You are on page 1of 8

1

Effect of Altitude on the Performance of Small


Centrifugal Compressors
Dries Verstraete and Kjersti Lunnan

Abstract: Small unmanned aircraft are currently limited to flight ceilings below 30,000 ft due to the lack of an appropriate
propulsion system. One of the most critical technological hurdles for the realisation of such small high altitude platforms
is the impact of the low Reynolds number conditions at altitude on the performance of small centrifugal compressors. The
current article investigates the influence of Reynolds number on the efficiency and pressure ratio of two small centrifugal
compressors using a 1D meanline performance analysis code. The results show that the efficiency and pressure ratio of
the 60 mm baseline compressor at the design rotational speed drops with 6-9% from sea-level to 70,000 ft. The impact on
the smaller 20 mm compressor is slightly more pronounced and amounts to 6-10%. Off-design changes at low rotational
speeds are significantly higher and can amount to up to 15%. Whereas existing correlations show a good match for
the efficiency drop at the design rotational speed, they fail to predict efficiency changes with rotational speed and can
therefore not be used to predict compressor maps at altitude.

Introduction
A wide variety of unmanned aerial vehicles (UAVs) are currently in operation around the world and their applications are
rapidly expanding. Very small platforms with limited range or
endurance are typically used for the so-called over the hill type
surveillance missions at low altitudes whereas extremely large
UAVs potentially fly for days to weeks at altitudes up to 65,000
ft [1]. The latter category of UAVs can serve a multitude of
roles ranging from meteorological and environmental science
studies over military intelligence and homeland security support to police search and rescue operations [2]. For most of
those applications light weight, small platforms are required
to keep the overall mission cost low [3, 4, 5] while extended
altitude capability is of prime importance for meteorological
research [4] and to extend the communications range of airborne communication relay payload [6, 7]. As shown in Figure 1 small platforms that operate above roughly 20,000 ft are
however not readily available, mainly due to the lack of a suitable propulsion system [7] and the Micro Propulsion Group
of The University of Sydney conducts research to support the
development of a small (<30 kg) high altitude (> 30,000 ft)
platform and its propulsion system.
High altitude propulsion system designs are constrained by
the desire to maximise vehicle endurance on station, and to
minimise the vehicle gross takeoff weight [9, 10]. This requires
both high engine thermal efficiency and low installed engine
weight [9]. The results of an extensive study of propulsion system options for very high altitude unmanned aircraft under
NASAs Environmental Research Aircraft and Sensor Technology (ERAST) program indicate that primary candidates for
this type of application are a highly turbocharged reciprocating
engine, and a semi-closed cycle gas turbine [10]. For UAVs as
small as the ones under consideration, recuperated open-cycle

Dries Verstraete. School of Aerospace, Mechanical and Mechatronic Engineering, The University of Sydney, Sydney, Australia
Kjersti Lunnan. School of Aerospace, Mechanical and Mechatronic Engineering, The University of Sydney, Sydney, Australia
Fourth Australasian Unmanned Systems Conference : 18 (2014)

Aim
Sea Ferret
Finder
Aerosonde
ScanEagle

Fig. 1. Altitude Capabilities of Small UAVs (based on data from


[8]

gas turbines seem suitable alternative candidates [11]. Compared to internal combustion engines, gas turbines namely offer the potential for higher reliability, longer engine life, and
superior compatibility with kerosene-based fuels [11]. However, achieving the desired thermal efficiency is challenging,
even with recuperation [11]. On the other hand, turbocharged
spark ignited or diesel engines offer a good adaptability for
low operating speed, a relatively low fuel consumption and
low manufacturing costs compared to other propulsion choices
[2, 13]. Turbocharger pressure ratios as high as 64, requiring
multi-stage intercooled turbochargers, can however be necessary [9] to overcome the significant drop in power output and
rise in specific fuel consumption with altitude of those engine
types [12].
All of the considered alternative propulsion systems use centrifugal compressors, either as part of the gas turbine engine
core or as part of the turbocharger for the piston engines, and
the impact of the low Reynolds number flow conditions at altitude on the performance of the compressor forms one of the
most important technological hurdles for the realisation of small

Fourth Australasian Unmanned Systems Conference , 2014

high altitude platforms. Low Reynolds number inlet conditions


namely promote laminar boundary layers and extend the laminar/turbulent transition flow region, resulting in greater flow
separation and higher flow blockage [13]. Both effects result
in a performance deterioration of the compressor and the entire propulsion system.
The current article investigates the impact of altitude operations on the performance of small centrifugal compressors and
is organised as follows. First a concise review of previous work
on the impact of Reynolds number on the performance of centrifugal compressors is given. The 1D performance analysis
method employed in the current article is described next and
the obtained results are described and discussed.

