You are on page 1of 24

AIAA 2015-4216

Propulsion and Energy Forum


July 27-29, 2015, Orlando, FL
51st AIAA/SAE/ASEE Joint Propulsion Conference

Design and Analysis of a High-Pressure Turbopump at


Purdue University

David Stechmann1, Charles Sese1, Rishika Duvvur1, Yu Huang1, Michael Bilyeu2, Daniel Goldberg2, Michael
King2, Allen Zhang2, and Timothee Pourpoint3
Purdue University, West Lafayette, IN, 47907

Most high-performance rocket engines employed on launch vehicles today use turbopumps to
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

supply propellants to the combustion chamber at high volume flow rates and high pressures.
While extensive experience exists in the rocket propulsion industry with these devices, relatively
little experience exists in an academic setting. As such, Purdue University has embarked on an
effort to design and built a high-performance turbopump capable of using a high-pressure gas
generator to pump a storable oxidizer at a nominal flow rate of 170 GPM to a discharge
pressure of 1400 psi. The overall configuration consists of a single-stage partial-admission axial
impulse turbine driving a single-stage centrifugal pump and inducer assembly using a single
shaft. The configuration will initially be tested using water and a high-flow hot-air source at
Purdue’s Zucrow Laboratories. During development, extensive analysis and design attention
was devoted to fluid dynamic simulation and cavitation suppression in the inducer and impeller,
structural stress analysis and safety factors in all fixed and rotating components, overall system
performance and power matching, case clearances, bearing load balancing, rotordynamic
analysis, dynamic seal placement and leakage estimation, and manufacturing considerations.
While the turbopump has yet to be tested, numerous lessons were learned during the
development process. All this point, all critical design details have converged and the system is
capable of meeting all of the performance and operational requirements. Hardware
manufacturing is pending.

Nomenclature
ATM = Automatic Topology and Meshing
D = Diameter (in)
DMLS = Direct Metal Laser Sintering
Ds = Specific Diameter
GPM = Gallons Per Minute
MPL = Minimum Power Level
Ns = Specific Speed
P = Pressure (psia)
RPM = Revolutions Per Minute
RPL = Rated Power Level
RL = Redline
T = Temperature (˚F)
y+ = Non-dimensional wall distance

1
Graduate Student, Department of Aeronautics and Astronautics, Purdue University 701 W. Stadium Ave. West
Lafayette, IN 47907-2045.
2
Undergraduate Student, Department of Aeronautics and Astronautics, Purdue University 701 W. Stadium Ave.
West Lafayette, IN 47907-2045.
3
Associate Professor, Department of Aeronautics and Astronautics, 500 Allison Road West Lafayette, IN 47907.

1
American Institute of Aeronautics and Astronautics

Copyright © 2015 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
I. Introduction
Most high-thrust rocket engines used on launch vehicles today employ turbopumps.1 These pumps are
designed to increase the delivery pressure of propellants to the combustion chamber making it possible to
achieve high efficiency operation, compact engine designs, and lightweight propellant tanks. Turbopumps
are a practical solution for this type of application because they can provide very high specific power and
very high power density – traits that are highly beneficial in any rocket system. Despite the near ubiquity
on commercial rockets however, little experience exists in an academic setting with these types of systems.
Given Purdue’s storied history with propulsion technology and the current experience with rocket
combustors and rotating machinery for aircraft engines, it is only natural that Purdue should extend this
experience to turbopumps. Doing so will not only advance Purdue’s own scientific and intellectual capital,
it will enable development and future testing of this type of technology in an environment that is very
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

conducive toward learning, knowledge transfer, and innovation.

II. Design

Objectives
The primary objective of the project described in this paper was to develop a turbopump with operating
characteristics and performance capabilities similar to those of other high-performance rocket engine
turbopumps. At the same time, the turbopump in question would need to be compatible with the testing
facilities available at Purdue’s Zucrow Laboratories. To keep this one semester long class project simple,
water was the initial intended working fluid for the pump, and the turbine was designed to operate on hot
air supplied from an external air heater in place of a storable oxidizer and a gas generation system
respectively. Based on these general parameters, the following detailed objectives were as follows:
1. The pump needed to be capable of pumping water at a flow rate of 170 GPM and a pump discharge
pressure of 1400 psia at its nominal operating condition. Based on system optimization trade studies
described later, a shaft speed of 24,000 RPM was set as the nominal operating point.
2. The pump needed be capable of surviving brief periods (several minutes) of operation at its redline
condition. Based on the same system optimization studies and limitations associated with the
Zucrow test facilities, a shaft speed of 28,800 RPM was set as this redline condition. Additional
redline parameters based on pump and turbine affinity laws follow from this including a redline
discharge pressure of 2,000 psi. These are documented in more detail later.
3. The turbopump had to operate within the high-pressure air supply limits associated with Zucrow
laboratories included a maximum supply pressure of 600 psia at 900˚ F and a maximum mass flow
rate of 5 lbm/s through the turbine.
Secondary objectives for the turbopump were as follows:
1. The turbopump and all associated flow paths needed to include instrumentation capable of
accurately characterizing the turbine and pump performance curves. These curves would be based
on flow rates, pressures, and power input / output at various engine speeds and would be used to
build an engine performance and operating map.
2. The turbopump assembly needed to be modular such that changes to the turbine or pump could be
implemented without complete disassembly or test stand removal. This was desirable to facilitate
experimental upgrades or improvements during testing if necessary.
3. The turbopump needed to be designed so it could be easily modified at a later point to pump a
room-temperature storable oxidizer like hydrogen peroxide or a room-temperature fuel like RP-1.

Configuration and Layout


The overall layout of the final turbopump design and rotor configuration is shown in Figures 1, 2, and 3.
The general configuration consists of a single-stage partial-admission axial impulse turbine driving a single-
stage centrifugal pump assembly using a single shaft. This arrangement was selected for its simplicity. In

2
American Institute of Aeronautics and Astronautics
this arrangement, the centrifugal pump assembly pumps water at moderate flow rates to the required
discharge pressure while the turbine provides shaft power to the pump at the required rotation rate. The
turbine assembly produces this shaft power by accelerating hot high-pressure air from the Zucrow air-heater
system through supersonic (convergent – divergent) turbine nozzles before discharging tangentially into
blades located on the turbine disk. Aft of the turbine disk, the exhaust air (which is once again subsonic)
passes through a waffle-grid flow-straightener before entering an annular plenum leading to a choked exit
nozzle. The choked exit nozzle combined with the choked turbine nozzles makes it possible vary the turbine
power by changing manifold pressure without appreciably changing the pressure ratio across the turbine
nozzles. This simplifies design and analysis, and it means the turbine should be able to operate close to the
optimum design point over a wide range of supply pressures and temperatures.
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

Figure 1: An isometric view of the turbopump mounted on the test stand. Key components are noted. A
standard 12” ruler is provided in the lower left corner for scale.

