., = .
and = = . are given, the swirl number will depend on the swirl angle ±
and the pressure ratio ∏ = .

−1
( ±) 2−1 1 1− ∏
= ,2
≡ (8)

2.2 Structural model

a) b)
Figure 3. Schematic structure a) and model including sagging b)

In Figure 3a) a schematic compliant structure of a GFB is shown. The complex


structure is reduced towards a 2D Timoshenko plate for the top foil and a linked
spring model, based on Iordanoffs (16) bump model, Figure 3b). To avoid sagging
effects the Young-modulus of the plate is increased by a factor of 4, ref. (14). The
overall structure can be explained by a finite element model, Eq. (9). The compliant
part ℎ is a subset of the displacement vector .

ℎ ⋂ = (9)

Where the pressure force vector is based on the pressure difference of ( , ) −


( ).

2.3 Dynamic performance


To estimate the dynamic bearing parameters for a given load w and steady speed
condition (Ω =const.) a perturbation method is applied. This method was introduced
by Lund (17) first. Assuming a harmonic perturbation with the perturbation
frequency and small amplitudes (∆ , ≪ ) acting around the equilibrium state
( = ( , , , ) ). In addition, the pressure field and the film thickness are affected
by the perturbation, as given by Eq. (11) and (12), where i=x,y.

= +∆ (10)

= +∆ (11)

ℎ = ℎ + ∆ (ℎ + ℎ , ) (12)

440
Where ℎ = cos and ℎ = sin . The term ℎ , is the perturbed compliant film
thickness. It is based on the perturbed displacements, analogously to Eq. (9). The
global stiffness matrix is changed to a complex stiffness matrix; to introduce
viscoelastic damping effects of the bump structure, while damping of the top foil is
neglected, see Eq. (13). The structural viscoelastic damping is related to the
dissipated energy caused by friction contacts inside the compliant structure and can
be described by the structural loss factor γ.

= + (1 + ) (13)

Substituting the perturbed approaches (Eq. (10)-(12)) into the Reynolds equation
(Eq. (3)) while neglecting terms of higher order leads to a zero and a first order
Reynolds equation. A hybrid finite difference scheme is applied to calculate the
pressure field ( , ) and ( , ). For a detailed description see ref. (17)(18).
Using the perturbed pressure field to calculate the linearized stiffness and damping
matrix.

cos cos
+ =− d d (14)
sin sin

Note the described model has been validated by experimental data of Ruscitto et al.
(19) for a non pressurized case, ref. (14). For evaluating the dynamic performance
of a GFB system a simple linear oscillator model is introduced, which has been
widely used to analyse GFBs, ref. (17)(18), where is the journal mass.

0 0
+ + = (15)
0 0

Introducing the impedance ℎ = + and considering a free oscillation yields


to an eigenvalue problem. It is solved by using a harmonic state of solution and
assuming non trivial solutions. The eigenvalue is defined by = ± . Positive
real parts indicate an unstable behaviour of a linear system. For = 0 the neutral
stability is gained. To determine the neutral stability for a steady speed condition
( = .) the excitation frequency ratio is varied, due to the dependence of the
bearing parameter. It enables the calculation of the critical excitation frequency
, and the critical modal stiffness , . Note for neutral stability the modal
damping is zero. Finally, these parameters can be used for introducing the critical
mass parameter to evaluate stability (11).

,
= 2
(16)
,

3 NUMERICAL RESULTS

Table 1 lists the geometrical data of the investigated GFB. Simulations are based on
ambient conditions ( = 293.15 K, = 1 bar) and a nominal clearance c of
31.8 μ is assumed. Different bump stiffness factors are applied, calculated with
the model of Iordanoff (16) ( , = 8.76 ∙ 10 N/ mm and , = 3.87 ∙
10 N/ mm ) and a structural loss factor of γ=0.2 is considered. The mesh grid
includes 130 circumferential and 40 axial nodes. Further, a higher node density
with an axial element length of = /120 is introduced for 5 axial nodes at each

441
edge along circumferential direction to achieve higher numerical stability due to
huge pressure gradients in this area. Three different side feed pressures were
observed (1.0, 1.5 and 3.0 bar), while the excitation frequency ratio ⁄
was varied between 0.1-10.

Table 1: Geometrical data of a GFB (17)

Parameter Variable Value Unit


bearing radius 19.05 mm
bearing length 38.1 mm
bump thickness 0.1016 mm
bump pitch 4.572 mm
bump number 26 -
half bump length 1.778 mm
clearance c 31.8 μm
foil thickness 0.1016 mm
Young modulus 2.07 ∙ 10 N/mm²
Poisson ratio 0.27 -

1.0E+07
30 N
p=1bar
1.0E+06

1.0E+05
critical mass parameter

1.0E+04
15 N unstable
region
1.0E+03

1.0E+02

1.0E+01
5N
1.0E+00
6000 12000 18000 24000 30000 36000 42000
rotor speed rpm

Figure 4. Critical mass parameter vs. rotor speed without side feed
pressurization

For a set of loadings (5, 15 and 30 N) the critical mass parameter versus the rotor
speed is shown in Figure 4. Furthermore, three horizontal dashed lines are plotted.
They are related to the journal mass and correspond to the loading ⁄ ).
Note, in the first analysis no feed pressure is supplied. The area underneath each
curve is called stable region. An increase of mass parameter or rotor speed beyond
this threshold will cause an unstable behaviour. The critical mass parameters with
values of 10 are stable and hold for an adequate logarithmical display. As shown in
the diagram increased loads lead to a higher stability range while higher speed
numbers reduce the range of stability. The onset speed is estimated at 14.800,
16.500 and 22.100 rpm for the loading cases 5, 15 and 30 N. As reported in (14),
higher loading along the vertical direction raises direct stiffness and damping
values . Whereas the stiffness difference of the cross coupled stiffness (
) is slightly reduced for a sub synchronous excitation ( 1). Note

442
has a major impact on the stability: Small values of results in less
coupling effects and can raise the range of stability (20). In addition, higher
excitation frequency ratios and raised rotor speeds affect damping and stiffness,
while damping decreases and stiffness is conversely affected.

In Figure 5 the critical mass parameter versus the rotor speed is shown for the
same loading set. In addition, different levels of side pressure without a pre-swirl
are listed. The same effects as mentioned above due to loading and rotor speed are
plotted. Increasing side feed pressure raises the range of stability slightly, for 5 and
15 N. Only for 30 N and speeds of 6.000-23.000 rpm the GFB shows lower stability
ranges for side feed pressurization.