Impact of Reynolds Number on Centrifugal


Compressor Performance
Considerable attention has been paid to the impact of Reynolds number on compressor performance as Reynolds number variations also impact the use of scaled model test data
to predict full-scale machine performance [14] and a number
of literature reviews on this subject have been conducted recently (a.o. [14, 15]). Research shows that changes in compressor performance due to Reynolds number variations involve a
change in efficiency, a shift in flow rate and a change in the
pressure rise. Of those the effect on efficiency has been most
thoroughly investigated and correlations similar to equation 1
have often been suggested for the varation in peak efficiency
with Reynolds number

n
1
Reref
= a + (1 a)
[1]
1 ref
Re
where is the compressor efficiency, Re the Reynolds number,
a is the fraction of losses that are independent of the Reynolds
number (typically between 0 and 0.5 [15]) and n the Reynolds
ratio exponent, with a value between 0.16 and 0.5 [15]. The
subscript ref refers to the reference conditions used for scaling. As the identification of values for a and n has caused substantial difficulties, this expression has recently been replaced
by equation 2 when it was identified that the physical variation
of n was related to the variation in friction factor [15]. This
lead to:

n
cf
1
= a + (1 a)
[2]
1 ref
cfref
where cf is the friction factor of an equivalent fully turbulent
pipe flow. According to the authors of [15], considerable error
still remains if a is assumed to be constant and they derived an
alternative expression in the form of:
= Bref

cf
cfref

[3]

where Bref is the inefficiency due to friction losses at the reference conditions, which can be determined as [14]:
Bref = 1 ref
This, according to the authors of [15, 14] has as a major advantage that the correlation is not directly coupled to any model

for the losses. As the correlation depends on the friction factor


of an equivalent fully turbulent plate flow, the Reynolds number for a turbomachinery component has to be transformed to
that of a plate. For centrifugal components this can be done as
follows [14]:
Rep =

1L 1
Re
4 b2 sin0

[4]

where Rep is the equivalent plate Reynolds number, L is the


characteristic length which is taken as the blade length [14], b2
is the blade height at the rotor outlet, 0 is the design stagger
angle and the flow coefficient of the compressor. Once the
equivalent plate Reynolds number is known, the well-known
Colebrook equation can be used to determine the pipe friction
factor [14].
Whereas less ubiquitous similar laws can be found for changes
in flow coefficient and pressure coefficient . Equation 5 describes the shift of best efficiency to lower flow coefficients at
reducing Reynolds numbers [14]:

1 ref
=C

cfref

[5]

where the constant C is a function of stagger angle and blade


spacing [14]. The variation in pressure coefficient is finally
given by [14]:

=
ref
ref

[6]

Method
For the current work, a one dimensional (1D) design and
performance analysis strategy analysis of a range of compressor sizes is performed. The 1D analysis is based on the impeller design and performance analysis strategy from [16] using the accompanying software COMPAERO, which is available from [17]. The strategy detailed in [16] consists of two
distinct steps. The first step consists of a preliminary design
code, which generates the initial candidate geometry for the
impeller to match the desired performance goals [16]. The preliminary design strategy consists of empirical relationships developed through extensive experimentation [16]. The empirical
relationships are valid for flow coefficients between 0.003 and
0.2 and for stage pressure ratios up to 3.5 [16]. The predicted
performance goals do not represent optimum achievable efficiencies but rather represent good efficiency levels and stable
operating conditions [16].
Once the compressor geometry is generated the second step
of the strategy is initiated: the 1D performance analysis. The
main aim of this step is to predict the performance of the impeller under design and off-design conditions. The mean-line
aerodynamic performance analysis consists of a combination
of basic theoretical relationships and empirical correlations used
to model the losses, as shown in Figure 2. Based on the specified geometry, the performance analysis determines the tip flow
conditions. Once these are known, the ideal discharge conditions are determined from which the blade work is calculated.
After that total pressure losses are determined from the loss
models which results in the thermodynamic conditions at the