The pump subassembly consists of a vortex suppressor to minimize up-stream vortex coupling from the
pump rotation, an axial inducer, a radial impeller and a pump case split into forward and back components
as shown in Figure 2. The axial inducer boosts the incoming flow pressure by approximately 10% of the
total pump capability to prevent cavitation in the main impeller. Low pressure or low-speed centrifugal
pumps often do not need an inducer, but the operating speeds and pressures of interest to propulsion
engineers make it difficult to achieve high-performance engines without an inducer or other type of boost
pump.1 After the flow passes out of the impeller radially at high pressure, it is routed to the pump discharge
orifice using a double-tongue volute design. This approach, while more challenging to design than a single-
tongue volute, is generally capable of minimizing radial loads on the pump and the bearings – especially at
off design conditions.2 For high-pressure pumps subjected to high loading, this can be a significant
advantage. The pump case is designed to withstand worst-case pressure loads and deflections while
maintaining sufficient clearance from the shaft, impeller, and inducer even with the expected shaft thermal
expansion during operation. The pump case also houses the front-face and back-face dynamic impeller shaft
seals to minimize leakage from the high-pressure impeller outlet back to the low-pressure inlet. These

3
American Institute of Aeronautics and Astronautics
dynamic seals are positioned based on the impeller pressure distribution, the inducer axial load, and the
turbine axial load to minimize net axial load on the entire rotor assembly.
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

Figure 2: A section view of the turbopump assembly without the test stand support cradle.

Figure 3: An isometric detail view of the rotor components including the inducer, impeller, shaft, and turbine.

The turbine subassembly consists of dual inlets into the air supply manifold, a windage shield supporting
four multi-vane stator inserts (turbine nozzles), the turbine itself, the turbine retaining shaft cap, the turbine
case, the turbine exit flow guide, and the annular exhaust nozzle as shown in Figure 2. The manifold inlets
are placed tangentially to the manifold and 180 degrees apart to provide a consistent pressure distribution
within the manifold and reduced manifold pressure loss within the available space. The turbine case and
stator vanes are located as close to the turbine as possible to minimize leakage around the turbine while
simultaneously maintaining clearance under worst-case thermal expansion and structural deflections. Both
the turbine subassembly and the pump subassembly are bolted to seal housings on either side of the shaft.
These seal housings each contain four sets of dynamic shaft seals, one purge gas inlet, two drain pathways,

4
American Institute of Aeronautics and Astronautics
and one lubrication inlet / outlet. These seals and the associated purge and drain pathways make it possible
to keep the hot air, bearing lubrication oil, and pump fluid separated despite small amounts of expected
leakage through the dynamic shaft seals. The seal housings also contain Belleville springs designed to
provide the appropriate level of axial pre-load to the bearings, and these components transfer axial bearing
loads to the central shaft housing and into the test stand.

Operating Specifications
In principle, a turbopump capable of meeting the aforementioned performance objectives could be
designed to rotate at a wide range of speed. There are trade-offs involved with different design points
however, so a non-linear iteration algorithm was used to determine the best design speed and other operating
and geometric parameters for this particular application. The principle intent of this iteration was to match
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

the required pump power with the delivered turbine power while accounting for the following effects,
design considerations, and performance or structural limitations:

1. Turbine geometric diameter limits. While large turbine diameters can result in improved efficiency
at low rotation speeds, they also result in larger turbine cases and manifolds. Since cost was a
concern, it was determined early in the development process that a turbine diameter of 6” or less
would likely help reduce the size and cost of the resulting components by a noticeable amount.
2. Turbine disk centrifugal stress limits. High-speed turbines with large diameters are subject to very
high centrifugal loading. This loading generally increases quadratically with diameter and speed.3
As such, the turbine diameter and speed were both limited to keep these stresses from becoming
excessive during the design iteration process.
3. Turbine specific speed and specific diameter relationship. Due to the nature of turbine flow-path
geometry, the turbine type (axial or radial), turbine operating efficiencies, required admission area,
and blade / diameter width ratio will all depend on the relationship between turbine specific speed
and specific diameter. Figure 4 provides a graphical schematic of the complex interrelation between
these different parameters. During development and power-balance analysis, an effort was made to
keep both the turbine specific speed and specific diameter such that reasonable efficiency could be
obtained. The red dot on Figure 4 also indicates the point at which the final turbine configuration
was designed to operate at nominal conditions.
4. Turbine isentropic velocity ratio. This parameter relates the turbine rotational velocity at the blade
locations with the turbine nozzle exit velocity. Depending on the type of turbine and the admission
area, this ratio also influences turbine efficiency. For a single-stage axial impulse turbine, the
optimum nozzle isentropic velocity ratio should be approximately 0.5. This can be seen in the left
side of Figure 5. While it was not possible to match this value in this case given the other design
constraints during turbopump development, the final operating ratio of 0.26 is still capable of
providing a turbine efficiency of ~60%. This can also be seen in the left portion of Figure 5.
5. Turbine nozzle pressure ratio. The nozzle pressure ratio will influence the amount of power the
turbine can deliver, but the pressure ratio must be set so the inlet pressure must never exceed the
limitations of the laboratory facilities. Similarly, the outlet pressure must never fall below the
pressure needed to sustain sonic conditions at the choked turbine exhaust exit (as noted earlier).
6. Shaft stress limits. The shaft must be designed to transmit all of the necessary power from the turbine
to the pump during operation. The power transfer capability of the shaft will be driven by the
rotational speed, shaft diameter, and the shaft material stress limits. To avoid shaft design challenges
later, conservative shaft stress limits were considered in the optimization process. These limits
subsequently dictated a minimum shaft diameter limit of about 1.0” for the different speed
configurations considered. It should be noted that the actual shaft ended up being somewhat larger
for rotordynamic considerations. This is detailed later.
7. Bearing speeds. In order to keep the design simple, the initial intention was to use commercial ball
bearings if possible. Unfortunately most commercial bearings have low speed capability relative to
the high-speeds needed in turbopumps. As such, bearings commonly used in high-speed machining

5
American Institute of Aeronautics and Astronautics
applications were selected since these generally have some of the highest speed ratings.
Nevertheless, at the shaft diameter needed to transmit the required pump torque (1 inch or greater),
commercial bearings with speed ratings above 30,000 to 40,000 RPM were still difficult to locate,
so it was necessary to keep the maximum speed below this threshold if possible.
8. Pump specific speed. The pump specific speed relates speed, volumetric flow rate, and head rise as
shown in the right portion of Figure 5. It can be seen from this that the pump specific speed has a
significant impact on overall pump efficiency. As such, a specific speed of 600 was considered a
minimum for this particular application. The red dot and line on the fight plot in the right side of
Figure 5 indicates the nominal pump specific speed and anticipated efficiency while operating at the
rated power level. It should be noted that this efficiency estimate was based entirely on past empirical
data and not on any computational fluid dynamic analysis (discussed later).
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

NsDs Turbine Chart
100
Specific Diameter (Ds)

10

0
0.001 0.01 0.1 1 10 100 1000 10000
Specific Speed (Ns)
Figure 4: Turbine specific speed (Ns) and specific diameter (Ds) relationships.4 The nominal operating design
point (rated power level) of the turbopump is marked in red.

Figure 5: Left – Empirical turbine efficiency as a function of nozzle isentropic velocity ratio – the ratio of
turbine nozzle exit velocity to blade tip velocity.2 Right – Empirical centrifugal pump correlations for specific
speed and efficiency.2 The nominal operating point (rated power level) is marked in red.

