Anda di halaman 1dari 8

Sensors and Actuators A 113 (2004) 8693

A microturbine for electric power generation


Jan Peirs, Dominiek Reynaerts , Filip Verplaetsen
Department of Mechanical Engineering, Katholieke Universiteit Leuven, Celestijnenlaan 300B,
Leuven 3001, Belgium
Received 15 September 2002; received in revised form 25 July 2003; accepted 11 January 2004
Available online 25 February 2004

Abstract
A single-stage axial microturbine has been developed with a rotor diameter of 10 mm. This turbine is a first step in the development
of a microgenerator that produces electrical energy from fuel. The turbine is made of stainless steel using die-sinking electro-discharge
machining. It has been tested to speeds up to 160,000 rpm and generates a maximum mechanical power of 28 W with an efficiency of
18%. When coupled to a small generator, it generates 16 W of electrical power, which corresponds to an efficiency for the total system
of 10.5%.
2003 Elsevier B.V. All rights reserved.
Keywords: Microturbine; Microgenerator; Power MEMS; EDM

1. Introduction

2. Microturbines-scale effects

Most portable devices use batteries for their power supply. Current Li-ion batteries have energy densities up to
0.5 MJ/kg but still offer limited autonomy to for instance
laptops and cellular phones. Charging times also pose problems. Fuel, on the other hand, offers a much higher energy
density of about 45 MJ/kg, and the reservoir can easily be
refilled. Therefore, several groups are working on the development of micro power generators based on fuel cells [13],
thermo-electric devices [4,5], Stirling engines [6,7], reciprocating internal combustion engines [8,9], Wankel motors
[10], and gas turbines [1114].
Specific about the microturbine presented in this paper is
that it is an axial turbine produced with electro-discharge
machining (EDM). The microturbine developed at MIT [11]
is a radial turbine with a rotor diameter of 4 mm, produced
lithographically in Si or SiC. The microturbine developed
at Stanford [12] is an axialradial turbine with a rotor diameter of 12 mm. The silicon nitride rotor is produced by a
gel-casting technique using a wax mould. Teams at Tohoku
University [13] and the University of Tokyo [14] use as well
radial as axial-radial designs.

The reduction of scale has several effects on the performance and construction of the turbine, and on the fuel
choice.

Corresponding author. Tel.: +32-16-322640; fax: +32-16-322987.


E-mail address: Dominiek.Reynaerts@mech.kuleuven.ac.be
(D. Reynaerts).

0924-4247/$ see front matter 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.sna.2004.01.003

2.1. Increasing power density with miniaturisation


Dimensional analysis shows that the power P generated
by a turbomachine is proportional to the density of the
gas, the fifth power of the diameter D, and the third power
of the rotational speed n:
P D5 n3

(1)

The power per unit volume (V D3 ) is thus:


P
D2 n3
V

(2)

For a known pressure ratio and constant inlet conditions, the


speed of the fluidum at the exit of the nozzles is a constant,
independent of the size of the nozzles. Therefore, the circumferential speed of the turbine is constant, independent
of the turbine size. This means that at optimal working conditions, size and rotational speed are inversely proportional:
D n = constant
The power density is thus inversely proportional to size:

(3)

J. Peirs et al. / Sensors and Actuators A 113 (2004) 8693

P
1

V
D

(4)

The power density of turbomachines increases thus with


miniaturisation. By replacing one turbomachine by k smaller
machines with each one kth of the power, the total volume
and mass can be reduced by a factor k. This mass reduction
is advantageous for aeronautical and space applications. Replacing one large turbomachine by several smaller ones also
improves the reliability, especially when a few redundant
devices are added [12].
However, the above scale laws should be interpreted with
care as they are strictly speaking only valid for constant
Reynolds numbers. As will be shown in Section 2.3, the
Reynolds number decreases with miniaturisation, which has
a negative influence on the power density.
The power density of miniature turbomachines is also limited for technological reasons. Small turbomachines cannot
be made with the same relative accuracy and detail as large
ones, so the performance will be worse than predicted by
the scale laws.
2.2. High rotational speeds
The rotational speed is inversely proportional to the diameter. For a turbine diameter of 10 mm the rotational speed
corresponding to sonic flow at the outlet of the nozzles is
already 325,000 rpm. For a turbine of 5 mm diameter the rotational speed rises to 650,000 rpm. This is clearly beyond
the limit of ball bearings such that air or fluid bearings are
required [15,16].
The speed is also limited by the bursting speed of the
rotor. This bursting speed (in terms of circumferential speed)
is a constant for a certain rotor geometry and material, and
thus independent of size. The burst limit corresponds thus
to a rotational speed inversely proportional to size. Thus,
the bursting speed and the speed resulting from the pressure
ratio follow the same scale law.
2.3. Low Reynolds number
The Reynolds number Re characterises the flow and is
defined as:
Re =