30 N
=
1.0E+03
critical mass parameter

1.0E+02

15 N

1.0E+01
p=1.0 bar 5N
p=1.5 bar
p=3.0 bar
1.0E+00
10000 15000 20000 25000 30000 35000 40000
rotor speed rpm

Figure 5. Critical mass parameter vs. rotor speed, influence of a side feed
pressure (without swirl = )

Figure 6 shows bearing parameters of three different feed pressure levels 1, 1.5
and 3 bar, whereas the 1 bar values are based on a GFB without side feed
pressurization (ambient condition). All values are displayed versus the excitation
frequency ratio ( = ⁄ ). The GFB is loaded by a vertical load of 30 N. As shown
in the figure, increasing rotor speed, excitation frequency ratio and feed pressure
yields higher direct stiffness. The damping is higher due to feed pressures
compared to the case of p=1 bar. Furthermore, an increase of rotor speed or
excitation frequency ratio significantly drops the direct damping values. Those
effects of pressurization on the direct parameters are more notable for low rotor
speeds. It should be mentioned, that similar results for and were observed.

However, for a sub synchronous excitation range the difference of the cross coupled
stiffness is bigger for a side feed pressurized condition, compared to the case p=1
bar. The difference rises with increasing rotor speed and decreases towards zero for
higher excitation frequency ratios ( 1.). As mentioned before, the difference of
the cross coupled stiffness has a major impact on stability. High values can
decrease the range of stability. Hence, the overall behaviour results in lower
stability in the speed range between 6.000-20.000 rpm for a medium loaded
bearing (30 N), as shown in Figure 5. This effect is not notable for loadings of 5 and
15 N, due to the higher damping and stiffness values for a side feed pressurization,
similar results have been reported by Kim et al. (10). However, the work of Kim et
al. (10) was limited to 4.9 N loading. This work was extended to 15 N and 30 N

443
(medium loaded conditions). The numerical results of the analysed bearing design
indicate no benefit of a side feed pressure on the stability for 30 N bearing loading.

6.E+06
30
30NN

5.E+06
direct stiffness kyy in N/m

4.E+06 higher rotor speed

3.E+06

2.E+06

1.E+06
6.E+03
direct damping cxx in N/m

4.E+03

higher rotor speed

2.E+03

0.E+00
3.E+06
p=1 bar, 10.000 rpm
cross-coupeld stiffness kxy-kyx in N/m

p=2 bar, 10.000 rpm


p=3 bar, 10.000 rpm
2.E+06
p=1 bar, 70.000 rpm
p=2 bar, 70.000 rpm
p=3 bar, 70.000 rpm
1.E+06

0.E+00

higher rotor speed


-1.E+06
0.1 1 10
excitation frequency ratio η

Figure 6. Bearing parameters at different levels of feed pressure


vs. a GFB at ambient conditions (without swirl = ; W=30 N)

444
0

5N
rotor speed
0.2 increased
15N

5N, p=1bar
vertical excentricity ex

0.4
30N 5N, p=1.5bar

loading
5N, p=3.0bar
0.6 15N, p=1.0bar
15N, p=1.5bar
15N, p=3.0bar
0.8 30N, p=1.0bar
30N, p=1.5bar
30N, p=3.0bar
1

nominal clearance
1.2
0 0.1 0.2 0.3 0.4 0.5 0.6
horizontal excentricity ey

Figure 7. Influence of side feed pressure on the static Trajectory


(without swirl = 0)

Figure 7 shows the static trajectories of the static equilibrium state for different
loading and rotor speed conditions versus the different side pressure levels. For
increasing load (acting in vertical direction) the displacements are raised and
dimensionless eccentricities ( = ⁄ , = , ) are bigger than the nominal
clearance, which is possible due to the compliant structure. An increase of rotor
speed leads the centre of mass towards the centre position ( = 0, = 0), while
higher side feed pressure yields to higher eccentricities along the y axis. This effect
is more notable for higher pressures and loadings, and indicates the increased cross
coupling effects, due to feed pressure.

Finally, a pre-swirl is applied for the second part of this investigation. The
simulations are based on the same side feed pressures as mentioned above. A swirl
angle ± = 20° and two different pressure ratios Π = 1.33 and 1.16 are used. The
pressure ratios are related to the pressures ps=1.5 and 3 bar. Figure 8 shows the
relative deviation of the critical mass parameter related to the swirl free state
versus rotor speed. At low speed conditions and 15 N (ps=1.5 bar) the swirl
increased the mass parameter up to 3.57 %. However, at high speeds the pre-swirl
of positive and negative swirling angles decreases the range of stability.

445
15N, p=1.5bar, alpha+
5
15N, p=1.5bar alpha-
4 15N, p=3.0bar, alpha+
3.57 15N, p=3.0bar, alpha-
3 5N, p=1.5bar, alpha+
rel. deviation ∆߰ܿ‫ ݐ݅ݎ‬in % 5N, p=1.5bar alpha-
2 raised stability range 5N, p=3.0bar, alpha+
5N, p=3.0bar, alpha-
1 30N, p=1.5bar, alpha+
30N, p=1.5bar alpha-
0

-1

-2

-3
decreased stability range
-4

-5
4000 24000 44000 64000
rotor speed in rpm

Figure 8. Relative deviation of the critical mass parameter with


swirling ࣘ ് ૙ related to the swirl free case ࣘ ൌ ૙

4 CONCLUSIONS

This paper has presented a model to investigate the dynamic performance of a gas
foil bearing affected by a side feed pressurization. Increasing pressure results in
higher direct stiffness and damping, while cross coupling effects have been
increased for higher loadings and side feed pressures. However, due to higher
amounts of damping the overall stability has been improved for loadings of 5 and
15 N, but loadings of 30 N have shown a stable behaviour only for high speeds
>20.000 rpm.

In addition, the results have shown that pre-swirling has slightly decreased the
range of stability. Only for loadings of 15 N and 1.5 bar side pressure and low rotor
speeds <10.000 rpm the stability has been improved. The current investigation was
limited on linear stability; future research will be based on time transient
simulations to confirm the impact of a side feed pressurization on the rotor dynamic
performance including nonlinear friction contacts. Further, a test rig is under
construction at the Berlin Institute of Technology to estimate linearized bearing
parameters and dynamic performance of different side feed pressure levels to prove
theoretical findings.

5 REFERENCE LIST

(1) DellaCorte, C., and Valco, M. (2000). Load capacity estimation of foil air
journal bearings for oil-free turbomachinery applications. pp. 795–801.
(2) Heshmat, H., (1994). Advancements in the performance of aerodynamic foil
journal bearings: High speed and load capability. Journal of Tribology,
116(2), pp. 287–294.
(3) Ku, C.-P., and Heshmat, H., (1994). Effect of static load on dynamic structural
properties in a flexible supported foil journal bearing. Journal of Vibration and
Acoustics, 116, pp. 257–262.

