Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/355285309

Parameterized, numerical design of a two-wheel Curtis steam turbine for


small scale WHR

Article in MATEC Web of Conferences · January 2021


DOI: 10.1051/matecconf/202134500031

CITATION READS

1 212

2 authors:

Philipp Streit Andreas Weiß


Ostbayerische Technische Hochschule Amberg-Weiden Ostbayerische Technische Hochschule Amberg-Weiden
14 PUBLICATIONS 38 CITATIONS 27 PUBLICATIONS 238 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Philipp Streit on 25 October 2021.

The user has requested enhancement of the downloaded file.


MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

Parameterized, numerical design of a two-wheel


Curtis steam turbine for small scale WHR
Philipp Streit1*, and Andreas P. Weiß1
1OTH Amberg-Weiden, Center of Excellence for Cogeneration Technologies, Kaiser-Wilhelm-Ring
23, 92224 Amberg, Germany, +49 9621 482 3531

Abstract. In contrast to the current trend of converting waste heat into


electricity in the small power range below 100 kWel by means of an ORC
plant, the authors are pursuing the concept of a micro steam power plant
equipped with a micro turbine. Water avoids many of the problems often
associated with organic working fluids, such as flammability, toxicity,
greenhouse gas effect and high fluid costs. However, water vapor makes
turbine design more challenging. The physical reasons for this are repeated,
and thereby it becomes clear why a velocity compounded two wheel Curtis
turbine has been chosen. The used in-house 1D turbine design tool is briefly
introduced. More focus is put on the shortcomings of the implemented 1D
loss model and their negative impact on the current turbine design.
Consequently, the authors continued actual turbine design by a
parameterized approach in 3D CAD/CFD. This approach is explained, and
finally, the CFD flow field and the performance maps of the designed turbine
are discussed. The turbine is currently under construction and will be
installed in 2022 in a waste heat recovery (WHR) plant in
Nuremberg/Germany.

1 Introduction
Waste heat recovery (WHR) is an important building block towards an increase in energy
efficiency in industry and for a successful energy transition. Pehnt et al. 2010 [1] reported
that about 80 TWh of waste heat >140 °C are rejected yearly only in the German industry.
Assuming a conversion efficiency of about 10%, which is realistic for small WHR units, 8
TWh per year could be extracted. This is the annual electricity production of a 1000 MW
steam power plant. In cases where the waste heat temperature is below 300 – 350 °C, an ORC
system is currently the preferred solution for waste heat recovery [2]. In large Combined
Cycle Power Plants (CCPP), a steam turbine reuses the waste heat of the upstream gas turbine
(500 - 600 °C), which leads to an overall electric efficiency of more than 60% [3]. The main
advantage of an Organic Rankine cycle (ORC) is the higher vapor pressure of an organic
fluid compared to water for a given heat source temperature [2]. Thanks to this, a low
temperature heat source can be exploited more effectively. However, the maximum cycle
temperatures of ORC are limited to 300-350 °C due to the danger of degradation of the
working fluid. This issue limits the achievable maximal cycle efficiency in cases in which

* Corresponding author: ph.streit@oth-aw.de

© The Authors, published by EDP Sciences. This is an open access article distributed under the terms of the Creative Commons
Attribution License 4.0 (http://creativecommons.org/licenses/by/4.0/).
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