uL
,

(5)

with u a characteristic speed, L a characteristic dimension of


the flow channels, and the kinematic viscosity. The speed
can be considered independent of size as it depends only
on the pressure ratio. The Reynolds number is thus proportional to size and, therefore, decreases with miniaturisation.
For small turbines the flow will be less turbulent and more
laminar. This means that the viscous friction losses will be
higher and that mixing of the fuelair mixture will be slower,
both having a negative impact on efficiency and power
density.

87

2.4. Fast start-up and stop


The small inertia of the rotor allows start-up and stop
of the turbine within a fraction of a second. This allows
power regulation using pulsed operation [17]. In that case,
the turbine can operate at its optimal speed and generate a
fixed voltage when coupled to a generator.
2.5. Increased heat transfer
The increasing surface-to-volume ratio results in higher
heat transfer. The higher thermal losses have a negative effect
on the efficiency of the turbine, and may even cause flame
extinction. At very small sizes, the heat generated by the
combustion minus the heat loss is not longer sufficient to
ignite the mixture. Another effect is that thermal insulation
between the hot parts and the cold parts becomes more and
more a problem.
2.6. Shorter residence time
The residence time of the fuelair mixture is proportional
to the size of the turbine. This means that the time for mixing and combustion decreases for smaller turbines. When
a conventional turbine would be made 500 times smaller,
the residence time would be reduced to the characteristic
kinetic reaction time of the fuelair mixture (0.010.1 ms)
[18]. Therefore, the relative size of the combustion chambers should be increased and fuels with shorter combustion
time and shorter combustion delay should be used.
In current combustion chambers, mixing takes a large part
of the residence time. Therefore, it would be an improvement to pre-mix air and fuel before they enter the combustion chamber. A disadvantage of pre-mixing is that the
fuel-rich and stable primary zone in the combustion chamber disappears. Stable combustion of the lean mixture can
be obtained by the use of hydrogen as fuel and by the use of
catalysts.
Hydrogen is an ideal fuel in many aspects. Compared
to hydrocarbon fuels, hydrogen has a higher mass-specific
combustion energy, a higher evaporation speed, a higher diffusion speed, a shorter reaction time, a considerably higher
flame propagation speed, wider ignition limits, a lower ignition energy, and lower radiation losses. The wide ignition
limits remove the need for a relatively rich primary combustion zone.

3. Turbine design
In a first phase of the project, the problem has been scaled
down to a turbine powered by compressed air. Compressor, combustion chamber, and generator have been left out
and will be addressed in a later phase. The microturbine is
a single-stage axial impulse turbine (Laval turbine). Expansion of the gas takes place in the stationary nozzles and not

88

J. Peirs et al. / Sensors and Actuators A 113 (2004) 8693

Fig. 1. Microturbine design.