446
(4) Rubio, D., and San Andrés, L., (2007). Structural stiffness, dry friction
coefficient, and equivalent viscous damping in a bump-type foil gas bearing.
Journal of Engineering for Gas Turbines and Power, 129(2), pp. 494–502.
(5) Heshmat, H., (1991). Analysis of compliant foil bearings with spatially variable
stiffness. In AIAA, SAE, ASME, and ASEE, Joint Propulsion Conference, 27th,
Sacramento, CA.
(6) Heshmat, H., Shapiro, W., and Gray, S., (1982). Development of foil journal
bearings for high load capacity and high speed whirl stability. ASME J. Lubr.
Technol., 104(2), pp. 149–156.
(7) Kyuho, S. and Kim, T. H., (2011). Thermohydrodynamic analysis of bump-
type gas foil bearings using bump thermal contact and inlet flow mixing
models. Tribology International Tribology International 48 (2012) 137–148.
(8) DellaCorte, C., Valco, M. J., Radil, K. C., and Heshmat, H., 1999, Performance
and Durability of High Temperature Foil Air Bearings for Oil-Free
Turbomachinery, Report No. NASA/TM-1999-209187.
(9) Radil K, DellaCorte C, Zeszotek M., (2007). Thermal management techniques
for oil free turbomachinery systems. Tribology Transaction;50:319–27.
(10) Kim, T. H., and San Andrés, L.,(2009). Effect of side feed pressurization on
the dynamic performance of gas foil bearings: A model anchored to test data.
Journal of Engineering for Gas Turbines and Power, 131(1), p. 012501.
(11) Schiffmann, J., and Spakovszky, Z. S., (2013). Foil bearing design guidelines
for improved stability. Journal of Tribology, 135, pp. 011103–1– 011103–11.
(12) Kim, T. H., and San Andrés, L., (2008). Effect of mechanical preloads on the
dynamic performance of gas foil bearings. Proceedings: IJTC2008-71195,
STLE/ ASME International Joint Tribology Conference, Miami, Florida.
(13) Lee, Y. B., Kim, T. H., Kim, C. H., Lee, N. S., and Choi, D. H., (2004).
Dynamic characteristics of a flexible rotor system supported by a viscoelastic
foil bearing (vefb). Tribology International, 37, pp. 679–687.
(14) Hoffmann, R. Pronobis, T. and Liebich, R., (2014). The impact of modified
corrugated bump structures on the rotor dynamic performance of gas foil
bearings. Proceedings of ASME Turbo Expo 2014.
(15) Black, H. F., Allaire, P. E., and Barrett, L. E., (1981). Inlet flow swirl in short
turbulent annular seal dynamics. Proceedings of the Ninth International
Conference in Fluid Sealing, BHRA Fluid Engineering.
(16) Iordanoff, I. (1999). Analysis of an aerodynamic compliant foil thrust bearing:
Method for a rapid design. Journal of Tribology, 121, pp. 816–822.
(17) Lund, J. W., (1968). Calculation of stiffness and damping properties of gas
bearings. Journal of Lubrication Technology, 90(4), pp. 793–803.
(18) Kim, D., ( 2007). Parametric studies on static and dynamic performance of air
foil bearings with different top foil geometries and bump stiffness
distributions. Journal of Tribology, 129(2), pp. 354–364.
(19) Ruscitto, D., Mc Cormick, J., and Gray, S., (1978). Hydrodynamic air
lubricated compliant surface bearing for an automotive gas turbine engine i -
journal bearing performance. Technical Report NASA CR-135368.
(20) Gasch, R., Nordmann, R., and Pfützner, H.,(2006). Rotordynamik. Springer.

447
The Modified Phan-Thien and Tanner
model applied to turbochargers thrust
bearing
B Rémya,b, T Lamquina, B Bou-Saïdb
a
Honeywell Turbo Technologies, COE Shaft & Bearings, France
b
Université de Lyon, CNRS INSA-Lyon, LaMCoS, France

ABSTRACT

In this paper, an unsteady thermohydrodynamic formulation of a turbocharger’s


thrust bearing contact is presented. The Modified Phan-Thien and Tanner model
takes into account complex rheological characteristics of the lubricant, transient
aspects and inertia effects due to the very high rotational speed. An extensive
rheological testing campaign was performed to get data as input for the numerical
model. The results obtained from this model are compared to experimental data
provided by a thrust bearing rig. It shows acceptable correlation between prediction
and test data of axial thrust load versus thrust bearing oil film thickness for two
typical turbocharger frame sizes.

NOMENCLATURE

, , : Cartesian coordinates [ ] : Time [ ]


( , , ) : Velocities [ ⁄ ] : Density [ ⁄ ]
ℎ : Film thickness [ ] : Pressure [ ]
: Stress tensor [ ] : Averaged flow rate [ ⁄ ]
: Inertia terms [ ⁄ ] : Dynamic viscosity [ . ]
: Heat capacity [ ⁄ ⁄ ] : Temperature [ ]
: Oil relaxation time [ ] , , : Models parameters [−]
: Shear rate [ ] : Critical shear rate [ ]
: Thermal conductivity [ ⁄ ⁄ ]

1 INTRODUCTION

With increasing focus on engine downsizing, high performance and low fuel
consumption, turbochargers technology has to take up several design challenges to
reach automotive engines expectations. One of them consists in analyzing the
phenomenon of “turbo lag” that delays the boost response and penalizes the
customer drivability. The understanding of the bearing system load capacity and
friction in steady state becomes not sufficient in extreme operating conditions and a
transient formulation of the problem must be solved accounting for complex
rheological characteristics and very high rotational speed effects.

Engine lubricants are submitted to a wide range of temperatures, pressures and


shear rates. Although a low viscosity saves energy, lubricants need to have
appropriate viscosity to aid cold and start/stop functionality and need to have
sufficient viscosity to provide enough capacity at high loads and temperatures

_______________________________________
© The author(s) and/or their employer(s), 2014
449
where wear might otherwise occur. The ideal lubricant should have a flat viscosity-
temperature relationship. But for mineral oils, the viscosity drops with the
temperature rise. Therefore, viscosity-index (VI) improvers in small percentage
consisting in long-chain polymers are added to the solvent (1). These additives
bring a satisfactory viscosity-temperature relationship but convey a non-Newtonian
behavior to the lubricant.

One of the most widely used rheological models that can represent those two
lubricant properties is the Modified Phan-Thien and Tanner (MPTT) model (2). In
this paper, the thrust bearing contact of a turbocharger is studied. The MPTT model
is combined with two viscosity laws accounting for the shear-thinning effects (Cross
model (3)) and the viscosity dependency on temperature (Walther and McCoull
model (4)). High rotational speeds, up to 300 000 rpm, yield to account for inertia
effects. A specific software has been developed to solve numerically these
equations. Some numerical parameters are based on an extensive testing campaign
on a Couette rheometer. The first time step of the analytical model is compared to
test data collected on a thrust bearing rig in terms of axial load and power loss.

2 ANALYSIS

Throughout this paper, we use thin film theory assumptions. For the considered
thrust bearings, the ratio of the thickness over the radius varies between 10-2 and
10-3. Thus, all second order terms such as (ℎ/ ) are neglected. In addition, the
flow is supposed incompressible and laminar in the region of contact where
Reynolds number varies from 10 to 1000 being in the lower range of the laminar to
turbulent transition. All external forces are negligible and there is no slip of the fluid
at the walls. Note that ( ), ( ) and ( ) respectively are the tangential, radial
and vertical directions of the contact.