higher waste heat temperatures are available, like in the glass, steel or cement industry or
downstream of a biogas engine. Water as a working fluid does not suffer from this limitation.
Furthermore, water cannot burn, is not toxic, does not spoil the atmosphere and is
significantly cheaper than any organic working fluid. Thus, the research project in which the
reported work has been carried out aims for a small steam Rankine cycle with about 50 kWel
power using a micro turbine as an expander [4]. The major goal of the current project is to
reduce specific costs (€/kWel) of the existing WHR system [5].
The challenge of designing a steam turbine for small shaft power (< 100 kW el) is at least
twofold. As pointed out by Macchi & Astolfi [2], steam provides a very high isentropic
enthalpy drop for a given maximum and minimum cycle temperature compared to any
organic fluid. The specific work 'his which can be processed by one turbine stage is equal to
a constant C times the square of the rotational speed u on mean diameter D, i.e proportional
to u2 (see Table 1). This is true for any turbine. The value of the constant C depends on the
turbine type. Live steam parameters are 400 °C/17.5 bar in the current project. Exhaust or
condensing pressure is 0.2 bar (60 °C), steam mass flow rate is about 80 g/s. So, the isentropic
enthalpy drop, which must be processed by the turbine, is about 880 kJ/kg [6]. If the constant
C = 1, what is true for a 50 % reaction turbine stage, u = 938 m/s (Table 1). Such a high value
of u is not feasible. The wheel would burst due to centrifugal forces – no matter which
material is used.
The second challenge in designing micro steam turbines is caused by the high enthalpy
drop as well. The power of a turbine is equal to the mass flow rate times the enthalpy drop.
Thus, the higher the enthalpy drop, the smaller the mass flow rate for a given or required
power. The mass flow rate determines the wheel diameter D and the blade height hblade, which
are the main dimensions of the turbine. The diameter D should be as big as possible to achieve
a high u combined with an acceptable angular velocity Z or rotational speed n, respectively.
However, this leads to very small blade heights, which limits this approach. To sum up, the
smaller the aimed power output and thus the corresponding mass flow rate, the smaller the
turbine wheel diameter D and thereby the higher the angular velocity or rotational speed must
be.
In small turbines for power output < 100 kWel, the rotational speed is in the magnitude of
several 104 rpm [5,7–9]. This fact makes bearing design and generator design challenging.
Sophisticated aerodynamic or magnetic bearings must be applied [5,10]) and a high speed
generator must be employed, which is expensive and contributes typically two third to the
specific costs (€/kW) of a small turbo generator. Of course, these physical obstacles can be
overcome by a multi-stage design like in big steam power plants (10 – 1000 MW).
Unfortunately, this solution is too costly and too laborious for those small machines.
Therefore, the authors looked for other approaches. In the history of turbomachines [11,12],
various turbine architectures have been introduced and were employed till the concentration
of power generation started at the beginning of the 20 th century. The multi-stage axial turbine
has finally prevailed in huge thermal power plants of the several 100 MW class.
Already when the steam turbine started to displace the piston steam engine at the end of
the 19th century, three different main turbine concepts existed. There was the simple impulse
turbine of de Laval [13] (Table 1). It can be easily derived that its stage enthalpy drop is 2 x
u2 in such an impulse stage without pressure drop across the buckets [14]. Therefore, for the
given project boundary conditions (400°C, 17.5 bar, 0.2 bar, 0.080 kg/s steam), the required
circumferential speed is 663 m/s compared to 938 m/s of the 50 % reaction turbine, which
was introduced originally by Parsons [15] (Table 1).
Thanks to the constant pressure over the rotor blading, impulse and Curtis stages can be
designed to work with partial admission. This means that a part of the total arc of the annulus
is blocked off. The small mass flow rate is concentrated on a section of the rotor wheel.
Thereby, too small blade heights can be avoided [8,14].