between the rotor blades. This type of turbine has been chosen because of its simple construction.
Fig. 1 shows an exploded view and an assembly of the
microturbine design. The compressed air enters via a standard pneumatic connector (1) and expands over the stationary nozzles (3) where it is deflected in a direction tangential
to the turbine rotor (5). After the air has passed the rotor
blades, it leaves the device through the openings in the outlet
disc (6). Screwing the pneumatic connector in the housing
(8) presses the stationary nozzle disc against a shoulder in
the housing. The rotor blades, wheel and axis are one monolithic part. The rotor is supported by two ball bearings (4),
one mounted in the stationary nozzle disc and one mounted
in the outlet disc. The outlet disc is locked in the housing
by a circlip (7).
The diameter of the turbine rotor is 10 mm. The housing
has a diameter of 15 mm and is 25 mm long. All parts, except the pneumatic connector and the circlip, are made of
stainless steel.
The nozzles are designed for subsonic flow and, therefore,
have a converging cross section. Sonic speed is reached for a
relative supply pressure of 1 bar. The exit losses (remaining
kinetic energy in the exhaust) are minimal when the turbine
is designed for a u/c1 ratio of 0.5, with u the circumferential
speed and c1 the absolute speed at the nozzle exit. At 1 bar,
c1 reaches sonic speed resulting in an optimal turbine speed
of 420,000 rpm. As this is too high for the bearings, the
turbine has been designed for a u/c1 ratio of 0.25, and is
operated below its optimal speed of 210,000 rpm.

4. Turbine production
The different parts of the turbine are produced by turning
and EDM. The nozzle disc and rotor are the most complex
parts. In a first step, their cylindrical surfaces are machined
on a lathe. In a second step, the nozzles and blades are
created by die-sinking EDM as illustrated for the rotor in
Fig. 2. The rotor is clamped in a rotary head which is indexed
with steps of 30 . A prismatic copper electrode with a cross
section having the shape of the air channels between the
blades is sunk into the turbine wheel by EDM. The electrode

Fig. 2. Machining of the rotor blades by EDM.

Fig. 3. Subassembly of nozzle disc, turbine rotor, and bearings. The rotor
has a diameter of 10 mm.

is produced by wire-EDM. Fig. 3 shows a subassembly of


nozzle disc, rotor, and bearings.

5. Generator
The turbine has been coupled to a small brushless dc
motor that is used as a three-phase generator (see Fig. 4).
The motor (Faulhaber, type 1628 T024B K312) has about
the same size as the turbine: 16 mm in diameter and 28 mm
long. Turbine and generator are coupled to each other by
an elastic tube that serves as a flexible coupling.

6. Mechanical output
Torque and power of the turbine have been tested up to a
speed of 100,000 rpm. For this purpose, a 30 mm diameter

J. Peirs et al. / Sensors and Actuators A 113 (2004) 8693

89

3.5
1 bar

Torque (Nmm)

3
2.5

0.8 bar

0.7 bar

1.5

0.6 bar
0.5 bar

Friction torque

0.5
0

0.2 bar
0

20

40

Fig. 4. Turbine coupled to the generator.

Inertia wheel

100

120
3
x 10

Fig. 6. Torque generated by the turbine as a function of speed and supply


pressure.
30
1 bar

Mechanical power (W)

25
20
0.8 bar
15

0.7 bar
0.6 bar

10

0.5 bar

5
0

Friction loss
0.2 bar
0

20

40

60
80
Speed (rpm)

100

120
3
x 10

Fig. 7. Mechanical power of the turbine as a function of speed and supply


pressure.
20
1 bar

18
16
Turbine efficiency (%)

brass wheel has been fixed to the turbine axis, as shown


in Fig. 5. An optical sensor measures the rotation of the
wheel in a contactless way: two vanes on the wheel interrupt the optical path of a photosensor. The turbine is tested
by switching on the pressure and accelerating the turbine to
100,000 rpm. The torque is then derived from the acceleration and the moment of inertia of the wheel and turbine rotor.
As the turbine passes through the whole speed range, acceleration, torque and power are known as a function of speed.
When the turbine is rotating at full speed, the pressure is
switched off and a new measurement is done while the turbine slows down. This gives the friction torque as a function
of speed. Friction mainly occurs between the wheel with
vanes and the surrounding air. The friction torque and power
are added to the results of the acceleration test to obtain the
total torque and power of the turbine.
Figs. 6 and 7 show torque and mechanical power as a
function of speed for different supply pressures up to 1 bar.
The maximum torque and power are, respectively, 3.7 Nmm
and 28 W. The dashed lines represent the friction losses determined with the deceleration test.
At 1 bar, the turbine consumes 8 Nm3 /h of compressed
air, which corresponds to a power consumption of 152 W
when assuming an ideal isentropic expansion. This means
that the mechanical efficiency of the turbine lies around 18%.
Fig. 8 shows the turbine efficiency as a function of speed
for different supply pressures.
The dips in the characteristics at high speed are caused
by the measurement method as they always occur at the

60
80
Speed (rpm)

0.8 bar

14

0.7 bar
0.6 bar
0.5 bar

12
10
8
6
4

Turbine

Vane

Photosensor

Fig. 5. Set-up to measure the mechanical output of the turbine. The output
torque is derived from the acceleration of the inertia wheel.