method does not allow the specification of a tip clearance factor. Tip clearance corrections for efficiency are therefore applied in post-processing based on ref. [18]:

t
=k

[9]

where t is the tip clearance, and h is the blade height at the impeller exit. A common value for the constant k for large-scale
compressors is 0.3, whereas experiments show that for ultra
micro gas turbines in the millimeter scale range the value of
k increases to around 0.65 [18]. For the current work a value
of 0.3 is adopted for k regardless of the impeller size and the
tip clearance t is set to 30 % of the blade height for the smallest compressor [19]. When increasing the compressor size, the
tip clearance height t is held constant and it is assumed that
the tip clearance losses are independent of pressure ratio and
operating conditions.

Fig. 2. Flow chart for the impeller performance analysis, adopted


from [16]

tip [16]. This step, indicated in grey in Figure 2, includes the


Reynolds number effects investigated here. If the tip mass balance is correct, other parasitic losses are added which results
in the absolute flow properties at the tip.
Wall friction total pressure losses, and thus Reynolds number effects, are modelled in COMPAERO using a general formulation for the skin friction coefficient that encompasses both
laminar and turbulent flow and takes into account the influence
of surface finish [16]. Skin friction coefficients are correlated
as a function of Reynolds number based on the pipe diameter
and three well established models are used [16]. When the pipe
Reynolds number is less than 2000, the following correlation
is used for laminar flow:
cfl =

16
Red

[7]

For pipe Reynolds numbers higher than 2000 turbulent flow


correlations are used and a distinction is made between turbulent flow over a smooth surface and turbulent flow over a rough
surface [16]. For rough surfaces, the following correlation is
employed:
h e i
1
p
= 2 log10
[8]
3.71d
4 cft,r
where cft,r is the turbulent, rough skin friction coefficient, e is
the peak-to-valley surface roughness, and d is the pipe diameter on which the Reynolds number is based [16]. Transitions
between laminar and turbulent and smooth and rough zones are
modelled as weighted averages of their respective values [16]
and surface roughness is considered to be significant when:
Ree = (Red 2000) e/d > 60
Whereas COMPAERO incorporates loss models for the most
prevalent types of losses, its centrifugal compressor analysis

Results
Below the results of the current study are given for 2 different compressor sizes, which are defined by their blade length.
Blade length is selected here as the indicator for compressor
size as it determines the Reynolds number of the compressor. The baseline compressor has an impeller blade length of
60 mm. The second compressor has a blade length of 20 mm
so that the investigated Reynolds numbers are much smaller
than for the baseline compressor. Additional compressor parameters are given in Table 1. As shown in the table, the design rotational speed for both compressors is selected to yield
a pressure ratio of approximately 2.2. For each of the compressors 4 sets of results are presented. First the impact of altitude
(Reynolds number) on the peak efficiency of each compressor is presented for a range of rotational speeds. The effect of
altitude on the flow coefficient where the efficiency peaks is investigated next. Finally the effect of altitude on the compressor
pressure ratio is given and compressor maps for sea-level and
altitude conditions are discussed. However, before presenting
the results obtained for those 2 compressors, the methods of
COMPAERO are first validated.
Table 1. Geometric compressor parameters
Blade passage length 60 mm
Number of full blades []
6
Number of splitter blades []
6
Disk diameter [mm]
80.0
Hub inlet diameter [mm]
28.0
Shroud inlet diameter [mm]
56.0
Inlet blade angle [ ]
30.0
Discharge blade angle [ ]
43.5
Design rotational speed [RPM] 90,000
Design mass flow rate [kg/min]
14.2
Design pressure ratio []
2.20
Design flow coefficient []
0.102
Tip clearance t/h [%]
15.5

20 mm
7
7
31.5
11.4
21.8
30.2
42.4
230,000
2.05
2.18
0.094
30.0

Adiabatic Efficiency

0.78
60mm Compressor

0.76

Aungier's Blade Code


0.74

Pelz & Stonjek

0.72
0.70
0.68
0.66

70,000 RPM
0

200,000

400,000

600,000

800,000

1,000,000

0.78

Adiabatic Efficiency

Validation of COMPAERO
Before using the COMPAERO code to explore the influence of Reynolds number on the efficiency of centrifugal compressors, the code was first validated using experimental data
presented in reference [14]. As the COMPAERO code calculates the blade angles internally rather than allowing the user
to specify the exact geometry and geometric data of the tested
compressor is limited, geometry matching between COMPAERO and the experimental results is limited to external diameters at the inlet and exit planes of the compressor. Despite
this restriction, the maximum adiabatic efficiency is predicted
accurately as shown in Figure 3. The figure shows the variation in adiabatic efficiency with flow coefficient for both the
experimental data and the results of the 1D blade code.