2
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

The third well-known turbine concept is the velocity compounded Curtis stage [16]. The
nozzles and the first wheel might be considered to be identical to the described impulse stage,
however, the main difference is the required circumferential speed u, which is only half of
that of the impulse stage. Thus, the required circumferential speed is only 331 m/s which is
feasible with a steel wheel (Table 1). It has to be mentioned, that the efficiency potentials of
these three turbine architectures differ significantly. Due to higher relative Mach numbers,
higher necessary flow deflections compared to a reaction turbine, the impulse turbine
achieves only about 80 % efficiency, the Curtis about 70 % compared to 90 % of the reaction
turbine. All these turbine concepts can and have been implemented by an axial or by a radial
architecture. The radial architecture might be considered to be advantageous because it
supports stage enthalpy processing by the centrifugal pressure field for a radial inflow turbine
[14]. In contrast, volume flow rate increase due to expansion is facilitated in a radial outflow
turbine [17].
In the current project, the authors have investigated analytically and by an in-house 1D
turbine design tool (1DTDT) [18] various un-common turbine architectures like multi-stage
radial-outflow turbines [17], quasi-impulse cantilever turbines [19] and radial inflow Curtis-
turbines for the given expansion task (400 °C/17.5 bar, 0.2 bar, about 80 g/s steam). Few
uncommon turbine architectures have been promising from an aerodynamic point of view.
However, mechanical pre-investigation showed that only the presently implemented
architecture, a two-wheel axial Curtis-turbine [5], is actually feasible. It is equipped with a
titanium rotor. Thus, the authors aim for an improved axial Curtis turbine with increased
efficiency and reduced costs compared to its predecessor. In the following, the parameterized,
mainly numerical design of this turbine is introduced and discussed.

Table 1 Comparison of different turbine concepts and their velocity triangles for the same stage
enthalpy drop ο݄௜௦ (value of u for the given project boundary conditions)
Reaction turbine Impulse Turbine Curtis turbine (2 wheels)

Turbine
type

velocity
triangles

his 1 x u2 2 x u2 8 x u2

ο݄୧ୱ ଴Ǥହ ͳ ο݄୧ୱ ଴Ǥହ


ሺο݄୧ୱ ሻ଴Ǥହ ൬ ൰ ൬ ൰
‫ݑ‬୭୮୲ ʹ ʹ ʹ
938 m/s 663 m/s 331 m/s
partial
not feasible feasible feasible
admission
efficiency
90% 80% 70%
potential

3
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

2 Methodology

2.1 1D Turbine design tool (1DTDT)


In order to comprehensively and reliably assess and compare different turbine concepts, a
computational tool has been developed for the following considered turbine concepts:
x Axial and radial (cantilever quasi) impulse turbines
x Axial or radial two-wheel velocity compounded Curtis turbines
x Radial inflow, axial outflow reaction turbines (90° IFR)
It has been developed so that the main geometry data and the efficiency of a turbine can be
assessed automatically and quickly with the following inputs to the model:
x Working fluid and mass flow rate
x Pressure at the turbine inlet and outlet and temperature at the inlet
x Wheel mean diameter and rotational speed
The tool is based on a 1D meanline model with an in-house developed appropriate simple
loss model, which is described in more detail in [20,21]. The 1DTDT calculates efficiency
and power output. The loss model for nozzles and buckets allows the prediction of the static
thermodynamic conditions of the working fluid at the nozzle throat, nozzle exit and at the
blading inlet and outlet by applying REFPROP fluid properties [6]. Hence, the velocity
triangles at the mean diameter and the required cross areas can be determined in the form of
nozzle angle, height and width, blade height and the blade angle at inlet and outlet. Based on
these few geometry specifications, an experienced designer is able to build up the 3D model
of the turbine in CAD. With the available 3D CAD model of the turbine, the next design step,
the analysis and optimization of the turbine flow by means of Computational Fluid Dynamics
(CFD), can be carried out. The turbine design data of the 1DTDT is shown in Table 2.

Table 2. Turbine main design data


Parameter Unit Curtis-Steam-Turbine
working fluid - steam
wheel diameter Dmean mm 160
rotational speed n rpm 36,000
degree of admission ε % 35
design pressure ratio PR (ts) - 87.5
degree of reaction r % 20
nozzle exit Mach number Ma2 - 2.22
rotor inlet relative Mach number Mar2 - 1.62
stator inlet absolute Mach number Ma1 - 0.56
stator outlet absolute Mach number Ma2* - 1.19
rotor inlet relative Mach number Mar2* - 0.68
expected shaft power kW 38
predicted total to static isentropic efficiency 53.6
%
(1D loss model) ηts,1D

4
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

2.2 CFD Physical models and boundary conditions

To carry out these analysis and turbine optimizations, appropriate CFD software has to be
chosen. For that reason, the software Fine/Turbo by NUMECA [22] which the authors
already used for previous investigations [23,24], was used for the simulations. For the rotor
stator interface, a full non-matching frozen rotor was used. As a turbulence model, the
Spalart-Allmaras model was applied. The turbine has to be meshed 360° because due to the
partial admission, it is not possible to simulate only a segment of the whole turbine. Table 3
shows the boundary conditions and physical models applied to the CFD simulations.