2.1 Modified Reynolds equation


As inertia effects are considered, a direct integration of the momentum equations is
impossible. Therefore, we follow the method developed by Tichy and Bou-Saïd (5)
which consists in averaging these equations through the film thickness. Using
Leibniz integral rule, one obtains:

+ + = −ℎ + − [1]

0 = [2]

+ + = −ℎ + − [3]

Where

= = = = =

Note that is the total deviatoric stress tensor. In our study, it is to be expressed
as the sum of a Newtonian stress and a polymeric stress based on the MPTT
model.

The continuity equation is also averaged through the thickness:


+ + = 0 [4]

450
The main assumption of this kind of method is to assume the velocity profiles.
Given the boundary conditions ( = ; = 0 ; = 0 at = 0 and = 0 ; = ; =
0 at = ℎ), we superpose a Couette flow and a Poiseuille flow determined without
inertia terms. Proceeding to similar arrangements to Tichy and Bou-Saïd
developments on the averaged momentum and continuity equations, a Modified
Reynolds equation is obtained:

ℎ + ℎ [5]
ℎ ℎ
= 12 +6
ℎ ℎ
− 2ℎ + + + 2ℎ + +

− ℎ − + +2 + − 12ℎ +
ℎ ℎ
+ 2ℎ − + −

+ℎ − + −

with = ℎ and = ℎ the averaged Poiseuille velocities respectively


in the radial and tangential directions.

The difference with the classical Reynolds equation stands in the second part of the
right side member. The third and fourth terms account for the inertia effects and
the surface acceleration. The fifth term represents the variation of the viscosity in
the contact. Finally, the polymeric contribution to the pressure field stands in the
two last terms. At this state of the development, the polymeric stress tensor needs
to be known to obtain the pressure field. We choose to solve the Modified Phan-
Thien and Tanner model.

2.2 Modified Phan-Thien and Tanner model


The MPTT model considers the fluid as a Newtonian solvent including a homogenous
and continuous network that has a non-Newtonian behavior (2). In the case of
multigrade engine oils, the solvent corresponds to the basic mineral oil whereas the
polymeric network stands for the long polymeric chains of the additives improving
the VI. The global behavior of such a fluid is non-Newtonian. That is practically
illustrated by shear thinning effects. The polymeric chains, or segments, interact
between each other at their junctions. These junctions are not permanent but are
continuously being created and destroyed and thus can result in loose end
segments. The more numerous the loose ends, the smaller the viscosity.

The MPTT model can be expressed as (6):

= + [6]

= ( + )= [7]
1 1
+ . − − . − . − + = [8]
2 2

where tr = exp represents the possibility to break polymeric


chains at very high shear, (−) being the trace function. The sum of the solvent
viscosity and the polymeric viscosity constitutes the fluid viscosity . In Eq.8,
is the polymer relaxation time, and ξ are dimensionless parameters respectively

451
standing for the elongation and the slip between the polymeric network and the
solvent. Several models such as the Upper Convected Maxwell model and the
White-Metzner model can be found when setting and ξ to specific values (7).

To deal with the shear-thinning effect, we introduce Cross equation (3):



= + [9]
1+( / )
where is the dynamic viscosity, and the viscosities respectively under no
and infinite shear rate, the critical shear rate and a parameter.

As for the viscosity dependency on the temperature, the Walther and McCoull
model was chosen (4):

= 10 − [10]
where is the kinematic viscosity under no shear rate, the temperature, , ,
three parameters. Although only the polymeric part of the lubricant should suffer
shear-thinning while the temperature should mainly affect the solvent part, we
apply these two viscosity laws on the global viscosity of the oil. Indeed, it was only
possible to perform rheological measurements, and therefore parameters fitting, on
the final mixture of the lubricant. To apply the Walther and McCoull model, the
temperature is calculated everywhere in the contact using the energy equation.

2.3 Energy equation


The equation for the thermohydrodynamic (THD) model is a 3D thin film energy
equation (8) that describes the heat transfer in the film thickness:

+ + + = + + [11]

In order to avoid instabilities during the


Initial conditions (t=0)
numerical calculation, the Richtmyer
method (9) is introduced. It consists of
calculating the temperature by using
Assume velocity profiles
forward or backward finite differences
depending on the flow direction. Thus, Solve MPTT model
eight cases are implemented according to
the velocities signs combinations. It avoids
Calculation of Modified
calculation instabilities at the outlet of the Reynolds equation
thrust bearing, near by the outer radius. Obtain pressure field

2.4 Numerical resolution Calculation of energy


equation
All equations are discretized using finite
differences. A flow chart of the global
Obtain temperature field

method, presented in Figure 1, consists in Viscosity update


an iterative process between velocities
and pressure field calculations. A Convergence
Convergence? NO
substitution method is used to solve the on pressure?
MPTT model and provide the non-
Newtonian stress tensor. The pressure and Performances
New rotational
calculation
the temperatures in the film thickness are speed and/or film
thickness
determined by using the over-relaxed t = t + dt
(t<tend)
Gauss-Seidel method. The convergence
criteria is reached when the maximum of
the local pressure ratio is lower than 10-4. Figure 1: Flowchart of
We can then calculate the global bearing numerical solution

452
performances and proceed to the next time step calculation. The resolution of this
problem is achieved using Fortran 90 and presented for the first time step, which is
equivalent to the steady state case.

3 EXPERIMENTAL STUDY

3.1 Rheological measurements


Three types of rheological measurements were performed on the Castrol SLX
Longtec Longlife II engine oil. An Anton Paar Physica MCR 301 rheometer with co-
axial cylinders creating a Couette flow has been used for the test.

3.1.1 Viscosity-temperature
The viscosity of the lubricant was measured for temperatures going from -20°C to
150°C. For each intermediate temperature, the fluid was submitted to a shear rate
ramp from 0.03s-1 to 3000s-1. Averaged and repeated measurements were
compared to Walther and McCoull model based on the two data points given by
Castrol and based on three experimental data points (see Fig.2).

These rheological experiments demonstrate that Walther and McCoull model is a


very accurate and robust model to calculate the viscosity as a function of the
temperature, particularly for high temperatures. It is of course even better when its
coefficients are based on three experimental data points. Based on the two data
points from the oil manufacturer and by assuming that = 0.80, we have = 7.80
and = 3.03. By fitting three experimental points using the least squares method,
we obtain = 0.58, = 7.91 and = 3.07 which remain near by the first triplet. In
the numerical programs presented further in this paper, we use the last set of
parameters.