Fourth Australasian Unmanned Systems Conference , 2014

0.76
0.74
0.72
0.70
0.68

80,000 RPM
0

200,000

400,000

600,000

800,000

1,000,000

Adiabatic Efficiency

0.78
0.76

Altitude effect on peak efficiency


As COMPAERO gives acceptable results for the peak efficiency prediction at lower Reynolds numbers, the 1D blade
performance analysis code is used to predict the change in
peak efficiency with Reynolds number for the 2 compressors
of Table 1. The results of the analysis for 3 different rotational
speeds for the 60 mm compressor are shown in Figure 4. The
figure also gives results using the method from Pelz & Stonjek [14]. For each of the RPMs the range in Reynolds numbers
corresponds to a change in altitude from sea level to 70,000 ft.

Adiabatic Efficiency

As shown both the maximum efficiency and the flow coefficient where this maximum efficiency occurs match well with
the experimental data. The drop in efficiency as lower mass
flow rates is however much steeper for the COMPAERO results
than for the experimental data, and the compressor designed
with COMPAERO chokes at lower flow coefficients. Both of
these discrepancies can be attributed to differences in blade angles at the inlet and exit plane of the compressor. For the higher
Reynolds number experimental test case presented in [14] a
bigger discrepancy was found for the peak efficiency. This can
be attributed to the higher dependency of the friction coefficient to surface roughness for the fully turbulent flow regime
[14, 16]. As COMPAERO does not allow specifying a specific surface roughness, the peak efficiencies could not be fully
matched. As the main objective of the current work is the investigation of changes in peak efficiency with altitude (Reynolds
number) for small compressors and thus low Reynolds numbers, the validation is however judged to be acceptable. After
all the higher Reynolds number case of [14] represents a compressor with an outer diameter of 2.24 m, or almost two orders
of magnitude larger than the compressors under investigation.

= 2.20

0.72
0.70
0.68

Fig. 3. Validation with centrifugal fan from [14]

Reference Point

0.74

90,000 RPM
0

200,000

400,000

600,000

800,000

1,000,000

0.78
0.76
0.74
0.72
0.70

100,000 RPM
0

200,000

400,000

600,000

800,000

1,000,000

Reynolds Number

Fig. 4. Change in efficiency with Reynolds number for a 60mm


diameter compressor

As shown in Figure 4, the method from ref. [14] predicts


a slightly higher drop in efficiency with altitude at the design RPM of 90,000. However the trends between both tools
are consistent. At off-design RPMs, slightly higher discrepancies are however found between the absolute efficiency levels whereas efficiency changes however follow similar trends.
As the method from Pelz & Stonjek only predicts changes in
efficiency with Reynolds number, its application to different
RPMs seems to fail to account for change in efficiency levels
with pressure ratio. At higher RPMs the flow Mach numbers
namely increase which results in higher pressure ratios but also
in higher levels of losses and thus lower efficiencies. Whereas
this is reflected in the results of COMPAERO shown in Figure 4, this trend is not apparent in the results obtained with the
method of Pelz & Stonjek [14].
In a second case a much smaller compressor is analysed to
investigate the change of efficiency at much lower Reynolds
numbers. Results for 4 different RPMs for the smaller compressor are shown in Figure 5. The results in Figure 5 show that
for this compressor COMPAERO also predicts a slightly lower
efficiency loss with altitude at the design RPM of 230,000 than
the method of Pelz & Stonjek. As for the larger compressor
the method of Pelz & Stonjek does not predict changes in efficiency with RPM and larger variations occur at off-design.

0.72
20mm Compressor

0.70
Adiabatic Efficiency

Aungier's Blade Code


0.68

Pelz & Stonjek

magnitude of this shift, the variation in efficiency with flow


coefficient of both the 20 mm and 60 mm compressor is determined for a range of altitudes. The results are depicted in
Figures 6 and 7.