Table 3. Physical Models and Boundary Conditions of the CFD Simulations


Model or Condition Parameter
Mathematical Model Turbulent Navier Stokes
Modelling of turbulence Spalart-Allmaras
Rotor-Stator Interface Full Non-Matching Frozen Rotor
Efficiency Definition Total to Static Isentropic
Fluid Model Water (Condensable Fluid)
Inlet Boundary Condition Absolute Total Pressure (17.5 bar)
Outlet Boundary Condition Static Pressure (0.2 bar)

2.3 Mesh study


To assess and estimate the quality of the mesh, different grid accuracies were analysed and
compared to each other. Figure 1 shows the inlet and outlet mass flow rate and the total-to-
static isentropic efficiency over a varying number of grid points for the design boundary
conditions. It becomes clear that for a very low number of grid points (0.7 Million), the mass
flow rate extremely differs from the design mass flow rate (82 g/s). The mass flow rate course
flattens out at about 81 – 82 g/s at 6.1 up to 53.8 Million grid points. The efficiency varies
between 61.54 % and 61.20 % in the case of 53.8 to 6.1 Million grid points, respectively. The
efficiency value for 0.7 million grid points is not conclusive due to the high mass flow rate
difference. The mesh with 48 Million grid points was used as a compromise between
accuracy and computing time. The still non-negligible deviation from inlet to outlet mass
flow rate is very likely to be caused by high unsteady effects in the flow field due to partial
admission. An unsteady time averaged CFD simulation confirmed these assumptions since a
smaller deviation could be achieved (82 g/s at the inlet and 82.8 g/s at the outlet).

5
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

Fig. 1. Mass flow rate and total to static isentropic efficiency as a function of the number of grid
points

2.4 Parameterized turbine design


In Figure 2 (top), the axial pressure distributions through the whole turbine at specific axial
positions for different versions of the Curtis meridional flow channel are shown. The dotted
line with the squared markers (1D-Design (15%)) and the dotted line with the triangular
markers (1D-Design (20%)) show the static pressure evolution designed by means of the
1DTDT. The solid line with squared markers (CFD V1 (15%)) and the solid line with the
triangular markers (CFD V16 (20%)) represent the pressure distributions achieved by 3D-
CFD simulations. It has to be addressed that in the parametrization process, the reaction in
the stator wheel was up scaled from 15 % to 20 % – the versions in Fig. 2 are named
accordingly. If the degree of reaction is at 15%, it implies that 85 % of the entire turbine
isentropic enthalpy drop is converted into kinetic energy in the supersonic nozzles and 15 %
in the stator wheel – 20 % reaction follows accordingly. Both wheel passes are originally
designed for constant pressure flow in the 1DTDT. The predicted blade heights (1DTDT)
should ensure the aimed pressure distribution. However, it becomes clear that there is a large
deviation between the designed (1D-Design (15%)) pressure distribution and the achieved
(CFD V1 (15%)) pressure distribution in 3D-CFD (Fig. 2). The biggest differences occur
between the inlet of the first rotor wheel and the stator wheel outlet. In contrast to the constant
pressure (impulse) design, the pressure drops from about 0.5 to 0.2 bar over the first rotor
wheel in V1. This pressure drop leads to a high reaction in the first wheel and increasing
relative velocity over the blade channel. After this pressure drop over the first rotor wheel,
the static pressure remains at the outlet conditions till the outlet area.

6
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

Fig. 2. Static pressure (top) and half blade height (bottom) at different positions through the turbine
(mean diameter Dm = const.)