In-house experimental data

Castrol exprimental data

Walther & McCoull model based on three in-house experimental data

Zoom-in between 20°C and 80°C Zoom-in between 100°C and 150°C
0.05 0.008

0.045

0.007
0.04

0.035
Viscosity [Pa.s]

Viscosity [Pa.s]

0.006
0.03

0.025
0.005

0.02

0.015
0.004

0.01

0.005 0.003
40 50 60 70 80 90 100 100 110 120 130 140 150
Temperature [°C] Temperature [°C]

Figure 2: Castrol SLX Longtec Longlife II (0W30) Viscosity – temperature


model correlation (a) Zoom-in between 40°C and 100°C -
(b) Zoom-in between 100°C and 150°C

3.1.2 Shear-thinning effect


Some measurements were performed to obtain the parameters constituting the
Cross equation (Eq.9). Current turbochargers speeds can reach up to 300000 rpm
which correspond to shear-rates of 107s-1. Unfortunately, the highest shear-rates

453
provided by the rheometer we used could not exceed 3000 s-1. For that reason, we
only could observe the beginning of the shear-thinning effect. Thus the parameter
∞ standing for the viscosity at infinitely high shear-rate and which is the critical
shear-rate standing for the transition between the high and the low viscosity plates
have been obtained from literature (10) for this oil submitted to turbocharger
solicitations in term of shear rate magnitude. Using different oil than Castrol SLX
Longtec Longlife II in simulations would be critical and need specific tests to be
conducted too.

3.1.3 Relaxation time


To access the relaxation time of the lubricant, we used the oscillatory mode of the
Anton Paar rheometer to run a frequency sweep and obtain its loss and storage
moduli (11-12). But the oil having a very low viscosity and the machine being
limited to 100Hz, we got data in that range, which is a little bit far from the rotating
frequency of turbochargers. Although those results tend to confirm that the
relaxation time of engine oils is between 10-3 and 10-6 s (13-15).

3.2 Thrust bearing rig


A thrust bearing rig (see Fig.3) developed in house is used to compare
experimental data to the prediction model. The thrust bearing to be tested is
seating in a cartridge; the axial load is provided by a pneumatic actuator and
measured by a force transducer; the thrust oil film thickness is measured by using
proximity probes; a second load cell is implemented to measure the thrust oil
resistive torque (and consequently thrust power loss) and many other operating
parameters such as the rotational speed, the oil inlet temperature or the oil inlet
pressure are being under control.

Actuator

Load cell

Cartridge
Driver

Figure 3: Thrust bearing rig

4 PARAMETRIC STUDY – EXPERIMENTAL COMPARISON

Correlation between the prediction code and test data in term of axial load and
power loss are presented for two typical turbocharger thrust bearings (see Fig.4)
operating in steady state. Speed ranges of the studied thrust bearings A and B,
start-up events taken aside, respectively are [500 – 3000 Hz] and [500 – 2500 Hz].
Therefore, the simulation conditions presented in Table 1 are relevant to
turbocharger operating conditions. Results are summarized in Figure 5.

454
Figure 4: Thrust bearings tested on the rig for numerical correlation

Table 1: Tested thrust bearings geometries and operating conditions


Supply Supply
Number ID OD Speed Oil
pressure temperature
of pads [mm] [mm] [Hz] Grade
[bar] [°C]
Bearing
6 8.6 14.15 2500 0 120 0W30
A
Bearing SAE
4 13.25 19.65 1700 0 100
B 10W30

The prediction model takes into account operating condition of the two sides of
thrust bearing. Indeed, while the active side of the bearing works with a certain
thickness, the remaining clearance leaves a determined thickness on the other side
of the bearing which creates an axial load in the opposite direction and additional
power loss. This phenomenon occurs on the thrust bearing rig, as well as on
engine.

4.1 Case 1: 100% solvent – Effective temperature


For this case, the MPTT model is not used (i.e. 100% Newtonian solvent is
considered) but temperature effects on the viscosity are still taken into account.

The predicted axial loads correlate with the experimental data obtained on the
thrust bearing rig, especially when the film thickness increases (see Fig.5). The
consideration of the conduction problem in the solids could possibly reduce the
discrepancy observed at high axial loading.

With the assumptions made, the predicted axial load remains acceptable but we
observe a discrepancy for power loss that needs to be investigated.

4.2 Case 2: 100% solvent – THD analysis


In this case, the code is run considering that the lubricant is only made of solvent
(i.e. not using the MPTT model). The energy equation is solved to provide the
temperature at each node of the mesh, enabling the update of the viscosity. In this
thermohydrodynamic analysis, the code assumes that the oil pockets of the bearing
provide fresh oil at 120°C. The discrepancy between experimental data and
predictions in terms of axial load narrows but still exist nearby 15% at low film
thickness (see Fig.5). This may come from the fact that conduction is not calculated
in the solids and adiabatic boundary conditions are used for the fluid-solid
interfaces.

455
Bearing A Bearing B
1.4 1 1.6 1

0.9 0.9
1.4
1.2
0.8 0.8
1.2

Total axial load (dimensionless)


Total axial load (dimensionless)

Power loss (dimensionless)


Power loss (dimensionless)
0.7 0.7

1
0.6 0.6
0.8
~15%
0.5 0.8 0.5

0.6
0.4 0.4
0.6

0.3 0.3
0.4
0.4
0.2 0.2
0.2
0.2
0.1 0.1

0 0 0 0
7 12 17 22 27 7 12 17 22 27
Oil film thickness [µm] Oil flim thickness [µm]

Axial load - Averaged experim ental thrust bearing rig data


Axial load prediction -THD - 100% solvent
Axial load prediction -THD - 98% solvent, 2% polym ers
Power loss - Averaged experim ental thrust bearing rig data
Power loss prediction -THD - 100% solvent
Power loss prediction -THD - 98% solvent, 2% polym ers

Figure 5: Axial load and power loss correlations: Test data vs. predictions

Because of the very small size of turbocharger thrust bearings, it is impossible to


put thermocouple in the pad without disturbing the pressure field. Therefore we
decide to benchmark our code using an experiment done for large bearing by
Dadouche (16) and Ahmed (17). Based on their geometry, oil properties and
operating conditions, we compare their experimental data with our predictions.
They are within the range of 4% which is really acceptable and validates the
temperature model (see Fig.6).

Figure 6: Temperature correlations with Dadouche (16) and Ahmed (17)


experiments

4.3 Case 3: 98% solvent – THD analysis


We now consider that the engine oil is made of 98% of solvent and 2% of
polymers. This is a realistic assumption when discussing with oil manufacturers
(18). For the MPTT model, we set the relaxation time to 5.10-6 s. It suits the
analyses performed on several multigrade oils by Bates et al. (13) who considered
that the lubricants followed a Maxwell model as the relaxation times they found
were in the interval [10-5 ; 10-6]. The slip parameter ξ is set to 0.05 and we
consider that no polymer chain breaks ( = 0 ). In this case, the code predicts
perfectly the axial load (see Fig.5). However, the power loss is overestimated when
the thickness is lower than 15 µm. This gap with the experimental data points

456
narrows as the thickness increases and the predicted power loss finally tends to
stabilize around better values than previous predictions.