6
American Institute of Aeronautics and Astronautics
The net result of the iterative engine power balance analysis and the various design constraints
documented above is shown in Table 1. The operating specifications in this table illustrate the calculated
performance and running speeds at three different power levels: the minimum power level, the rated (or
nominal) power level, and the redline power level. The minimum power level is set by the minimum amount
of flow the primary high-pressure dome-regulator at Zucrow labs will be able to supply to the turbine (about
1.2 lbm/s) while still maintaining accurate control of the pressure. The redline condition represents the
highest flow rate that can be achieved with this regulated air system while running with an operating
temperature of 900˚ F. This temperature was noted earlier as a requirement, but it represents a fairly optimal
value for this particular application. It is high enough to provide reasonable turbine efficiency using a design
similar to other turbopumps, but it is low enough to avoid significant challenges associated with material
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

properties. In particular, this temperature makes it possible to avoid the use of Inconel or similar nickel
alloys in nearly all parts of the turbopump. This is a significant cost benefit.
It should be noted that the redline speeds and pressures shown in Table 1 represent the limit load
condition. As such, most of the FEA analysis used to validate the structural components was done using
boundary conditions derived from this operating condition. The original redline manifold pressure estimate
was set at 580 psi, but during the design process, the value needed was reduced to the 420 psi value shown
as more information was gained regarding the capabilities of the turbine nozzles from computational fluid
dynamic analysis. Nevertheless, the structural design continued to use the higher-pressure value since this
value still represented the upper capabilities of the air-supply system.

Table 1: Key operating parameters


Minimum Nominal
Power Power Redline
Shaft Speed (RPM) 15000 24000 28800
Pump Water Flow Rate (GPM) 106 170 204
Pump Discharge Pressure (psia) 555 1400 2010
Pump Specific Speed 692 692 692
Pump Tip Speed (ft/s) 281 450 541
Predicted Pump Efficiency (%) 57 62 64
Turbine Manifold Inlet Pressure (psia) 103 260 420
Turbine Outlet Pressure (psia) 23 59 92
Turbine Inlet Temp (˚F) 890 890 890
Turbine Mass Flow Rate (lbm/s) 1.2 3.0 5.0
Turbine Diameter (in) 5.9 5.9 5.9
Turbine Pressure Ratio 4.12 4.2 4.2
Turbine Exit Temperature (˚F) 435 433 400
Turbine Power Output (hp) 74 278 483
Turbine Specific Speed 10.5 16.7 20.2
Turbine Specific Diameter 2.9 2.9 2.9
Turbine Nozzle Exit Velocity (ft/s) 2325 2332 2319
Turbine Tip Speed (ft/s) 377 600 722
Turbine Nozzle Isentropic Velocity Ratio 0.16 0.26 0.31
Predicted Turbine Efficiency (%) 40 62 65
Shaft Torque (in-lbf) 310 726 1055

7
American Institute of Aeronautics and Astronautics
II. Detailed Fluid Dynamic Analysis
Detailed analyses of all parts within the assembly were conducted to ensure the design criteria and
operating parameters specified earlier could be met. Computational Fluid Dynamics (CFD) represented one
of the techniques employed. In particular, CFD methods and software tools were used to evaluate shaft
power requirements (or capabilities), pressure distributions, and overall pressure changes associated with
the inducer, the impeller, and the turbine. The cavitation performance of the inducer and the impeller were
also evaluated using these techniques. It should be noted that deflection data from structural analysis
(described later) was also used in the CFD analysis to ensure proper accounting of pressure losses due to
blade tip clearances. Given the small nature of this turbopump, these losses were non-trivial – especially in
the inducer and impeller. The methods associated with this CFD analysis and the results from the final
design iteration are described below.
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

Inducer Contour Design and Analysis Methodology


As explained earlier, the shaft speed selected to obtain the required pressure rise was 24,000 RPM at the
rated power level. At these speeds, a major concern in the design of the impeller was cavitation. The
presence of cavitation will result in a loss of pump head and lower efficiency. In order to avoid cavitation
in the impeller, the inducer was designed to increase pressure at the impeller inlet. In general, the maximum
cavitation head loss that should be allowed on an inducer is 3%, so this requirement was set here as well.2
In addition to the cavitation requirement, an inlet pressure of 15 psi was specified so that the pump would
be capable of running off of tank head alone without additional pressurization. Based on prior turbopump
configurations, the inducer was designed to give a head rise of 10% of the total head rise required to meet
the required performances.1 Additional parameters governing the inducer design are shown in Table 3.
In order to generate the Table 3: Inducer design parameters
necessary inducer contour, the
Impeller cavitation loss < 3%
commercial software CFturbo was
5 Inlet pressure 15 psia
used. This software was capable of
generating many different types of Shaft Speed 24,000 RPM
geometry for subsequent use with Shaft diameter 0.75 in
computational fluid dynamics Inlet diameter 0.75 in
analysis tools, finite element Outlet hub diameter Equal to Impeller inlet hub diameter
structural analysis tools, and Tip diameter Equal Impeller inlet tip diameter
computer-aided design tools. The Tip clearance 5% of blade height
parameters used by CFturbo to
generate the inducer design are 2: CFTurbo inducer geometric input parameters
listed in 2. After numerous
Hub diameter 0.75 in
iterations (many of which initial
Suction diameter 2.2 in
performed poorly during CFD
Inducer diameter 1.837 in
analysis), a highly capable inducer
was eventually created. The profile Outlet width 0.363 in
associated with this configuration is Length 1.5 in
shown in Figure 6. Number of blades 3
Using CFturbo it was possible to Blade thickness Hub: LE : 0.039 in TE : 0.049 in
export the geometry of the inducer Shroud: LE : 0.047 in TE : 0.059 in
o
directly in the form of curve for use Wrap angle 350.2
with Turbogrid meshing software. Leading Edge definition Ellipse : 6.00
Turbogrid subsequently made it Trailing Edge definition Ellipse : 1.00
possible to generate a structured
grid with an ATM optimized topology. This is a combination of O/H/J grid topology. In this case, there
was no need for control point adjustment. Since cavitation involves small air bubble close to the wall, a fine
mesh resolution was required. A y+ below unity is preferable in this situation to avoid the use of a wall

8
American Institute of Aeronautics and Astronautics
function which would reduce the accuracy associated with
computing the boundary layer. For this analysis, the y+ was kept
around 10 to accurately capture the cavitation while avoiding
high computational cost. It was also necessary to include an inlet
and outlet domain to avoid reverse flow.
During meshing and early analysis, a grid sensitivity analysis
was completed. Based on this analysis, it was apparent that by
choosing close to 1.9 million cells, the percentage of error could
be kept below 1%. Moreover, due to time constraints it was
preferable to have a smaller number of cells in order to iterate
on the design more quickly. It should be noted that the grid was
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

generated using only one flow passage, and periodic boundary


conditions were applied on the sides of the flow passage. In this
way, it was possible to model the full inducer while minimizing
computational requirements. Figure 7 shows the inducer grid Figure 6: Inducer geometry generated
by CFturbo
generated by Turbogrid and the single passage analysis
configuration.

Figure 7: Inducer grid generated using Turbogrid 14.5 (left and center), and the single-inducer flow passage
configuration used for analysis (right).