20

40

60
80
Speed (rpm)

100

120
3
x 10

Fig. 8. Efficiency of the turbine (compressed air to mechanical power).

90

J. Peirs et al. / Sensors and Actuators A 113 (2004) 8693


12

18
16

10
1 bar

Total efficiency (%)

Electrical power (W)

14
12

0.9 bar

10
8

0.8 bar
0.7 bar

6
0.6 bar

0.7 bar 0.8 bar


0.9 bar
0.6 bar

0.5 bar
0.4 bar

0.3 bar

0.5 bar

1 bar

0.2 bar

0.4 bar

0.2 bar

20

0.3 bar

40

60

80

100

Speed (rpm)

120

140

160

20

40

x 10

60

80

100

120

140

160
3

x 10

Speed (rpm)

Fig. 9. Electrical power generated by the total system (turbine plus


generator).

Fig. 10. Total efficiency (compressed air to electricity).

maximal speed, even for different loads and pressures. In


reality, power and efficiency increase further with speed to
reach their maxima theoretically at 210,000 rpm (for 1 bar).
These speeds can be reached using a smaller load.

value (6 + 7) is obtained with a deceleration test of the


turbine without generator and without external load. The
loss associated with the leak flow around the turbine wheel
(2) and the exit losses (8) are calculated from the known
air speeds. The expansion losses (1), incidence losses (4)
and blade profile losses (5) are calculated using friction and
loss coefficients known from large turbines and may be less

7. Electrical output
To measure the electrical power output of the system,
the generator is connected to a variable three-phase load
consisting of three potentiometers (range: 2 k , 10 turns).
In contrast with the mechanical tests, the electrical tests are
performed at constant speed. The speed of the turbine, which
is measured from the frequency of the generator voltage, is
controlled by varying the load. Fig. 9 shows the electrical
power measured for different supply pressures and speeds.
At a pressure of 1 bar, the maximal electrical power is 16 W
and is reached at a speed of 100,000 rpm. Measurements
show that the air flow and input power depend only on the
supply pressure and not on speed or load. Therefore, the
input power is the same as in the mechanical test at 1 bar,
i.e. 152 W. Fig. 10 shows the total efficiency (compressed
air to electricity) as a function of speed and for different
supply pressures. The maximal total efficiency is 10.5% and
is reached at a speed of 100,000 rpm.

Input power
152 W - 100 %
Expansion losses
15 W - 9.8 %
Leak flow around rotor
4 W - 2.6 %
Obstruction losses
1 W - 0.7 %
Incidence losses
2 W - 1.3 %

Mechanical
power
28 W - 18.4 %

Blade profile losses


48 W - 31.6 %
Ventilation losses +
bearing friction
2 W - 1.3 %
Exit losses
52 W - 34.2 %
Losses in coupling
2 W - 1.3 %
Generator losses
10 W - 6.6 %

8. Sankey diagram
The energy flow and the different losses are illustrated
in the Sankey diagram shown in Fig. 11. The diagram
is generated for a supply pressure of 1 bar and a speed
of 100,000 rpm. This corresponds to the working point
at which the maximal electrical power and maximal total
efficiency are reached. Input power, mechanical power,
electrical power and the combination of ventilation losses
(6) and bearing friction (7) are measured values. This last

Electrical power
16 W - 10.5 %

Legend
Measured

Calculated

Difference from
other values

Fig. 11. Sankey diagram for a supply pressure of 1 bar and a speed of
100,000 rpm.