5 CONCLUSION

A numerical model is being developed and uses a thermohydrodynamic analysis


coupled with the MPTT model and two viscosity laws. It has been presented for the
first time step but is under development for the unsteady formulation. It is
important to understand whether a quasi steady state approach is sufficient to
model a transient event or the unsteady solution is required. Numerous parameters
for the viscosity models are based on rheological experiments performed on Castrol
SLX Longtec Longlife II (0W30) but device limitations prevent us from obtaining a
precise relaxation time and the remaining properties were set from the literature or
industrial background. Rheological tests in high shear and high frequency ranges
would be a further investigation on the parameters’ accuracy for this oil.

For those two bearings, using adequate parameters for the MPTT model
demonstrates a very good correlation between the predicted axial load and
experimental data. As for the power loss estimations, the prediction doesn’t match
very well with the measured values. Those discrepancies may come from
inaccuracies in the mass flow rate determination and the adiabatic hypothesis for
the film temperature at the fluid-solid interfaces. Some investigations, such as fluid
recirculation or under-lubrication, are being considered to improve the predicted
torque and therefore the power loss. On the other hand, the precision of the power
loss measurement device can be questioned because of the very low friction torque
to measure. An additional in-house system based on the enthalpy measurement
(19) is to be mounted on the rig to double check this quantity. Investigations on
high speed torque meters are being considered as well. The developed model will
be used for the optimization of thrust bearing design performances (thrust load
capacity, power loss, oil flow…) at the development stage of turbocharger. The
target is to getting right the first time with product design prediction tool.

REFERENCE LIST

(1) WILLIAMSON B.P. et al. – The Viscoelastic Properties of Multigrade Oils and
their Effects on Journal-Bearing Characteristics, J. Non-Newtonian Fluid
Mech., 73, p.115-126, 1997
(2) PHAN-THIEN N., TANNER R.I. – A New Constitutive Equation Derived from
Network Theory, Journal of Non-Newtonian Fluid Mechanics, Elsevier, 1977
(3) CROSS M.M. – Rheology of Non-Newtonian Fluids: A New Flow Equation for
Pseudoplastic Systems. J. Colloid Sci., 20(5), p.417-437, 1965
(4) ASTM D341-09, Standard Practice for Viscosity-Temperature Charts for Liquid
Petroleum Products, ASTM International, 2009
(5) TICHY J., BOU-SAÏD B. – Hydrodynamic Lubrication and Bearing Behavior
with Impulsive Loads, Tribology Transactions, Vol.34(4), p.505-512, 1991
(6) BIRD R.B. et al. – Network Theories for Polymer melts and Concentrated
Solutions, Dynamics of Polymeric Liquids, John Wiley & Sons, New York, 1987
(7) EHERT P. – Contribution à l’Etude du Comportement de Mécanismes Lubrifiés
sous Chargements Transitoires, Thèse Mécanique, INSA de Lyon, 246p, 1993
(8) FRENE J. et al. – Lubrification hydrodynamique. Paliers et butées. Collection
de la Direction des Etudes et Recherches d’ED, Paris, Eyrolles, 488 p, 1990
(9) RICHTMYER R.D, MORTON K.W. – Difference Methods for Initial-Value
Problems, Second Edition, Interscience Publications, New York, 1967
(10) SAN ANDREAS L.A. – Internal Testing Report, Texas A&M University, Mech.
Eng. Dept., 2009

457
(11) KONTOGIORGOS V. – Calculation of Relaxation Spectra from Mechanical
Spectra in Matlab, Polymer Testing 29, p.1021-1025, 2010
(12) BAUMGAERTEL M., WINTER H.H. – Determination of Discrete relaxation and
Retardation Time Spectra from Dynamic Mechanical Data, Rheologica Acta,
28, p.511-519, 1989
(13) BATES T. et al. – A Correlation Between Engine Oil Rheology and Oil Film
Thickness in Engine Journal Bearing, SAE Technical Paper Series, 15p., 1986
(14) CROCHET M., WALTERS K. – Computational Rheology : A New Science,
Endeavour, 17:2, p.64-77, 1993
(15) LONGSTRUP T. et al. – The Relationship between Engine Oil Viscosity and
Engine Performance, ASTM International, 108 p., 1977
(16) DADOUCHE A. – Etude des Phénomènes Thermiques dans les Butées
Hydrodynamiques, Thèse Mécanique, 136p., 1998
(17) AHMED S.A. – Contribution à l’Etude des Effets Thermiques et des
Déformations dans les Butées Hydrodynamiques à Géométrie Fixe, Thèse
Mécanique Productique Transport, Université de Poitiers, 195p, 2008
(18) AYEL J. – Lubrifiants – Additifs à action physique ou physiologique, Technique
de l’Ingénieur, BM5354, 2002
(19) LAMQUIN T., GJIKA K. – Power Losses Identification on Turbocharger
Hydrodynamic Bearing Systems: Test and Prediction, Proceedings of ASME
Turbo Expo 2009: Power for Land, Sea and Air, June 8-12 2009, Orlando, FL,
USA, GT2009-59599

458
Influence of holding time in thermal
cycle proven by state-of-the-art
thermomechanical fatigue calculation
in the turbine housing of a turbocharger
K Shoghi
BorgWarner Turbo Systems Ltd, UK

There is a high demand from automotive regulators to reduce emissions of air


pollutants such as oxides of nitrogen (NOx) and oxides of sulphur (SOX). One
solution is to increase the exhaust gas temperature to start chemical reactions in
the exhaust manifold. This increase in temperature combined with thermal cycles
with different amplitudes to which the vehicle is subjected, could lead to
thermomechanical fatigue (TMF) of the manifold and the turbine housing. Different
vehicles have different duty cycles, for example forklift trucks have very cyclic
applications, whereas by contrast a truck travelling on the highway will have low
frequency of cycles. There are other factors which could reduce the TMF life of the
vehicle, for example increase in altitude above the sea level increases the
operating temperature of the turbine housing and hence a reduction in TMF life.

This paper examines a variety of thermal duty cycles obtained from different
vehicles and considers them for analysis. One cycle based on the results of the
analysis is selected for detailed study. Finite element analysis (FEA) is used to
analyse the mathematical model of the turbine housing and evaluate the
temperature and strain distribution in the housing. Experimental methods are used
to measure temperature in the housing and to validate the FEA results.

NOTATION

K= Thermal conductivity [W/mK]


Cp= Specific heat capacity [J/kgK]
ρ= Density [kg/m3]
FO =Fourier Modulus [Dimensionless]
t= Time [s]
x=Length of the element [m]

1. INTRODUCTION

During its life time in service the turbine housing of a turbocharger is subjected to
different thermal cycles. Two such cycles are given in Figure 1 and Figure 2. The
thermal cycles shown in Figure 1 are measured on vehicle ‘A’, as it can be observed
in this application the frequency of the cycles where the temperature changes from
minimum to maximum is high when compared with the cycles shown in Figure 2
measured from vehicle ‘B’. This is as expected since in vehicle ‘A’ application the
torque output from the engine can vary significantly depending on the number of
operations where a load is lifted or lowered whereas in vehicle ‘B’ the torque is

_______________________________________
© The author(s) and/or their employer(s), 2014
461
approximately constant. However, vehicle ‘A’ has a significantly lower mean
temperature than vehicle ‘B’.