Once generated, the flow passage was imported into CFX 14.5 as a rotating domain. The two-equation
k-ω Shear Stress Transport turbulence model was employed for this analysis. Despite the fine grid at the
surfaces, the boundary layer was modeled using wall functions as y+ > 10. In order to model the cavitation
in the inducer, the Rayleigh-Plesset cavitation model was used. The saturation pressure was set to 0.51 psia
at a temperature of 60˚ F to mimic the properties of water, and the computation was done in two steps. First
the cavitation model was disabled so it would be possible to obtain the value of the total pressure rise
without cavitation. The second step was to re-enable the cavitation model. By evaluating the differences
between the results from these two models, it was possible to calculate the pressure loss due to cavitation.
During this step, the converged solution from the first step was used as an initial condition to facilitate
faster convergence. It should be noted that it was necessary to define two fluids for the analysis: water as a
liquid and water as an ideal gas.
At the inlet, a stationary-frame static pressure was prescribed normal to the inlet. The mass flow rate
was prescribed at the outlet while the blades and hub were set as no-slips rotating walls. As the blades are
unshrouded on the inducer, the shroud was defined as a no-slip counter-rotating wall. Figure 8 shows the
single passage boundary configuration in CFX. Table 4 provides a few additional parameters used in the
simulation. After setup, solving the total energy equation (with compressibility effects considered)
completed the simulations.

9
American Institute of Aeronautics and Astronautics
After several inducer design and
analysis iterations using the
aforementioned methodology, a
design that met all of the
requirements was found. In this
design, the mass fraction of vapor
was below 1% compare to the mass
fraction of liquid water. The loss of
pressure head due to cavitation was
below 3%, and a total pressure
increase of 137.8 psia could be
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

obtained. Figure 9 shows the Figure 8: CFX 14.5 computational setup for the inducer
pressure distribution and the
Table 4: Summary of the inducer computational setup
cavitation region on the inducer
obtained from these simulations. The Rotational speed 24,000 RPM
red isosurfaces show the regions of Fluid Water: Liquid / Vapor
cavitation. They are primarily Inlet stagnation pressure 15 psia
situated at the leading edge blade Inlet swirl angle 0 degrees
tips. There are also some regions of Outlet mass flow rate 7.854 lbm/s
cavitation near the hub, but these
areas are much more minor. Overall, the blade pressure distribution computed using this analysis made it
possible to determine that the inducer would require 85 in-lbf of drive torque at 24,000 RPM.

Figure 9: Inducer CFD results showing the cavitation regions (in red) and the pressure distribution on the
final inducer design.

Impeller Fluid Design and Analysis Methodology


Following the inducer design and
Table 5: Impeller CFTurbo geometric parameters
analysis, impeller development
commenced in much the same way as Hub diameter 1.4752 in
before by using CFturbo. The input Suction diameter 2.2 in
CFTurbo parameters for the impeller Inducer diameter 4.6 in
are shown in Table 5. In this case, both Outlet width 0.145 in
shrouded and unshrouded impellers Length 0.55 in
were considered. Initially a shrouded Number of blades 6
design was favored due to the Blade thickness Hub: LE : 0.06 in TE : 0.071 in
performance penalties associated with Shroud: LE : 0.063 in TE : 0.071 in
tip leakage flow in an unshrouded Wrap angle 136.6o
impeller. Due to manufacturing Leading Edge definition Ellipse : 6.00
challenges however, an unshrouded Trailing Edge definition Ellipse : 1.00

10
American Institute of Aeronautics and Astronautics
design was considered more seriously for a period of
time. Eventually performance losses associated with
tip clearances in an unshrouded configuration (over
25% at this scale) drove the design back to a shrouded
impeller. Additionally, there were concerns regarding
rotordynamic coupling in an unshrouded impeller and
pump unsteadiness due to vortex shedding along the
blades – especially during startup. Since the pump
behavior is related to the gap distance but the gap
distance is difficult to measure after the pump is
assembled, there was also greater uncertainty in the
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

performance of an unshrouded pump. The geometry


of the final shrouded impeller iteration obtained from
CFturbo is shown in Figure 10. It should be noted that
the meridional contour was designed to have a
constant cross section along the impeller to minimize Figure 10: The final impeller geometry from
the likelihood of flow separation within the impeller CFturbo (shown without the shroud for clarity).
and to maintain as consistent a flow velocity as
possible between the inlet and the outlet.2
The same meshing and analysis methodology described earlier for the inducer was used for analyzing
the impeller. As before, the grid was generated for only one flow passage (two halves on either side of the
blade). Figure 11 shows one flow passage with the periodic mesh in yellow. Figure 12 shows the impeller
boundary conditions in CFX while further parameters for the simulation are shown in Table 6. It should be
noted that the velocity profile from the inducer outlet was used at the impeller inlet.

Figure 11: The impeller mesh generated using Turbogrid 14.5

Due to the pressure increase provided by the Table 6: Summary of the impeller computational setup
inducer, it was possible to avoid cavitation
entirely in the impeller. In general, the blade Rotational speed 24,000 RPM
loading on the impeller shows a continuous Fluid Water: Liquid / Vapor
increase in static pressure from the leading edge Inlet velocity profile From the inducer outlet
to the trailing edge due to the dynamic head Outlet mass flow rate 3.927 lbm/s
developed by the rotating pump impeller. The
pressure contours from this analysis are shown
in Figure 13.

11
American Institute of Aeronautics and Astronautics
Based on the impeller CFD analysis, a
torque of 350 in-lbf would be necessary
to drive the impeller at 24,000 RPM.
This amounts to a required power of 170
hp – equivalent to 93% efficiency. While
this efficiency is very high for a pump,
the computation does not take into
account back-face leakage flow losses,
volute pressure losses, or transient losses
coming from the interaction between the
volute and the impeller tip. These effects
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

must be taken into account to provide a


more realistic estimate of pump
efficiency. A rough estimate of these Figure 12: The CFX 14.5 impeller computational setup.
losses can typically bring the
performance down by 30% resulting in a total pump efficiency around 65 % - close to the value expected
from the empirical relations discussed earlier.6

Figure 13: CFD results on the impeller showing the pressure distribution on the 3D geometry and along the
meridional contour. The shroud is removed in this view for clarity.

Turbine Fluid Design and Analysis Methodology


The turbine in this turbopump was designed to use a multi-vane partial admission convergent-divergent
nozzle assembly coupled with symmetrical turbine rotor blades. The initial turbine analysis was
accomplished using a simple 1D compressible flow model and velocity triangles. The results of this 1D
analysis can be seen in the general table of operating parameters shown earlier. This approach was
employed because the turbine performance was assumed to be far more linear than the impeller or inducer
(due to cavitation challenges in these other components). This approach also made it possible to iterate the
turbine design much more quickly since the geometric complexity made iteration using full CFD analysis
very time consuming. Nevertheless, once the design was frozen, CFD was employed to evaluate the
accuracy of the 1D model and the velocity triangle analysis approach. For simplicity, only a 90-degree
section of the turbine (which includes one of the four nozzles inserts) was studied using this method. The

12
American Institute of Aeronautics and Astronautics
actual turbine rotor has 79 blades, but for convenience 80 blades were assumed so each 90-degree section
of the turbine rotor could drive 20 blades in the CFD simulation. The projected impact of this simplification
was considered negligible.
The stator side turbine CFD model was composed of 9 flow passages for each nozzle insert. The meshing
was done with ICEM CFD 14.5. An unstructured mesh was developed in four steps: First, a surface mesh
was generated using the Octree algorithm from the CAD geometry. Next, the surface mesh was smoothed
until an appropriate quality was obtained. Following this, a volume mesh was grown on the surface mesh
using a Delaunay algorithm. Finally, a smoothing algorithm was employed to increase the quality of the
mesh. The final mesh consisted of 8.2 million cell for the rotor side and 4.3 million for the stator side. This
mesh is shown in Figure 14 for the rotor analysis segment and the stator analysis segment.
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

Figure 14: The fully unstructured grid generated using ICEM CFD 14.5 for the turbine rotor segment (left)
and the stator vane segment (right).