J. Peirs et al. / Sensors and Actuators A 113 (2004) 8693


100
Helicopter
Tank
Power/mass (kW/kg)

accurate. The generator losses (10) are derived from the


manufacturers data sheets. The obstruction losses (3) and
the losses in the coupling (9) are derived as the difference
between the calculated and measured values.
The major losses are the blade profile losses and the exit
losses. The large blade profile losses can be explained by
the increased friction in miniature systems (small channels
and low Reynolds numbers). The high exit losses can be
explained by the low u/c1 ratio (0.25 instead of 0.5 in the
optimal case). Additionally, the turbine operates below its
optimal speed because the ball bearings limit the speed. Both
factors result in higher air speeds at the turbine exit, and
thus higher exit losses.

91

L-1

Ship

M -1/3

Power system

10

1
10

100

1000

10000

100000

Mass (kg)

9. Power density
Table 1 gives the masses of the different parts. The turbine housing and pneumatic connector are not optimised towards mass and are responsible for about 86% of the weight.
Therefore, the mass of the turbine can be substantially reduced by optimising these parts.
The mechanical power density of the turbine, defined as
the mechanical power output of the turbine (28 W) divided
by the mass of the turbine (36 g), is about 780 W/kg. As a
comparison, Fig. 12 shows the power density of commercial
gas turbines for helicopters, tanks, ships and power generators from General Electric, Rolls-Royce, and Pratt & Witney. Most of these turbines have a power density between 4
and 10 kW/kg. The tank turbine has a lower power density
of 1 kW/kg. The power density of the current microturbine
is thus 510 times lower than the power density of large turbines. This figure can be improved by optimising the mass
of the connector and housing, but on the other hand, a compressor and combustion chamber have to be added to obtain
a turbine comparable to the ones mentioned in Fig. 12. A
comparison can also be made with a silicon air turbine with
a rotor diameter of 4.2 mm developed at MIT [15]. It generates 5 W of power and achieves a power density of more
than 4 MW/m3 (about 2 kW/kg), about 2.5 times more than
the turbine presented in this paper.
Table 1
Masses of the different parts
Part

Mass (g)

Turbine
Pneumatic connector (1)
Ring (2)
Nozzle disc (3)
Small bearing (4)
Large bearing (4)
Rotor (5)
Outlet disc (6)
Circlip (7)
Housing (8)
Generator

36
15.8
0.77
1.78
0.03
0.07
1.63
0.35
0.27
15
30

Total (turbine + generator)

66

Fig. 12. Power density of commercial gas turbines. Data from General
Electric, Rolls-Royce, and Pratt & Witney. The line represents the evolution of the power density as predicted by the scale laws.

The line shown in Fig. 12 shows the evolution of the


power density as predicted by the scale laws derived
above. It is clear that the data does not correspond to this
scale law. Extrapolation predicts even a power density of
100300 kW/kg for the current microturbine, more than two
orders of magnitude above the measured values. As mentioned in Section 2.1, this can be explained by technological
limitations and increased friction losses.
The electrical power density of the microturbine, defined
as the electrical power produced by the generator (16 W)
divided by the combined masses of turbine and generator
(66 g), is about 240 W/kg.

10. Conclusion
A 10 mm diameter axial microturbine with generator has
been developed and successfully tested to speeds up to
160,000 rpm. It generates a maximum mechanical power of
28 W with an efficiency of 18%. Power and efficiency are
mainly limited by the maximal speed of the ball bearings.
The main losses are the blade profile losses (32%) and the
exit losses (34%). Higher speeds can considerably reduce
the exit losses and therefore increase efficiency and power.
Currently, the power density is 780 W/kg, about 510 times
lower than for large turbines. However, higher speeds and
optimisation of the housing can considerably increase this
figure. When coupled to a small generator, the system generates 16 W of electrical power, corresponding to a total
efficiency of 10.5%.

11. Future work


The first goal is to increase the efficiency of the turbine, mainly by decreasing the exit losses. Therefore,

92

J. Peirs et al. / Sensors and Actuators A 113 (2004) 8693

the allowable speed of the turbine will be increased to


200,000300,000 rpm by using special high-speed ball
bearings. Later on, air or fluid bearings can be introduced to attain even higher speeds. Another possible approach is to decrease the speed by using a multiple-stage
design.
A compressor and a combustion chamber will be added
to finally come to a microgenerator running on fuel. The
compressor is currently under development.