Figure 1 Thermal cycle measured for the application with vehicle ‘A’.

Figure 2 Thermal cycle measured for the application with vehicle ‘B’.

Thermal cycle testing on a turbine housing can be performed either on the engine
test bed or gas stand test rig. The thermal cycle test performed on the gas stand
offers more control regarding temperature and the turbine stage can operate
independently of the compressor stage due to the fact that in a gas stand test the
output of the compressor stage does not influence the inlet of the turbine stage as
it does in an engine test. However in the gas stand test, the effect of pulsation is
not simulated as it would be the case on the engine test bed.

The investigation carried out here considers a combination of the thermal cycles ‘A’
and ‘B’. Although there may be rapid spikes in temperature as shown in Figure 1,
under gas stand test condition these ramp-ups and cooling down in temperatures

462
are not very practical to implement. However FEA can simulate these conditions
where the time between maximum and minimum temperature is between 10-15
seconds. Some engine manufacturers may have their own test cycle with different
acceleration, nominal load, deceleration, and idle time allocated. A typical test
cycle is presented in Figure3 showing acceleration between 1-2, nominal load from
2-3 and idle from 4-5. This test cycle offers the higher temperatures as shown in
the thermal cycle ‘B’ and a moderate ramp-up as thermal cycle ‘A’. The test was
performed at two different gas stands due to different test constraints such gas
flow rate and temperature control. As it will be demonstrated in experimental work,
a second cycle was selected to allow for longer time for heat soak.

Figure 3 Typical test cycle used at the gas stand.

2. FINITE ELEMENT AND COMPUTATIONAL FLUID DYNAMICS MODELS

The finite element methodology for a simplified model for thermomechanical fatigue
was discussed by Oberste-Brandenburg et al [1]. However one important aspect is
the relationship between the ramp-up and ramp-down of the cycle, element size
considered, and step time used for the analysis which can influence the values of
plastic strain in the model. The minimum time step, Δt for the largest element size,
Δx in direction of heat flow can be calculated using equation(1). It is assumed that
Fourier modulus, F0 =1 as the components are almost in thermal equilibrium with
their surroundings [2].

4KΔt
F0 = 2
(1)
ρcp (Δ x)

The model is constructed and meshed as shown in Figure 4, consisting of cylinder


head, exhaust manifold, turbine housing and bearing housing.

The heat transfer coefficients were calculated using both 1-D and 3-D approach.
The methodology of 1-D approach was discussed in [1]. For the 3-D method, CFD
code was used to calculate the heat transfer coefficients and the solution was
coupled directly to structural FEA code. The coupled transient heat transfer and
structural model was constrained as shown in Figure 4. In order to prevent over
constraining the turbine housing and the manifold, as these will result in artificial
stress concentration at these locations, it is important to allow the growth of the
structure from the central position. Hence all degrees of freedom in X, Y, and Z
directions can be constrained using the surface and the constructed lines on the

463
cylinder head to prevent rigid body motion without over constraining the model.
The cycle shown in Figure 3 was used in the transient heat transfer model. To
account for the accumulated plastic strain in the turbine housing at different
temperatures, multilinear kinematic hardening model, with temperature dependent
material, was used.

Figure 4 Showing the mesh and constraint of the turbine housing assembly.

The Ansys CXF code was used to construct the CFD model of the turbine housing
assembly. There are different ways in which the fluid model can be coupled with the
structural model to obtain the same results. For example transient CFD coupled
with steady state heat transfer and structural analysis or conjugate heat transfer
coupled directly with structural analysis and finally steady state CFD analysis could
be used with different time frames combined with transient heat transfer and
structural analysis. This is the option which has been considered due to the
computing cost and speed compared with other methods. The mesh of each fluid
and solid part consisting of: rotating fluid (the space between the turbine blades),
stationary fluid (the gas passage of the turbine housing and manifold) and solid
(turbine housing, manifold, bearing housing, and cylinder head) was imported into
the CFX separately. The boundary condition of the model is defined by fixing the
mass flow rate at the inlet to the turbine housing and the average static pressure at
the outlet which determines the pressure ratio required for the solution.

The total thermal energy equation with a turbulent k-ε model was used for the
analysis.

Figure 5 Giving temperatures in the turbine housing for the time steps at
300 and 800 seconds.

464
The corresponding temperature for different time steps in the cycle from the
transient heat transfer analysis is given in Figure 5.

As shown in Figure 6 the FEA results have taken more time for the turbine housing
to cool down. This is due to the thermal inertia stored in the turbine housing
assembly.

Figure 6 Showing the thermal cycle output from an FEA file.

The corresponding plastic strain was obtained from the structural part of the model
as the criterion for crack initiation and failure is based on the limit of this plastic
strain. In order to simulate the effect of heat soak, a second cycle was considered
with a ramp-up time of 200 seconds as this cycle offers more time for the heat soak
when compared with the 100 second cycle. This is given in Figure 7.

It is worth noting that these cycles are designed for use with different gas stands
which may have different characteristics. A cycle with a ramp-up time of 100
seconds may be possible on one gas stand, but not be possible with another. As it
shall be demonstrated in experimental work, the temperature is still climbing after
400 seconds and as the strength of the turbine material decreases with the rise in
temperature, the 100 seconds ramp-up cycle would not be suitable for this test,
unless the heat soak time was increased.

Figure 7 Showing the thermal cycle with a ramp-up of 200 seconds.

465
3. EXPERIMENTAL WORK AND GAS STAND TESTING

The tests were performed on two different gas stands. One with capability of using
a shorter ramp-up time while the other requires longer ramp-up and ramp-down
times. The turbine housing was fitted with thermocouples as shown in Figure 8. The
test was performed on a gas stand using the cycle shown in Figure 3. After 1200
cycles the housing was examined for cracks.

Figure 8 Positions of thermocouples on the turbine housing.

Mass flow Temperature


control control

Figure 9 Actual test cycle on a gas stand.

The cycle given in Figure 3 was simulated during the gas stand test, however in
order to prevent the final temperature of the cycle from exceeding its intended
maximum value, initially during the ramp-up the cycle was controlled by means of
mass flow through the inlet up to a nominal value, the final temperature of the
cycle was controlled by means of temperature control as given by Figure 9. It was
necessary to increase the time of the cycle to achieve the temperature required at
the end of the cycle. It is worth stating that if a tolerance on the final temperature
is allowed, then it is possible to simulate the desired cycle.

In order to see the effect of heat soak a cycle with the ramp-up of 200 seconds was
tested on a gas stand as discussed in section 3. The results are shown in Figure 10.
It is important to note that the average temperature of the turbine inlet gas is
higher than the metal temperatures measured by the thermocouples. Also that the
maximum gas temperature for each thermocouple occurs at the same time interval
because if the cooling stage of the cycle starts before the maximum temperature is
achieved, the true values of the strain in the structure will not be representative of
the true nature of the cycle.