0.66
0.64
0.62
0.60
0.58
0.56

200,000 RPM
0

100,000

200,000

300,000

400,000

0.72

Adiabatic Efficiency

0.70
Reference Point

0.68

= 2.18
0.66
0.64
0.62
0.60
0.58

230,000 RPM
0

100,000

200,000

300,000

400,000

Adiabatic Efficiency

0.72
0.70
0.68
0.66
0.64
0.62
0.60

260,000 RPM
0

100,000

200,000

300,000

400,000

Adiabatic Efficiency

0.72
0.70
0.68
0.66
0.64
0.62

290,000 RPM
0

100,000

200,000

300,000

400,000

Reynolds Number

Fig. 5. Change in efficiency with Reynolds number for a 20mm


diameter compressor

Fig. 6. Change in optimum flow coefficient with altitude for a 60


mm compressor

As shown, the optimum flow coefficient at sea level is around


0.095 for both compressors and an almost linear drop occurs
with Reynolds number (altitude). At 70,000 ft the optimum
flow coefficient for both compressors is around 0.09. Using
those results the slope / of equation 5 is calculated as
15.0 for the 60mm compressor and 22.1 for the 20mm compressor. The steeper slope for the smaller compressor can be attributed to the higher sensitivity of efficiency to altitude for that
compressor. A comparison between Figures 6 and 7 shows that
the drop in efficiency with altitude is much more pronounced
for the 20 mm compressor than for the 60 mm compressor.
This can be attributed to the lower range of Reynolds numbers
at which the smaller compressor operates, which results in an
increased importance of skin friction and a more prominent
change in friction coefficient with Reynolds number reduction.
Whereas the change in flow coefficient for both compressors is
the same, this leads to a higher value for the slope /.
Altitude effect on pressure ratio
The operating Reynolds number not only affects the efficiency and optimum flow coefficient but also has an impact on
the pressure ratio of the compressor as the friction (pressure)

A comparison between Figures 4 and 5 shows that the effect of altitude is slightly larger for the smaller compressor. At
70,000 ft the efficiency drops by almost 8% for the compressor
with a blade length of 20 mm, compared to a drop of 6% for
the 60 mm compressor. The peak sea-level efficiency is additionally about 6% lower for the smaller compressor, which predominantly stems from the larger relative tip clearance height.
For the smaller compressor tip clearance is 30% of the blade
height whereas for the larger compressor this is reduced to
15.5%. Variations in efficiency with RPM are similar for both
compressors.
Altitude effect on optimum flow coefficient
As shown by equation 5, the drop in efficiency with altitude
(Reynolds number) is accompanied by a shift in the flow coefficient at which the peak efficiency occurs. To examine the

Fig. 7. Change in optimum flow coefficient with altitude for a 20


mm compressor

Fourth Australasian Unmanned Systems Conference , 2014

losses will increase with a reduction in Reynolds number. The


impact of the operating Reynolds number on the compressor
pressure ratio is given in Figures 8 and 9. The figures show the
variation in normalised pressure ratio and adiabatic efficiency
at the design RPM and flow coefficient. The figures show that
the pressure ratio for both compressors drops by up to 5% at
70,000 ft compared to sea level values.

Fig. 10. Compressor Pressure map for the 60 mm compressor

Fig. 8. Change in efficiency and pressure ratio with altitude for a


60 mm compressor

Fig. 11. Compressor Efficiency map for the 60 mm compressor

Fig. 9. Change in efficiency and pressure ratio with altitude for a


20 mm compressor

Figures 8 and 9 also show that the normalised pressure ratio


drops less than the normalised adiabatic efficiency and the drop
in both the pressure ratio and adiabatic efficiency is relatively
modest up to 40,000 ft. Once higher altitudes are reached a
very sharp drop in both pressure ratio and efficiency however
occurs. This general trend is in line with computational results
presented in [13] for a transonic centrifugal impeller.
Altitude effect on compressor map
The combined effect of Reynolds number on pressure ratio, efficiency and mass flow rate can be investigated using socalled compressor component maps. Figures 10 and 11 show
the component maps obtained for the 60 mm compressor at
sea-level and 70,000 ft. The blue line in Figure 11 represents
the location of peak efficiency at each rotational speed whereas
the red line indicates the location of the maximum pressure ratio. As shown, a small loss in the obtained pressure ratio occurs at all rotational speeds and the loss increases for increas-