In general, a hybrid grid would be better suited to compute the behavior of the boundary layer in this
application, but due to computational cost, layers of prisms were not included. In this computation, only a
general estimation of the performance of the turbine was expected, so no study of the complex flow in the
turbine was required (including transient behavior). As such, the fully unstructured mesh was implemented
with a steady-state solution sequence to provide reasonable performance predictions. The frozen rotor was
also used as the frame-mixing model. In this model, the frame of reference is changed but the relative
orientation of the components across is fixed. It should be noted that the final model produced a steady state
solution to the multiple frame of reference problem with some account of the interaction between the two
frames. However, the transient effects and the transient losses at the frame change interface were not
modeled.
Once the turbine and
stator mesh was
completely assembled,
the simulation was run at
24,000 RPM using the
two-equation k-ω SST
turbulence model and
wall functions to model
the boundary layer. The
flow at the leading edge
of the stator was defined
as entering tangentially
since the turbine
manifold to the nozzle
insert was delivering it Figure 15: The CFX 14.5 turbine computational setup
in that orientation. It was

13
American Institute of Aeronautics and Astronautics
not possible to set a boundary tangentially to a surface, so it was necessary to model part of the inlet domain.
Finally, we decided to model the side of the turbine as free slip walls in order to avoid increasing the error
if no slip walls were set up. The model of these boundary conditions is shown in Figure 15.
It should be noted that one additional problem encountered with the turbine CFD model was the
periodicity of the rotor side. Since this was a partial admission turbine, the stator admission was only located
at four positions around the turbine rotor. As such, it was not possible to set periodic boundary conditions
on the stator or the rotor. If periodic boundary conditions had been used, it would have effectively modeled
a full-admission turbine and provided incorrect results. A small amount of error in the pressure and velocity
distributions resulted from this assumption, but this error was considered relatively minor overall.
Compared with a simple 1D analysis using velocity triangles, the turbine CFD analysis predicted a
nozzle exit Mach number about 40% greater than expected and a mass flow of 39% greater than expected.
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

Some of this error in the 1D analysis method can be attributed to the simplified averaging process used to
compute choked flow area and nozzle area ratios on a curved and twisting domain surface. Nevertheless,
the CFD analysis computed a turbine efficiency of 62%, which was very close to the predicted value. Given
these results, it was possible to simply reduce the turbine manifold pressures from the original design point
and still obtain the same overall system performance capability. This adjustment was already accounted for
in Table 1.
A graphical view of the Turbine CFD results is shown in Figure 16. It should be noted that it is possible
to see the acceleration of the flow on the suction side of the turbine blades. These acceleration regions
contribute to the overall increase in performance of the turbine beyond that predicted using the simpler
model. It is possible to see some asymmetry in the turbine flow, but this is not realistic. This asymmetry is
likely due to the fact that periodic boundary conditions could not be used despite the fact that the simulation
only includes part of the turbine disk. It should also be noted that this analysis does not account for leakage
flow around the rotor disk, so real-world turbine performance may be lower than these results. Given the
ample margin on manifold pressure available however and the conservative design, it should be possible to
simply increase the supply pressure back to the original target value if necessary to compensate for some
of these losses.

Figure 16: Turbine CFD results showing the Mach number and pressure distribution

14
American Institute of Aeronautics and Astronautics
III. Detailed Structural Analysis
In addition to computational fluid dynamic analysis, it was also necessary to evaluate the structural
stresses, design margins, and overall deflections of the fixed and rotating structural components using Finite
Element Analysis (FEA) techniques. The results associated with this analysis were used to improve
component designs to the point where material strength capability was consistently 50% above stresses at
the worst-case redline load conditions. In other words, all components ultimately were designed to have a
structural safety factor of greater than or equal to 1.5. The deflection results were also used to adjust casing
dimensions in an effort to provide sufficient clearance between the rotor and the stator during operation. It
should be noted that the pressure distributions associated with the CFD results were imported into the finite
element analysis software to provide accurate loading on the impeller, inducer, and turbine components.
Since the effects of pressure distribution and structural deflections were coupled, several design and
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

analysis iterations were required before the fluid and structural analysis results converged. The general
finite element analysis process and the application of this process to the inducer, impeller, shaft, and turbine
are described below. The finite element process and results associated with the fixed structural components
are not included in detail below in the interest of saving space. In general however, the fixed structure
design and analysis was not as challenging as the rotating structure, and because the static FEA analysis
techniques associated with the fixed structural components were consequently more typical of what would
be encountered in a standard commercial application.

Inducer Design and Structural Analysis


The inducer is manufactured using direct metal laser sintering of 15-5 PH stainless steel. In this state, it
will have a yield strength of nearly 170 ksi at room temperature at the desired age hardened condition.7 To
accommodate clocking features (designed to enable different inducer rotation angles relative to the impeller
during testing), splines are machined into the bottom face of the inducer as shown in Figure 17. The radial
splines transfer torque from the shaft through the impeller. To ensure that the inducer could withstand the
stresses expected in service, a finite-element analysis of this component was completed using a refined
adaptive mesh with 1.0 million elements. An initial coarse mesh analysis was used to determining regions
of high stresses prior to the mesh refinement. In particular, the mesh quality on the blades was refined by a
factor of 3 to capture high stress gradients in this region more accurately.
The inducer structural boundary conditions associated with this analysis included a compression support
along the side surfaces of the splines and the shaft contact surface to simulate contact with the impeller.
The hub edges were also constrained axially for the same reason but were free to rotate. A rotational velocity
of 28,800 RPM (redline) was applied along the z-axis in a counter-clock-wise direction to account for
centrifugal stresses, and finally the fluid pressure distribution obtained from the aforementioned CFD
analysis was applied along the surface of the blades. Since the water or storable oxidizer entering the inducer
will be at room temperature, the inducer was not subjected to any significant amount of thermal loading.
Based on the inducer finite element analysis, peak stresses in the inducer will result in a minimum safety
factor of 2.59. The highest stresses will be near the interface of the blades and hub as seen in the left side
of Figure 17, however the majority of the inducer will experience much lower stresses. The stress
concentrations at the blade roots are a result of very small fillets, however the minimum safety factor is still
greater than the required safety factor, and as a result the final inducer design meets all static strength
requirements. In general, worst-case inducer deflections are also small (less than 0.007”) as shown in the
right side of Figure 17. Case clearance in this area is greater than 0.01,” so no contact with the inducer case
should occur under load.

15
American Institute of Aeronautics and Astronautics
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

Figure 17: The inducer safety factor (left) and deflections (right)

Impeller Structural Analysis


The impeller is manufactured from 15-5 PH stainless steel much like the inducer, so it also has a yield
strength of 170 ksi at room temperature.7 The impeller structural analysis is an important factor that affects
the pump design as a significant amount of torque is applied to this component. Due to the large torque
loads, the central impeller hub employs a “polygon” shaft interface. This interface geometry helps minimize
stress concentrations that can occur in more conventional spline designs along the spline edges, and it has
been used in may other industrial applications to transfer high load using a small diameter shaft interface.8
The polygon-style hub interface is also more tolerant of manufacturing imperfections.8 For the impeller
FEA, a refined adaptive mesh consisting of 1.4 million elements was employed. As before, an initial coarse
mesh analysis helped determine regions of high stresses so the mesh could be refined in these areas.
The structural boundary conditions used on the impeller were similar to the inducer and consisted of
compression along the hub edges and shaft interface as well as a centrifugal load due to the redline rotation.
Hub inner surface and blade pressures were also applied based on the computed CFD pressure distribution.
Finally, the back-face pressure was assumed to vary as a function of radius according to the ideal couette
flow assumptions.2 Like the inducer, the impeller operates near room temperature and is constantly cooled
by water - hence there was no consideration of thermal loading.