466
800

700

600
TC1
Temperature °C

500
TC2

400 TC4

TC6
300
TC7
200
TC8

100 Avrg at turbine inlet

0
0 500 1000 1500
Time (Seconds)
Figure 10 Test cycle obtained from a gas stand test.

4. VALIDATION

The comparison between the FEA thermal cycle with the ramp-up of 200 seconds
and the experimental work on the gas stand with the same ramp-up time for
position 1 is given in Figure 11. As there are two thermocouples fitted on the inlet
flange of the turbine housing (please refer to Figure 8 in section 3), this position
was selected for the comparison with the FEA results. It can be observed that there
is a good correlation between the experimental work and the FEA model. The FEA
model for the more severe cycle given in Figure 3 discussed in section 3 with the
ramp-up of 100seconds is shown in Figure 12. It is interesting to note that the
values for the temperature from the FEA model measured at the inlet flange of the
turbine housing are nearly the same. This result is consistent with the temperature
measurements on the previous cycle.

The housing was sectioned along its circumference through the divider wall
between the scrolls in the volute. The correlation between the high strain locations
obtained from the FEA model and the turbine housing section is given in Figure 13.
It is important to observe that the mode of failure starts at the tongue in the
turbine housing, this is the case in majority of the applications. It is due to the fact
that the tongue heats up much quicker than the rest of the turbine housing and the
resulting temperature difference combined with the geometry of the volute results
in crack initiation at this location. However it is important that the crack does not
propagate to the outer skin as this will result in leakage of exhaust gas through the
housing and loss of performance and failure. The study of fatigue crack growth and
failure prediction is discussed by F.Laengler et al [3] and S.Bist et al [4].

467
Figure 11 Comparisons between test and FEA for the thermal cycle
with 200 seconds ramp-up.

Figure 12 FEA for the thermal cycle with 200 seconds


ramp-up at 600 seconds.

Figure 13 Comparisons between test and FEA.

468
SUMMARY

1- Both F.E.A. and CFD methodology were used to simulate thermal cycles in a
turbine housing assembly of a turbocharger.
2- Two different thermal cycles were simulated.

CONCLUSIONS AND FUTURE WORK

1- As it was demonstrated by experimental methods, FEA and CFD codes there


must be enough time allowed for the heat soak during the thermal cycle test,
otherwise the full effect of the damage due to thermal stresses and strains in the
structure will not be captured.
2- There was good correlation between experimental results and theoretical work.
3- The FEA model successfully predicted the mode of failure and this matched very
well to the location of cracks developed in the turbine housing during the
durability test.

Experimental work is planned to measure the values of strain in the housing by


using:
(i) Strain gauges
(ii) Neutron diffraction methodology (this method is ideal for the locations where it
is not practical to fit a strain gauge).

Further test on engine test bed is recommended and investigation into effect of
pulsation in the turbine housing assembly.

REFERENCES

1. C.Oberste-Brandenburg, K.Shoghi, M.Gugau, and F.Kruse, On the influence of


thermal boundary conditions on Thermo Mechanical Analysis of turbine housing
of a turbocharger. 10th International Conference on Turbochargers and
Turbocharging. 15-16 May 2012. London, pp.86-93.
2. Ansys theory manual, chapter 3, Transient Thermal Analysis Nov 2009. P47.
3. F.Laengler, A.Scholz, H.Alesksanoglu, T.Mao, Validation of a phenomenological
lifetime estimation approach for application on turbine housing of turbochargers.
9th International Conference on Turbochargers and Turbocharging. 19-20 May
2010. London, pp.193-205.
4. S.Bist, R.Kannusamy, P.Tayal, E.Liang, Thermomechanical fatigue crack growth
and failure prediction for turbine housings. 9th International Conference on
Turbochargers and Turbocharging. 19-20 May 2010. London, pp.207-215.

469
11th International Conference on
Turbochargers and Turbocharging
Combustion Engines & Fuels Group Organising Committee
Dr Kian Banisoleiman (Chairman) Lloyd’s Register
Dr Roland Baar Technische Universität Berlin
Andrew Banks Ricardo
Steve Birnie BorgWarner
Dr Chris Brace University of Bath
Dr Geoff Capon Ford
Dr Ennio Codan ABB
Gavin Donkin Honeywell
Dr-Ing Dietmar Filsinger IHI
Pierre French Cummins Turbo Technologies
Dr Seiichi Ibaraki Mitsubishi
Per-Inge Larson Scania
Prof Ricardo Martinez-Botas Imperial College London
Takashi Otobe Honda
Prof Joerg Seume Hanover University
Dr Les Smith Jaguar Land Rover
Prof Stephen Spence Queens University Belfast
Bernd Wietholt Volkswagen
11th International Conference on
Turbochargers and Turbocharging

13–14 MAY 2014


THE BRITISH MUSEUM, LONDON

Conference Proceedings sponsored by:

AMSTERDAM  BOSTON  CAMBRIDGE  HEIDELBERG  LONDON


NEW YORK  OXFORD  PARIS  SAN DIEGO
SAN FRANCISCO  SINGAPORE  SYDNEY  TOKYO
Woodhead Publishing is an imprint of Elsevier
Woodhead Publishing is an imprint of Elsevier
80 High Street, Sawston, Cambridge CB22 3HJ, UK
225 Wyman Street, Waltham, MA 02451, USA
Langford Lane, Kidlington, OX5 1GB, UK

First published 2014, Woodhead Publishing


© The author(s) and/or their employer(s) unless otherwise stated, 2014
The authors have asserted their moral rights.

This book contains information obtained from authentic and highly regarded sources.
Reprinted material is quoted with permission, and sources are indicated. Reasonable
efforts have been made to publish reliable data and information, but the authors and the
publisher cannot assume responsibility for the validity of all materials. Neither the
authors nor the publisher, nor anyone else associated with this publication, shall be liable
for any loss, damage or liability directly or indirectly caused or alleged to be caused by
this book.
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 the publisher.
Permissions may be sought directly from Elsevier’s Science & Technology Rights
Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333;
email: permissions@elsevier.com. Alternatively you can submit your request online by
visiting the Elsevier web site at http://elsevier.com/locate/permissions, and selecting
Obtaining permission to use Elsevier material.

Trademark notice: Product or corporate names may be trademarks or registered trade-


marks, and are used only for identification and explanation, without intent to infringe.

British Library Cataloguing in Publication Data


A catalogue record for this book is available from the British Library.

Library of Congress Control Number: 2014940342

ISBN 978-0-081000-33-5 (print)


ISBN 978-0-081000-34-2 (online)

For information on all Woodhead Publishing publications


visit our website at http://store.elsevier.com

Produced from electronic copy supplied by authors.


Printed in the UK and USA.
Printed in the UK by 4edge Ltd, Hockley, Essex.

S-ar putea să vă placă și