The origin of the significant deviation between 1D-design and 3D-CFD pressure distributions
is the applied 1D loss model. The 1D loss model is not only needed to predict turbine
efficiency. Furthermore, it is needed to determine the fluid properties like density at any axial
position in the turbine. Thus, too high losses result in too low fluid density and thereby too
big flow areas or blade heights, respectively. The loss model of the 1DTDT has mainly been
developed and applied previously for single wheel axial and radial impulse turbines, working
with organic fluids [21]. So far, deviations this big between 1DTDT and 3D-CFD have not
occurred yet. Therefore, the authors decided to continue the optimization of the Curtis turbine
directly in 3D CAD and 3D CFD, applying a parameterized design approach.
By adjusting the meridional flow channel or the blade heights (Fig. 2 bottom) step by
step, a desired axial pressure distribution (CFD V16, Fig. 2 top) could be achieved. The
prismatic blading remained mainly unchanged – besides rounding the leading edges of the
stator and second rotor wheel. As mentioned above, reaction or enthalpy drop across the

7
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

stator wheel has been increased from 15 % to 20%. The V16 stator wheel is working as a
transonic nozzle. Furthermore, additionally, a slight reaction has been introduced over the
two rotor wheel passes, although axial thrust forces should be avoided. This slight streamwise
pressure drop should ensure that there is always a small but clearly defined axial force in the
streamwise direction on the bearings. The necessary changes in blade heights are significant
for the stator wheel and the second rotor wheel but much smaller for the first rotor wheel.
This confirms the previous reasonable results for axial and radial impulse turbines [19]
designed by the 1DTDT. Total-to-static isentropic efficiency has been increased from 57.5%
(CFD V1) to 61.5% (CFD V16).
To achieve and to obtain the V16 design, it was necessary to model the turbine in a
parameterized way and to iteratively exchange the information between CAD and CFD. The
steps, which were necessary for this, are described below in more detail.

Fig. 3. Comparison of the meridional channel in CAD and CFD

Figure 3 shows the meridional flow channel of the Curtis turbine in the CAD software Creo
(top) [25] and in IGG (Interactive Geometry Generator) (bottom) [22], the program to prepare
the geometry and to generate the mesh for the CFD simulation. In Fig. 3 (top), the meridional
channel in blue and the blades in grey can be seen. The red crosses mark the points where the
channel can be adjusted. The height of the turbine outlet changes accordingly to the height at
the outlet of the last rotor wheel. The inlet of the first rotor wheel is fixed due to a fixed
nozzle outlet height. This simplification was necessary because the expenditure of time for
manually meshing a new nozzle geometry was too high. The bottom part of Fig. 3 shows the
meridional channel in blue and the blades in grey in the IGG environment. In comparison to
the channel and blades at the top of the picture, only the blade heights differ because the
meshing process implies prismatic blade geometries, which stand out over the hub and shroud
curve of the meridional flow channel. The channel in IGG is taken directly from the CAD
software via export/import. In this way, it is possible to adapt the channel in CAD and
combine it with the existing blades in IGG. This implementation allows quick geometry
adjustments, mesh generation and thereby a quick gain in information. In this way, it was
possible, that the geometry, in this case the meridional flow channel, could be changed. The
meshing process ran afterwards and the simulation was then started. After the analysis and
evaluation of the results received from this CFD simulation, the authors could then adjust the
geometry and start the entire process again. The CFD simulation needed about 8 h so that the
overall turn-around-time for one geometry modification was about one working day
(including the night for CFD calcs).

8
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

3 Results
Due to a high expenditure of time, not all of the different versions of the turbine could be
analysed in detail. Because of this reason, only the most prospective version of the turbine is
analysed in the following.