Figure 18: The impeller back face (left) and front face (left) safety factor with the shroud removed for clarity.

16
American Institute of Aeronautics and Astronautics
Based on the impeller finite
element analysis, peak stresses will
result in a minimum safety factor of
3.9. The highest-stress areas will be
located at the edges of the back-face
protrusion for the dynamic labyrinth
seal as shown in Figure 18. High
stresses will also be present at the
hub contact interface and the hub
pressure-balance holes due to the
torque transfer in this region. Despite
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

the high stresses, the minimum


safety factor is still greater than the
required safety factor, so the impeller
meets all static strength
requirements. Worst-case
deformation is also fairly small at
0.008” (as shown in Figure 19) and it
is symmetrical along z-axis as Figure 19: The impeller deflections at redline load conditions.
expected. Radial case clearances are The outer shroud is removed for clarity.
over 0.031” near the impeller outer
edges, so no contact between the
impeller and the pump case should
occur under load.

Turbine Structural Analysis


The bladed turbine disk (or blisk) is made of 15-5 PH stainless steel and is manufactured in much the
same fashion as the pump impeller and inducer. Like the impeller, the turbine uses a polygon-style shaft
interface to minimize stress concentrations associated with the torque transfer. Since this is a partial
admission turbine, the turbine blades will pass in and out of fluid jets from the turbine nozzles during
operation and as such will be subjected to high-frequency load oscillations. These oscillations can lead to
high-cycle fatigue, so an outer turbine shroud is used to distribute blade tangential loading around the
circumference of the disk. This shroud increases disk centrifugal loads somewhat, but it reduces peak
transient stresses at the blade roots.
While 15-5 PH is not normally used for turbine components, the lower supply air temperature relative
to a conventional turbine in this application (only 900˚ F) made it possible to use a lower-cost corrosion-
resistant material like 15-5 PH. The yield strength of 15-5 will still be reduced to approximately 95.7 ksi at
this temperature, but with the current turbine design it is sufficient. It should be noted that the strength value
of 15-5 PH was set 15% lower than the reference value at this temperature to account for uncertainties
associated with the reference property data set.9
To verify that the turbine will meet static strength and deformation requirements under load, a refined
adaptive finite element mesh consisting of 1.1 million elements was constructed. As before, an initial
analysis with a coarse mesh helped to determine the regions likely in need of mesh refinement. In general,
the mesh was refined around the shaft interface and at the blade roots.
The loads and boundary conditions applied to the turbine were similar to those associated with the other
rotor components. A centrifugal load due to the 28,800 RPM counter clockwise rotation was imposed on
the disk, and a compression contact with the shaft was imposed along the hub edges and within the central
polygon interface region. Since this is a partial admission axial impulse turbine, only 43% of the turbine
disk is subjected to fluid loading at any point in time. To simulate this and the dynamic pressure of the fluid
jets impinging on the blades, a 90 psi pressure condition was applied to the inner surfaces of seven blades
in four sectors around the disk. While these blades will experience the highest temperatures from the jet

17
American Institute of Aeronautics and Astronautics
impingement, it is likely the entire turbine assembly will still be at a nearly uniform temperature after startup
due to fluid movement within the turbine cavity. As such, thermal stresses on the disk should be negligible
and were not considered in this analysis. Due to the nature of the air-heater supply system, startup thermal
transients should also be fairly benign.
The turbine deformation and safety factor results from this finite element analysis are shown in
Figure 20. In general, high stresses will occur at the shaft interface due to the large torque transfer in this
region. The interfaces between the central disk and the blades will also experience high stresses, although
these stresses are fairly uniform around the disk despite the sector loading due to the turbine shroud. Overall,
the minimum safety factor on this component is 2.39. This is greater than the minimum required value, so
the turbine meets all static strength requirements. The worst-case deformations are also fairly small at
approximately 0.018 inches as shown in the right side of Figure 20. These deformations will increase
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

somewhat during operation due to material creep at 900˚ F, but at these stresses and temperatures, the creep
rate of 15-5 PH will be less than approximately 0.015” over several hours of operation (the expected test
time). The radial clearance between the turbine and the turbine case is 0.045,” so even after a few hours of
operation, no contact or rubbing should occur.

Figure 20: Turbine safety factor (left) and deformation (right) during redline load conditions. Critical
stresses are present at the shaft interface and at the blade roots.

Shaft Structural Analysis


The shaft is made of 17-4 PH stainless steel. 17-4 PH is similar to 15-5 PH and was selected for similar
reasons, but it is ferromagnetic in nature and has yield strength of 110 ksi at 900˚ F. The shaft is circular in
cross-section at the inducer end but uses polygon interfaces at the impeller and turbine locations for torque
transfer. The bearings are located at the shoulders of the center portion of the shaft, and axial loads from
the turbine, impeller, and inducer are transferred through the shaft into the bearings at these shoulders. The
slots at the center of the shaft combined with the material ferromagnetic properties make it possible to
measure rotation speed using a variable reluctance sensor installed in the outer bearing housing.
In order to evaluate the stress distribution and deformations in the shaft, a refined adaptive finite element
mesh consisting of nearly 500,000 elements was created. Mesh refinements were incorporated at the fillets
between the shaft and polygon interfaces and at the bearing shoulders to capture stress gradients effectively.
It should be noted that while only part of the shaft will be exposed to the 900˚ F turbine gas temperatures,
the aforementioned strength value will be conservative for areas at low temperature and as such was used
for all shaft finite element analysis.
The primary loads and boundary conditions on the shaft were divided into two sets associated with
separate analysis cases. In the first case, torque was applied to the turbine end with the pump end fixed. In
the second case, torque was applied at the pump end with the turbine end fixed. In both cases, constant
temperatures conditions were applied at the turbine end (900˚ F), the turbine-side bearing interface (160˚

18
American Institute of Aeronautics and Astronautics
F), and the pump end (room temperature). This temperature profile was derived from the fact that heat will
be conducted from the turbine side of the shaft into the bearing and seal housing during operation. At this
point, the shaft will be cooled significantly through convective heat transfer to the bearing cooling oil and
the turbine-side seal purge gas. From here, the shaft will continue to cool to room temperature through the
bearing housing and into the pump case. Finally, a centrifugal load associated with the shaft rotation rate
was applied to the model. In general the stresses due to the shaft centrifugal load were negligible but the
boundary condition was included in the analysis anyway to be consistent with other rotor components. It
should be noted that while the bearing temperature value noted above is only an estimate, it is based on the
fact that the bearing oil flow rate will be adjusted during testing to ensure temperatures are no higher than
this. This assumption also makes it possible to avoid complex modeling of the convective heat transfer
behavior and simply assume fixed temperatures at the respective shaft regions.
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

Based on both finite element analysis cases, the highest stresses in the shaft will occur in a narrow region
where the shaft diameter and shape changes. This can be seen in Figure 21. The minimum safety factor for
Case 1 is 1.95 and for Case 2 is 2.25. These are both above the minimum required safety factor, so the shaft
meets all static strength requirements. Tangential deformation of the shaft is generally negligible and is not
shown here. Axial deformation is appreciable due to the thermal expansion of the shaft, however the
majority of this axial thermal expansion will occur between the turbine and the turbine-side bearing and
does not contribute to net movement of the shaft. The axial thermal expansion in this area was considered
when sizing the turbine rotor and turbine stator axial gap however. Some shaft thermal expansion also will
occur between the sets of bearings, however the dual bearing pre-load Belleville springs were sized to
provide enough clearance to allow this expansion. Similarly, the axial clearances between the pump and
the pump case were sized to provide enough clearance for the small amount of shaft thermal expansion
expected on this side of the turbine-side bearing assembly.