[11]

[12]
[13]

Acknowledgements
This research is sponsored by the Belgian programme
on Interuniversity Poles of Attraction (IAP5/AMS) initiated
by the Belgian State, Prime Ministers Office, Science Policy Programming. The authors assume the scientific responsibility of this paper. The authors wish to thank Michael
Poesen and Pieterjan Renier for their contribution to this
work.

[14]

[15]

[16]

References
[1] U. Groos, C. Hebling, U. Wittstadt, P. Hbner, Fuel cell development at Fraunhofer ISE, European Fuel Cell News 8 (4) (2002) 4
5.
[2] W.Y. Sim, G.Y. Kim, S.S. Yang, Fabrication of micro power
source (MPS) using a micro direct methanol fuel cell (DMFC)
for the medical application, MEMS 2001, in: Proceedings of the
14th IEEE International Conference on Micro Electro Mechanical
Systems, Interlaken, Switzerland, 2125 January 2001, pp. 341
344.
[3] K.-B. Min, S. Tanaka, M. Esashi, MEMS-Based Polymer Electrolyte Fuel Cell, Technical Digest Power MEMS 2002, Int. Workshop on Power MEMS, Tsukuba, Japan, 1213 November 2002,
pp. 6467.
[4] L. Sitzki, K. Borer, E. Schuster, P.D. Ronney, S. Wussow, Combustion
in microscale heat-recirculating burners, in: Proceedings of the Third
Asia-Pacific Conference on Combustion, Seoul, Korea, 2427 June
2001, pp. 473476.
[5] K. Yoshida, K. Kobayashi, T. Nakajima, D. Satoh, S. Tanaka, M.
Esashi, Micro-Thermoelectric Generator using Catalytic Combustor, Technical Digest Power MEMS 2002, Int. Workshop on Power
MEMS, Tsukuba, Japan, 1213 November 2002, pp. 98101.
[6] N. Nakajima, K. Ogawa, I. Fujimasa, Study on micro engines
miniaturizing stirling engines for actuators and heatpumps, micromechanics and MEMS, in: Proceedings of the IEEE Micro Electro
Mechanical Systems Workshop, Salt Lake City, UT, USA, February
1989, pp. 145148.
[7] S. Backhaus, G.W. Swift, A thermoacoustic stirling heat engine,
Nature 399 (1999) 335338.
[8] D.H. Lee, D.E. Park, J.B. Yoon, S. Kwon, E. Yoon, Fabrication and
test of a MEMS combustor and reciprocating device, J. Micromech.
Microeng. 12 (2002) 2634.
[9] K.D. Annen, D.B. Stickler, J. Woodroffe, Miniature internal combustion engine (MICE) for portable electric power, in: Proceedings
of the 23rd Army Science Conference, Orlando, Florida, USA, 25
December 2002, paper FP-02.
[10] K. Fu, A.J. Knobloch, F.C. Martinez, D.C. Walther, C. FernandezPello, A.P. Pisano, D. Liepmann, Design and fabrication of a
silicon-based MEMS rotary engine, in: Proceedings of the 2001

[17]

[18]