4 CFD Contour plots


Figure 4 shows the absolute (top) and relative (bottom) Mach number distribution for the
V16 at 36,000 rpm and the design pressure ratio at 50 % blade height. The nozzles accelerate
continuously as designed, with little anisotropic flow at the outlet to about Mach 2.5 to Mach
3. The flow passes through the first rotor wheel with shockwaves, probably induced by the
leading edges of the rotor blades. Flow separation occurs at the suction side of some of the
blades, and the flow stays attached at the pressure side. For the stator wheel, the flow enters
the channel with sub- to transonic flow and accelerates steadily to about Mach 1.5 to Mach
2 in the absolute reference frame. The flow then passes through the second rotor wheel and
exits the turbine through the outlet with about Ma 0.7. Surely, this flow field is representing
one specific position of the frozen rotor-stator interface. The results should not be overrated,
but nevertheless, the chosen approach is reliable and quick enough to be applied in a
reasonable turbine design process.

Fig. 4. Absolute (top) and Relative (bottom) Mach number plots for V16 at 36,000 rpm and design
pressure ratio (87.5)

9
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

4.1 CFD Performance maps

The designed Curtis turbine will not constantly work under design conditions in the WHR
plant. Condensing pressure and thereby overall turbine pressure ratio (PR) will vary, e.g. due
to seasons. In order to diminish efficiency losses due to varying PR, the rotational speed of
the turbine is usually adapted. Therefore, efficiency characteristics of V16 are generated by
means of the described CFD approach.
The total to static isentropic efficiency as a function of the rotational speed of the V16 is
shown in Figure 5. The maximum rotational speed of 36,000 rpm at the design point was
chosen due to mechanical integrity limitations. Generally, the efficiency shows a
comprehensible behavior over the entire rotational speed range. The efficiency of the turbine
increases with rotational speeds, 63 % are achieved at the highest considered rpm of 39,000,
which are not allowed in actual operation (mechanical integrity). The entire course lies on a
too high efficiency level from approx. 53 % to 62 %, because not all physics are considered.
Although the 360° CFD simulation takes into account steam leakages at the beginning and
the end of the impinged part of the rotor wheels, unsteady effects are not considered.
Furthermore, the tip gaps of the rotor wheels and the hub gap of the stator wheel are not
modelled – as can be seen in Fig. 3 - due to CFD performance reasons.

Fig. 5. Total to static isentropic efficiency as a function of the rotational speed for the V16

Figure 6 depicts the total to static isentropic efficiency as a function of the pressure ratio (PR)
for V16. The pressure ratio was adjusted by changing the static pressure at the outlet of the
turbine (i.e. condensing pressure, displayed at every point in the figure); the inlet pressure
stayed constant at 17.5 bar. From a lower to a higher PR, the efficiency increases up to the
design PR of 87.5 (17.5 absolute total inlet pressure and 0.2 bar static outlet pressure) and
decreases from this point to a PR of 175. For pressure ratios lower than approx. PR = 29, the
simulation did not converge because of the high unsteady flow effects and flow separations.
For these high outlet pressures, the second rotor wheel even increased the total enthalpy and
therefore worked as a compressor instead of a turbine. The efficiency drop for higher than
designed pressure ratios can be explained by the inability of the Laval nozzles to convert the
available higher enthalpy drop (PR > PRdesign) into kinetic energy. As expected, the turbine
efficiency shows a dependency on the pressure ratio but still performs well in the outlet

10
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

pressure range between 0.15 bar and 0.3 bar. The highest efficiency at the designed PR
reinforces the confidence of the authors in their design.

Fig. 6. Total to static isentropic efficiency as a function of the pressure ratio for the V16 with outlet
pressures (bar) for every point