Figure 21: Shaft safety factor for analysis conditions 1 (left) and 2 (right)

IV. Rotordynamic Analysis and Bearing Selection


Given the very high rotating speeds of this turbopump, a rotordynamic analysis of the inducer, impeller,
shaft, and turbine assembly was necessary to provide guidance regarding bearing placement and stiffness
requirements, disk placement, and shaft sizing. XLRotor was the primary software tool used for this
rotordynamic analysis.10 XLRotor made it possible to idealize the rotor assembly using a segmented one-
dimensional beam model of the shaft while the rotor disk components were treated as point masses at their
centers of gravity with associated mass moments of inertia. In this application, the mass values and mass
moments of inertia values associated with the inducer, impeller, and turbine were derived from CAD models
of the hardware and imported into XLRotor. Bearings were also represented as nodal boundary conditions

19
American Institute of Aeronautics and Astronautics
with finite stiffness “springs” located between the shaft node and a fixed “grounded” node. In this case,
bearing stiffness values were derived from data-sheets associated with the selected bearing type. Finally,
the length, diameter, and material stiffness properties of each shaft segment at operating temperature were
specified thereby to complete the rotordynamic model.
The damped critical speed map or Campbell diagram generated from the rotordynamic analysis model
is shown in Figure 22. This diagram plots operating speed relative to the first four critical speeds of the
assembly. Critical speeds represent regions where the impeller would be rotating at precisely the resonant
frequency of the assembly, and as such they should be avoided. The critical speeds change slightly
depending on the actual rotating rate of the shaft due to gyroscopic forcing.
In many modern high-performance rotating machines, it is possible to operate above or between any of
the critical rotordynamic speeds. With heavy external damping, it may even be possible to operate at a
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

critical speed. Nevertheless there are risks associated with this that were considered too high for this
experimental turbopump. As such, the design requirements were amended during development to include
a critical speed margin of 10%. In other words, the hardware would need to be designed such that the lowest
critical speed would remain at least 10% above the highest expected operating speed. As shown in
Figure 22, this goal was achieved after many design iterations and changes in shaft diameter, rotor disk
spacing, bearing spacing, bearing number, and bearing stiffness. As a result, the turbopump should be free
of any serious resonant vibrations throughout the expected operating range. It should be noted that
maintaining a critical speed margin of 15% between the operating speed and the nearest critical speed is a
common practice in industrial applications. However from discussions with engineers at Mechanical
Solution, margins greater than 5% were still considered relatively safe. The 10% requirement represented
a compromise between these points of view.

Figure 22: The critical speed map or Campbell diagram for the turbopump rotor. The nominal system
running speed and redline speed are noted as well.

All of the aforementioned rotordynamic analysis is based on the assumption of an idealized and perfectly
balanced system. If the system is not balanced well, significant vibration can still occur and cause other
damage to the assembly even without reaching a critical speed. As such, the production plan for the
turbopump called for balancing of the individual rotor components as well as the entire rotor assembly by
Precision Balance Company Inc. in Indianapolis, IN prior final integration with the fixed structure at
Purdue. The balancing process generally requires material removal, but the amount of material removed is
usually small, and the material is removed from an area where it will not impact other aspects of operational
behavior or performance.

20
American Institute of Aeronautics and Astronautics
It should be noted that initially a two-bearing design with a shorter shaft was considered for this
turbopump in an effort to minimize costs, but this did not produce an acceptable critical speed map. This
ultimately drove the decision to use four bearings. Following this design decision, angular contact ball
bearings with a 25˚ contact angle were selected based on their availability, load capability, desirable
stiffness characteristics, high-speed capability, and ease of integration with the oil lubrication system. The
final bearing deemed optimal was the Timken #3MM206WI. The critical speed map shown in Figure 22
reflects the precise specifications associated with this bearing model.

V. Seal Design and Purge System


Numerous dynamic seals are required in a turbopump to prevent fluid from flowing along the shaft form
one area to another. These seals will inevitably leak slightly during service, so a purge and drain system is
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

required to prevent pump fluid, turbine drive gas, and bearing oil from mixing at any points. Since the
turbopump is designed to pump a storable oxidizer (even though the initial pumping fluid will be water),
contact between the pump fluid and the bearing oil must be prevented. Contact between the hot high-
pressure turbine drive gas and the bearing oil must also be prevented. In order to achieve this goal, four sets
of knife-edge labyrinth seals were employed on the pump-side of the assembly, and flow-paths were
provided between each set of seals for purge and drain lines. Similar dynamic seals were also included on
the back-face and front-face of the impeller to minimize leakage flow from the pump outlet back to the
pump inlet. As noted earlier, the radial locations of these impeller dynamic seals were selected to minimize
rotor axial load. The pump-side dynamic seal sets as well as the purge, drain, and leakage flow paths are
shown graphically in Figure 23. This arrangement is mirrored on the turbine-side of the turbopump.

Figure 23: Dynamic seal locations (in white) as well as leakage and purge flow paths. The main propellant
leakage and drain paths are in dark blue, the bearing oil leakage and drain paths are in green, and the purge
gas injection and leakage paths are in light blue.

Commercially manufactured dynamic seals are usually customized due to the fact that most of the
sealing requirements and environments are different very from one application to the next. Additionally,
seals for high-pressure or high surface speed environments must have precise fits to operate effectively.
Initially numerous commercial dynamic seal designs were considered for this turbopump, but none of these

21
American Institute of Aeronautics and Astronautics
options could match the pressure, temperature, surface speed, oxygen compatibly, and dimensional
requirements associated with the turbopump. It is possible to establish relationships with seal vendors so
that customized seals can be created, but this is generally time consuming and expensive. Given these
challenges, an in-house knife-edge seal design was created. In this design, several highly modular Teflon
disks with trimmed ridges (to act as labyrinth “teeth”) were installed in each dynamic seal set. This can be
seen in Figure 23.
To evaluate the effectiveness of the in-house dynamic seal design, a seal test apparatus was printed using
a 3D plastic printer and subsequently tested using a low-pressure source of compressed air. Based on the
leakage rates derived from these tests and general scaling laws (where leakage rate scales with seal linear
contact area and the square of the pressure differential), the estimated leakage rates associated with the
pump impeller were less than 5% of the total pump flow rate. While not optimal, this amount of leakage
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

was considered acceptable since the majority of the leakage flow in this area would simply pass around the
back-face or the front-face seals on the impeller and return to the inlet of the impeller. Leakage elsewhere
was expected to be less than 1% based on the same set of calculations. This was due to the much lower
pressure drop across the shaft dynamic seals. Again, these results were considered acceptable given the
experimental nature of this turbopump. Nevertheless, dynamic seals represent a ripe area for innovation in
a turbopump such as this.