ASME International Mechanical Engineering Congress and Exposition, New York, 1116 November 2001, pp. 303308.
A.H. Epstein, S.A. Jacobson, J. Protz, L.G. Frchette, Shirtbuttonsized gas turbines: the engineering challenges of micro high speed
rotating machinery, in: Proceedings of the 8th International Symposium on Transport Phenomena and Dynamics of Rotating Machinery,
Honolulu, Hawaii, 2630 March 2000.
S. Kang, S.-J.J. Lee, F.B. Prinz, Size does matter, the pros and cons
of miniaturization, ABB Rev. 2 (2001) 5462.
K. Isomura, M. Murayama, H. Yamaguchi, N. Ijichi, N. Saji,
O. Shiga, S. Tanaka, T. Genda, M. Hara, M. Esashi, Component Development of Micromachined Gas Turbine Generators,
Technical Digest Power MEMS 2002, International Workshop on
Power MEMS, Tsukuba, Japan, 1213 November 2002, pp. 32
35.
E. Matsuo, H. Yoshiki, T. Nagashima, C. Kato, Development of Ultra
Gas Turbines, Technical Digest Power MEMS 2002, International
Workshop on Power MEMS, Tsukuba, Japan, 1213 November 2002,
pp. 3639.
L.G. Frchette, S.A. Jacobson, K.S. Breuer, F.F. Ehrich, R. Ghodssi,
R. Khanna, C.W. Wong, X. Zhang, M.A. Schmidt, A.H. Epstein,
Demonstration of a Microfabricated High-Speed Turbine Supported
on Gas Bearings, Solid-State Sensor and Actuator Workshop, Hilton
Head Island, SC, USA, 48 June 2000, pp. 4347.
E.S. Piekos, D.J. Orr, S.A. Jacobson, F.F. Ehrich, K.S. Breuer, Design
and analysis of microfabricated high speed gas journal bearings,
in: Proceedings of the 28th AIAA Fluid Dynamics Conference,
Snowmass Village, CO, USA, 29 June2 July 1997, AIAA paper
971966.
A.H. Epstein, S.D. Senturia, O. Al-Midani, G. Anathasuresh, A.
Ayon, K. Breuer, K.-S. Chen, F.E. Ehrich, E. Esteve, L. Frechette,
G. Gauba, R. Ghodssi, C. Groshenry, S. Jacobson, J.L. Kerrebrock,
J.H. Lang, C.-C. Lin, A. London, J. Lopata, A. Mehra, J.O. Mur
Miranda, S. Nagle, D.J. Orr, E. Piekos, M.A. Schmidt, G. Shirley,
S.M. Spearing, C.S. Tan, Y.-S. Tzeng, I.A. Waitz, Micro-heat engines,
gas turbines, and rocket enginesthe MIT microengine project,
in: Proceedings of the 28th AIAA Fluid Dynamics Conference,
Snowmass Village, CO, USA, 29 June2 July 1997, AIAA paper
971773.
I.A. Waitz, G. Gauba, Y.-S. Tzeng, Combustors for micro-gas turbine
engines, J. Fluids Eng. 120 (1998) 109117.

Biographies
Jan Peirs Graduated as mechanical engineer (K.U. Leuven, 1993). He
started his activities as a research assistant at the Division of Production
Engineering, Machine Design and Automation (PMA) of K.U. Leuven
in 1993. He received a PhD degree in mechanical engineering in 2001
from K.U. Leuven. Currently, he is working as a Postdoctoral researcher
at PMA. His research interests include the design of micro-actuators,
medical microsystems, micro-powergenerators, and micromechanical
systems in general.
Dominiek Reynaerts Graduated as mechanical engineer (K.U. Leuven,
1986). He started his activities as a research assistant at the Division
of Production Engineering, Machine Design, and Automation of K.U.
Leuven in 1986. Within the framework of the Erasmus student exchange
program, he stayed as a PhD student at the Scuola Superiore S. Anna
in Pisa in 1990. In 1993, he became a research manager of the division
PMA. He received a PhD degree in mechanical engineering in 1995 from
K.U. Leuven. Currently, he is associate professor at K.U. Leuven. His
current research interests include design and control of multi-fingered
robot grippers, shape memory alloy actuators, precision mechanics, and
micromechanical systems.

J. Peirs et al. / Sensors and Actuators A 113 (2004) 8693


Filip Verplaetsen received his Engineering degree from the Universiteit Gent, Ghent, Belgium, in 1992 and the Diplome en Administration des Entreprises from the Universit Catholique de Louvain,
Louvain-la-Neuve, Belgium, in 1993. He joined the Katholieke Universiteit Leuven , Louvain, Belgium in 1994 as a research assistant and
received the PhD degree in mechanical engineering in 1999. From 1999

93

till 2002, he worked as a Postdoctoral researcher of the Fund for Scientific Research, Flanders (F.W.OVlaanderen) at the same university and
became assistant professor in 2002. His research focuses on industrial
safety, explosion safety, heat transfer enhancement techniques, design of
thermal systems and turbomachinery.

Anda mungkin juga menyukai