ing RPM. This is in line with experimental results presented in


ref. [13].
As can be seen from Figure 11 the loss in efficiency with
altitude increases significantly at lower RPMs. At sea level
a small increase in efficiency can be observed for rotational
speeds lower than the design RPM. As the velocity magnitudes are smaller at lower rotational speeds, the corresponding
losses are reduced which results in this small efficiency rise.
At 70,000 ft this is no longer the case and the peak efficiency
at lower rotational speeds is lower than that at the design RPM.
This is a consequence of the much lower Reynolds number at
lower RPMs and the sharp drop in efficiency at low Reynolds
numbers. As shown in Figure 12 a similar trend can be observed for the 20 mm compressor albeit with a much sharper
drop in efficiency at low rotational speeds. As shown, a drop
of almost 15% in efficiency occurs at 110,000 RPM due to
the extremely low Reynolds number at low rotational speeds at

Fig. 12. Compressor Efficiency map for the 20 mm compressor

70,000 ft altitude.
As shown in Figure 10 a small decrease in maximum flow
rate capacity also occurs at altitude. The reduction in Reynolds
number namely leads to an increase in the boundary layer thickness which results in a choking of the blade passages at slightly
smaller mass flow rates. The reverse trend seems to be apparent at the point of maximum pressure ratio. Whereas this seemingly indicates an increase in mass flow rate at surge, the stationary meanline analysis of COMPAERO cannot be used to
accurately predict a highly dynamic phenomenon like surge.
The overal consequence of the changes in the compressor
map in altitude can have a significant impact of the overall engine performance. The drop in pressure ratio of the compressor
will namely result in the need for an increase in engine rotational speed to obtain a given operating pressure ratio. The reduction in efficiency on the other hand results in an increase
in compressor power. To operate at a given pressure ratio both
effects will require an increase in turbine power and will thus
require the engine to operate at a higher turbine inlet temperature to provide the same specific power. The reduction in mass
flow rate will further exacerbate this effect and will result in a
significant increase in turbine inlet temperature to provide the
same thrust or power output, which will result in a considerable
loss in engine life.

The thicker boundary layer at low Reynolds number conditions leads to a reduction in maximum mass flow rate
capacity at compressor choke
All of the above effects can have a considerable impact on
engine life as they will drive operation of the engine to higher
turbine inlet temperatures. Updated compressor maps should
thus be used for investigations of altitude performance of small
engines.