VI. Test Stand Configuration and Layout


In order to test the turbopump design and evaluate turbine and pump performance, a concept for both
a water supply system for the pump and a hot-air supply system for the turbine was established using the
facilities available at Zucrow Laboratories. The simplified plumbing and instrumentation diagram for this
arrangement is shown in Figure 24.

Figure 24: A simplified plumbing and instrumentation diagram of the turbopump test facility.

In this open-loop test configuration, water from a high-flow main line would enter the pump after passing
through a low-pressure supply cut-off valve and a screen filter. After passing through the pump, this water
would be discharged through a high-pressure out-flow valve with a known discharge coefficient. By
measuring pressure both before and after the high-pressure outflow valve, it will be possible to determine

22
American Institute of Aeronautics and Astronautics
the total pump flow rate. Additionally, by measuring both temperature and pressure before and after the
pump, it will be possible to calculate the total supplied pump power.
In addition to the water feed system shown in Figure 24, drive air for the turbine would be produced by
the existing air-handling system available at the laboratory. This system is capable of transferring gas from
high-volume, high-pressure compressed air tanks through a supply valve, a high-flow dome pressure
regulator, and a natural gas-fired air heater to obtain the desired turbine inlet conditions. The hot gas would
subsequently pass through a splitter assembly and into the turbine manifold. The pressure and temperature
of the supplied flow would be measured in the splitter just prior to entering the turbine manifold. After
passing through the turbine, the gas pressure and temperature would be measured again using ports supplied
on the test article in the turbine exhaust manifold before existing through the annular choked nozzle. By
measuring the pressure and temperature in the exhaust manifold prior to the choked nozzle, it will be
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

possible to compute the total supplied mass flow rate to the turbine. Additionally, by measuring pressure
and temperature both prior to the turbine and after the turbine, it will be possible to compute total turbine
power transfer. Together with the pump output power measurements, the turbine pressure and temperature
measurements should make it possible to build a turbine and pump performance map for a range of turbine
inlet pressures and water outlet pressures (adjustable by changing the water outflow valve discharge
coefficient).
It should be noted that a closed-loop water systems was initially considered for the turbopump test stand.
While designing this system, it rapidly became apparent that heat build-up within the water tank would be
non-trivial due to the shear amount of power being dissipated from the pump in this arrangement. Since
changes in water temperature can significantly affect pump cavitation behavior, this was a problematic
configuration. A heat exchanger was considered, but the additional complexity was considered inadvisable.

VII. Conclusions and Lessons Learned


Purdue has designed a high-performance turbopump capable of pumping a storable oxidizer at 170 GPM
to 1400 psi using a hot gas generator. While the turbopump has yet to be tested, at this point all critical
design details have converged and the system appears to be capable of meeting all of the targeted
performance and operational requirements. During this development process there were numerous
challenges and lessons learned regarding design and analysis methods, system integration, testing, and
manufacturing. While these are two exhaustive to mention in detail, a few select lessons-learned are as
follows:

1. The turbine blade geometry was generally easier to design than expected, and turbine performance
was not especially sensitive to minor changes in blade geometry. Additionally, simple 1D flow
models and velocity triangle analysis methods were capable of predicting turbine efficiency with
reasonable accuracy. Mass flow predictions using this technique were less accurate.
2. The rotordynamic analysis and critical speed margins had a significant impact on the system design.
Had the shaft and bearing assembly been sized for strength alone, the design would have had very
poor rotordynamic characteristics and critical speeds well within the operational performance
envelope.
3. There were many more areas of high stress on the fixed and rotating structural components than
anticipated, and many more design iterations were required on some of these components before the
design converged. On the rotating structure, these challenges were generally due to the high
performance and rotation rate requirements of the hardware. On the fixed structure, many of these
high-stress areas were near fastener locations, and this was driven by the need to bolt everything for
easy assembly and testing. If assembly, disassembly, and manufacturing constrains were relaxed,
then welded, brazed, or cast components would likely have had fewer challenges with regards to
stress margins and stress concentrations near fasteners.
4. Finite element boundary conditions can be very challenging to correctly implement in a system like
this since one component of a design rarely if ever operates in a “vacuum.” In this case, the highly

23
American Institute of Aeronautics and Astronautics
integrated design and high performance requirements exacerbated the challenge. As such, extra
attention to FEA boundary conditions was necessary in order to accurately account for the integrated
structural stiffness and non-linearity between components in the assembly (including surface
contact, contact with gaps, etc).
5. In a high-performance system such as this, it is far too easy to design a part based on analysis goals
or first principles and end up with something that cannot be built without very significant design
changes. As such, most of the components on this project were designed from a manufacturing and
assembly perspective, and then changes dictated by analysis results were incorporated later. This
approach worked very well.
6. Direct Metal Laser Sintering (DMLS) methods were considered for many of the complex rotor
components in this project. While DMLS a very capable manufacturing technology that has gained
Downloaded by AFFTC TECHNICAL RESEARCH LIBRARY on September 21, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2015-4216

a lot of attention recently, it is not without limitations. In particular, it is difficult to obtain the high
tolerances and surface finishes needed for a system like this without significant post-process
machining. It can also be difficult to remove support material from the types of components
considered in this project due to the nature of their small internal cavities. These limitations had a
significant impact on the design and manufacturing feasibility of the impeller in particular.
7. Case deflections are critical when evaluating the performance losses associated with tip clearances
on an impeller. In this application, the casing gaps during operation were often larger than the fixed
gaps due to the very high pump exit pressures.

Acknowledgements
The authors would like to acknowledge the contributions and guidance of Matthew Steiner as well as the
great insight and design experience of Ed Bennett and Travis Jonas at Mechanical Solutions. The authors
would also like to recognize the School of Aeronautics and Astronautics at Purdue University for funding
the rapid prototyping work that was used during the turbopump development process.

References
1
Sutton, G. P., History of Liquid Propellant Rocket Engines, AIAA, Virginia, 2006, Chaps. 3, 4.
2
Huzel and Huang, Design of Liquid Propellant Rocket Engines, AIAA, Washington, DC, 1992, Chaps. 6.
3
Roark, Formulas for Stress and Strain 7th Edition, McGraw-Hill, New York, 2002. Chap. 16.
4
Balje, O. E, Turbomachines, a guide to design, selection and theory, John Wiley and Sons, Indianapolis,
1981.
5
CF Turbo, Software Package, CF Turbo Software and Engineering, Dresden, Germany, 2014.
6
Japikse, D., Marscher, W. D., and Furst, R. B., Centrifugal Pump Design and Performance, Concepts
NREC, Vermont, 1997, Chap. 6.
7
AK Steel, 2014, 15-5 PH Stainless Steel [Product Data Sheet], Retrieved from:
http://www.aksteel.com/pdf/markets_products/stainless/precipitation/15-5_ph_data_sheet.pdf
8
Stoffel Polygon Systems, Polygon Interface Design Guidelines [White paper], New York, 2014.
9
ASM International, Properties and Selection: Irons, Steels, and High Performance Alloys, Vol. 1, 10th
Edition, ASM International, 1990.
10
XL Rotor, Software Package, Rotating Machinery Analysis, Inc., North Carolina, 2014.

24
American Institute of Aeronautics and Astronautics

You might also like