5 Conclusion
The physical challenges of designing a small-scale steam turbine were repeated. Thereby, it
has become clear why a two wheel velocity compounded axial Curtis turbine must be chosen
for the given design task. The applied design approach, consisting of an in-house developed
1D turbine design tool and a parameterized 3D CAD/CFD design, was introduced. The
shortcomings of the applied 1D loss model and their consequences for the turbine geometry
were addressed.
The predicted efficiency (1DTDT) of the V1 was just 53.6% compared to 57.5% in 3D
CFD. By adjusting and optimizing the meridional flow channel directly in 3D space, the
desired axial pressure distribution could be achieved with V16. Total to static isentropic
efficiency is finally 61.5% in 3D-CFD. Thus, the parametric design, directly in 3D space,
proved to be a practical and reasonable process for daily turbine development. Mach number
contours at design point showed the expected pattern but furthermore, area for improvements.
Finally, performance characteristics of the final design (V16) were introduced and evaluated.
The designed axial two-wheel Curtis is currently under construction. For the new design,
Titanium for the rotor can be avoided and thereby material and manufacturing costs reduced.
The new turbine will be installed 2022 in the WHR plant of the main sewage treatment plant
in Nuremberg/Germany. There, the waste heat of two 1 MW biogas engines is converted to
40 -50 kWel by means of a micro steam Rankine cycle.

The authors gratefully acknowledge the funding of the joint project KompACT on the techno-
economic optimization of an SRC plant for high temperature WHR by the Federal Ministry for
Economic Affairs and Energy in the 6th Energy Research Program on the basis of the decision by the
German Bundestag.

11
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

References
1. Pehnt DM, Bödeker J, Arens M, Jochem DE, Idrissova F. Die Nutzung industrieller
Abwärme – technisch-wirtschaftliche Potenziale und energiepolitische Umsetzung:
Bericht im Rahmen des Vorhabens „Wissenschaftliche Begleitforschung zu
übergreifenden technischen, ökologischen, ökonomischen und strategischen Aspekten
des nationalen Teils der Klimaschutzinitiative“. Heidlberg, Karlsruhe; 2010.
2. Macchi E, Astolfi M (eds.). Organic rankine cycle (ORC) power systems:
Technologies and applications. Duxford, UK, Amsterdam, Boston, Heidelberg,
London, New York, Oxford, Paris, San Diego, San Francisco, Singapore, Sydney,
Tokyo: Woodhead Publishing is an imprint of Elsevier; Elsevier; 2017.
3. Tanuma T. Advances in steam turbines for modern power plants. Oxford: Woodhead
Publishing; 2016.
4. Raab Florian, Opferkuch Frank, Klein Harald. Dezentrale Verstromung von Abwärme:
Forschungsarbeit zur Abwärmenutzung mittels Steam-Rankine-Cycle-Technlogie. In:
BWK Energie.
5. Kraus MH, Deichsel M, Hirsch P, Opferkuch F, Heckel C. Hermetic 40-kW-Class
Steam Turbine System for the Bottoming Cycle of Internal Combustion Engines. In:
ASME Turbo Expo 2016: Turbomachinery Technical Conference and Exposition, June
13-17, 2016, Seoul, South Korea. New York, N.Y.: ASME; 2016.
6. REFPROP: NIST Reference Fluid Thermodynamic and Transport Properties Database
(REFPROP). [March 30, 2021]; Available from: https://www.nist.gov/srd/refprop.
7. Seume JR, Peters M, Kunte H. Design and test of a 10kW ORC supersonic turbine
generator. J. Phys.: Conf. Ser. 2017;821:12023. https://doi.org/10.1088/1742-
6596/821/1/012023.
8. Klonowicz P, Witanowski Ł, Suchocki T, Jędrzejewski Ł, Lampart P. Selection of
optimum degree of partial admission in a laboratory organic vapour microturbine.
Energy Conversion and Management 2019;202:112189.
https://doi.org/10.1016/j.enconman.2019.112189.
9. Uusitalo A, Honkatukia J, Backman J, Nyyssönen S. Experimental study on charge air
heat utilization of large-scale reciprocating engines by means of Organic Rankine
Cycle. Applied Thermal Engineering 2015;89:209–19.
https://doi.org/10.1016/j.applthermaleng.2015.06.009.
10. Żywica G, Kaczmarczyk TZ, Breńkacz Ł, Bogulicz M, Andrearczyk A, Bagiński P.
Investigation of dynamic properties of the microturbine with a maximum rotational
speed of 120 krpm – predictions and experimental tests. J VIBROENG
2020;22(2):298–312. https://doi.org/10.21595/jve.2019.20816.
11. Meher-Homji CB. The Historical Evolution Of Turbomachinery 2000:281–322.
https://doi.org/10.21423/R1X948.
12. Stodola A. Die Dampfturbinen: Mit einem Anhang über die Aussichten der
Wärmekraftmaschinen und über die Gasturbine. 1st ed. Berlin: Springer; 2012.
13. De Laval CGP. Steam Turbine: Patent(522,066); 1889.
14. Weiß AP. Volumetric expander versus turbine - which is the better choice for
small ORC plants? In: University of Liège and Ghent University, editor. ASME-
ORC2015: Proceedings of the 3rd International Seminar on ORC Power Systems;
2015.