References

1. Grundlach J., Designing Unmanned Aircraft Systems: A Comprehensive Approach, AIAA Education Series, American Institute for Aeronautics and Astronautics, 2012.
2. Perez P.L., Boehman A.L., Performance of a single-cylinder
diesel engine using oxygen-enriched intake air at simulated highaltitude conditions. Aerospace Science and Technology, Vol:14,
2010, pp. 83-94.
3. Holland G.J., Webster P.J., Curry J.A., Tyrell G., Gauntlett D.,
Brett G., Becker J., Hoag R. and Vaglienti W., The Aerosonde
Robotic Aircraft: A New Paradigm for Environmental Observations, Bulletin of the American Meteorological Society,
Vol:82(5), 2001, pp. 889-901.
4. Curry J.A., Maslanik J., Pinto J.O., Drobot S., Cassano J. and
Holland G.J., Applications of Aerosondes for RIME, Extended
Conclusions
Abstracts of the Workshop on the Ross Island Meteorology Experiment, September 2001.
One of the most critical technological hurdles for the reali5. Anon., Airborne Communications Relay Could Become Primary
sation of small high altitude platforms is the impact of the low
Mission for Tactical UAVs, http://defense-update.com/features/
Reynolds number conditions at altitude on the performance
2010/january/airborne relays for uavs 110110.html, Accessed
of small centrifugal compressors. The current article investiAugust 2014.
gates the influence of Reynolds number on the efficiency and
6. Jenkinson L.R. and Marchman J.F., III, Aircraft Design Projects
pressure ratio of two small centrifugal compressors using a
for Engineering Students, Butterworth-Heinemann, 2003.
1D meanline performance analysis code. The investigations re7. Verstraete D., Gibbens P., and Wong KC., Is a small UAV with
ported here lead to the following conclusions
extended altitude capabilities feasible? AIAC14 Fourteenth Aus The reduction in Reynolds number at high altitude leads
tralian International Aerospace Congress, February 2010.
to a considerable loss in compressor efficiency. At 70,000
8. Anon., 2013 Worldwide UAV Roundup, American Institute of
ft the adiabatic efficiency of the compressor can be reAeronautics and Astronautics, 2013.
duced by up to 10% at the design rotational speed. Both
9. Rodgers C., Turbocharging a High Altitude UAV C.I. Engine
investigated compressors suffered to a similar extent with
37th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and
only a small increase in efficiency loss for the smaller
Exhibit, July 2001.
compressor.
10. Bettner J.L., Blandford C.S., and Rezy, B.J., Propulsion System
Assessment for Very High Altitude UAV Under ERAST NASA Under off-design operation, at much lower rotational speeds,
CR-195469, May 1995.
the loss in efficiency is further increased, especially for
11. Vick M.J., Heyes, A., and Pullen, K., Design Overview of a
the smaller compressor under investigation. At 50% of
Three Kilowatt Recuperated Ceramic Turboshaft Engine, Jourthe design rotational speed an efficiency loss of 15%
nal of Engineering for Gas Turbines and Power, Vol:132, 2010,
was observed when operating at 70,000 ft compared to
pp. 092301:1-9.
sea-level conditions for the same rotational speed. This
12. Shin Y., Chang S.-H., and Koo S.-O., Performance test and simcan be attributed to the significantly lower operational
ulation of a reciprocating engine for long endurance miniature
Reynolds number at low rotational speeds.
unmanned aerial vehicles, Proc. IMechE Part D: J. Automobile
The loss in efficiency with altitude is relatively small up
Engineering, Vol:219, 2005, pp. 573-581.
to 40,000 ft after which it starts to increase rapidly.
13. Zheng X., Lin Y., Gan B., Zhuge W., and Zhang Y., Effects of
Reynolds number on the performance of a high pressure-ratio
Operation at altitude also leads to a loss in pressure raturbocharger compressor, Science China Technological Sciences,
tio at a given rotational speed, albeit to a lesser extent
Vol:56(6), 2013, pp. 1361-1369.
than the loss in efficiency. Pressure losses of around 6%
are observed for the design rotational speed at 70,000 ft
14. Pelz P.F., and Stonjek, S.S., The Influence of Reynolds Number
altitude
and Roughness on the Efficiency of Axial and Centrifugal Fans -

15.

16.
17.

18.

19.

20.

21.
22.

Fourth Australasian Unmanned Systems Conference , 2014


A Physically Based Scaling Method Journal of Engineering for
Gas Turbine and Power, Vol:135, 2013, pp. 052601-1:8.
Casey M.V., and Robinson C.J., A unified correction method
for Reynolds number, size, and roughness effects on the performance of compressors, Proc. IMechE Part A: J. Power and
Energy, Vol:225, 2011, pp. 864-876.
Aungier R., Centrifugal Compressors: A Strategy for Aerodynamic Design and Analysis. ASME Press, New York, 2000.
Aungier R., Turbomachinery Aerodynamic Technology Design
& Analysis Software and Consulting Services. http://www.
turbo-aero.com/Pages/default.aspx, April 2012.
Nagashima T., and Teramoto S., Aero-Thermal Research Particulars in Ultra-Micro Gas Turbines, Micro Gas Turbines, Educational Notes RTO-EN-AVT-131, Paper 3, Neuilly- sur-Seine,
France, 2005.
Engin T., Gur M., and Scholz R., Effects of tip clearance and
impeller geometry on the performance of semi-open ceramic
centrifugal fan impellers at elevated temperatures, Experimental
Thermal and Fluid Science, Vol:30, 2006, pp. 565-577.
Verstraete D., Hendrick P., Djanali V., Ling J., Wong KC., and
Armfield S., Micro Propulsion Activities at The University of
Sydney, Powermems 2010 Conference, September 2010.
Schreckling K., Gas Turbine Engines for Model Aircraft, The
Modellers World Series, Traplet Publications Ltd., 2007.
Schreckling K., Home Built Model Turbines, The Modellers
World Series, Traplet Publications Ltd., 2005.

You might also like