12
MATEC Web of Conferences 345, 00031 (2021) https://doi.org/10.1051/matecconf/202134500031
Power System Engineering 2021

15. Harris FR. The Parsons Centenary—a Hundred Years of Steam Turbines. Proceedings
of the Institution of Mechanical Engineers, Part A: Power and Process Engineering
1984;198(3):183–224. https://doi.org/10.1243/PIME_PROC_1984_198_024_02.
16. OMERSCALES EFC. The Vertical Curtis Steam Turbine. Transactions of the
Newcomen Society 1990;62(1):157–8. https://doi.org/10.1179/tns.1990.008.
17. Spadacini C, Rizzi D. Radial outflow turbines for Organic Rankine Cycle expanders.
In: Macchi E, Astolfi M, editors. Organic rankine cycle (ORC) power systems:
Technologies and applications. Duxford, UK, Amsterdam, Boston, Heidelberg,
London, New York, Oxford, Paris, San Diego, San Francisco, Singapore, Sydney,
Tokyo: Woodhead Publishing is an imprint of Elsevier; Elsevier; 2017, p. 335–359.
18. Weiß AP, Popp T, Zinn G, Preißinger M, Brüggemann D. A micro-turbine-generator-
construction-kit (MTG-c-kit) for small-scale waste heat recovery ORC-Plants. Energy
2019;181:51–5. https://doi.org/10.1016/j.energy.2019.05.135.
19. Weiß AP, Popp T, Müller J, Hauer J, Brüggemann D, Preißinger M. Experimental
characterization and comparison of an axial and a cantilever micro-turbine for small-
scale Organic Rankine Cycle. Applied Thermal Engineering 2018;140:235–44.
https://doi.org/10.1016/j.applthermaleng.2018.05.033.
20. Weiß AP, Novotný V, Popp T, Zinn G, Kolovratník M. Customized Small-Scale ORC
Turbogenerators- Combining a 1D-Design Tool, a Micro-Turbine-Generator-
construction-kit and Potentials of 3D-Printing. In: The National Technical University
of Athens, editor. ORC2019: Proceedings of the 5th International Seminar on ORC
Power Systems; 2019.
21. Weiß AP, Novotný V, Popp T, Streit P, Špale J, Zinn G et al. Customized ORC micro
turbo-expanders - From 1D design to modular construction kit and prospects of
additive manufacturing. Energy 2020;209:118407.
https://doi.org/10.1016/j.energy.2020.118407.
22. NUMECA International - FINE™/Turbo. [March 30, 2021]; Available from:
https://www.numeca.com/en_eu/product/fineturbo.
23. Streit P, Popp T, Weiß AP. Simulation and analysis of the performance map of a
micro-ORC-turbine - Comparison with measurements. In: 18th Conference on Power
System Engineering, Thermodynamics & Fluid Flow: PSE 2019.
24. Streit P, Popp T, Winkler J, Scharf R, Weiβ AP. Numerical and experimental
investigation of different technologies for adjusting the swallowing capacity of a
cantilever ORC turbine. In: THERMOPHYSICAL BASIS OF ENERGY
TECHNOLOGIES (TBET 2020). AIP Publishing; 2021, p. 70001.
25. PTC Creo Parametric. [March 30, 2021]; Available from:
https://www.pdsvision.com/de/ptc-creo.

13
View publication stats

You might also like