Sunteți pe pagina 1din 9

Energy 34 (2009) 14841492

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Methodologies for predicting the part-load performance


of aero-derivative gas turbines
F. Haglind*, B. Elmegaard
Technical University of Denmark, Department of Mechanical Engineering, DK-2800 Kgs. Lyngby, Denmark

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 24 June 2008
Received in revised form
20 March 2009
Accepted 17 June 2009
Available online 8 August 2009

Prediction of the part-load performance of gas turbines is advantageous in various applications. Sometimes reasonable part-load performance is sufcient, while in other cases complete agreement with the
performance of an existing machine is desirable. This paper is aimed at providing some guidance on
methodologies for predicting part-load performance of aero-derivative gas turbines. Two different
design models one simple and one more complex are created. Subsequently, for each of these models,
the part-load performance is predicted using component maps and turbine constants, respectively.
Comparisons with manufacturer data are made. With respect to the design models, the simple model,
featuring a compressor, combustor and turbines, results in equally good performance prediction in terms
of thermal efciency and exhaust temperature as does a more complex model. As for part-load
predictions, the results suggest that the mass ow and pressure ratio characteristics can be well predicted with both methods. The thermal efciency and exhaust temperature, however, are not well
predicted below 6070% load when using turbine constants and assuming constant efciencies for
turbomachinery.
2009 Elsevier Ltd. All rights reserved.

Keywords:
Aero-derivative gas turbine
Performance
Part-load
Map
Turbine constant

1. Introduction
Prediction of the part-load performance of gas turbines is
advantageous in various applications, such as with stationary
power plants, aero engines, and mechanical drive and prime mover
applications. Sometimes reasonable part-load performance is
sufcient, while in other cases complete agreement with the
performance of an existing machine is desirable. Which methodology for part-load performance prediction is the more appropriate
to use, depends on the required accuracy and the availability of
performance data.
The study covered in this paper is part of research aimed at
nding alternative prime movers for large ships that impact the
environment less than todays slow-speed diesel engines. One of
the technologies under consideration is gas and steam turbine
combined cycles. Part-load performance of naval prime movers is of
great importance because of the considerable portion of the
running time spent at part-load. Therefore, models are developed
for accurately predicting part-load performance of gas turbines and
other components included in such plants.

* Corresponding author. Tel.: 45 45 25 41 13; fax: 45 45 93 52 15.


E-mail address: frh@mek.dtu.dk (F. Haglind).
0360-5442/$ see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.energy.2009.06.042

The paper is aimed at providing some guidance on methodologies for predicting part-load performance of aero-derivative gas
turbines. Such gas turbine type is considered because of its
common use for marine applications. For predicting part-load
performance, it is widely recognised that the use of component
maps is the most accurate way. However, this approach requires
that suitable maps are available, which, in practice, they rarely are
(other than for manufacturers). An alternative approach, which is
also considered here, is the use of a turbine constant for the turbine,
i.e. a constant that governs the relation among ow capacity,
pressure ratio and inlet temperature for the turbine. Such approach
simplies the calculation procedure and requires no component
information.
In order to compare different methods for performance
modelling, an existing gas turbine is selected. Performance data,
covering the load range, for this machine were provided by the
manufacturer. Two different design models one simple and one
more complex are created. Subsequently, for each of these
models, the part-load performance is predicted using component
maps and turbine constants, respectively. Comparisons with
manufacturer data are made.
Other authors have used similar approaches for predicting partload performance of gas turbines. For example, Zhu and
Saravanamuttoo [1] used generalized component maps and related
these to an actual machine. This approach is very similar to the one

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492

Nomenclature

Abbreviations
C
compressor,
CC
combustion chamber,
CT
compressor turbine,
DNA
dynamic network analysis (computer simulation
program),
E
exhaust,
I
inlet,
MGO
marine gas oil,
PT
power turbine.
Notations
COT
combustor outlet temperature ( C),
turbine constant,
CT
fraction cupper loss,
FCU
LHV
lower heating value (MJ/kg),
TIT
turbine inlet temperature ( C),

presented here. Pathak et al. [2] as well as Najjar et al. [3] developed
correlations for the performance based on available data for gas
turbines. Deidewig and Dopelheuer [4] used a similar approach
where they explicitly used the ideal gas law and assumed constant
polytrophic efciency.
In Section 2 the numerical simulation tool used for the calculations, including the additional development of the tool required for
the current study, is described. The selected gas turbine and the
design models are described in Section 3. In Section 4 the different
methodologies for predicting part-load performance are explained.
The performance results for part-load operation are presented and
discussed in Section 5. Finally, in Section 6 the conclusions are
outlined.
2. DNA the simulation tool
In this section the fundamentals of the simulation program used
for this work are described. Moreover, the development work of the
program for the purpose of this research is explained briey.

P
R
T
W

h
b
4
y

1485

pressure (bar),
gas constant (kJ/kg K),
temperature ( C),
mass ow (kg/s),
efciency,
auxiliary coordinate in component map,
mass ow coefcient (m2),
specic volume (m3/kg).

Subscripts
3
compressor turbine inlet,
4
power turbine inlet,
Cis
compressor, isentropic,
corr
corrected,
D
design,
exh
exhaust,
PL
part-load,
std
standard atmosphere (T 15  C, P 101.325 kPa),
th
thermal,
Tis
turbine, isentropic.

systems. Both steady state (involving algebraic equations) and


dynamic (involving differential equations) simulations can be
conducted.
2.2. Implementation of component maps in DNA
For the purpose of this work, the DNA component library has
been extended with steady state models describing the part-load
characteristics of compressors and turbines of real gas turbine
engines. These models are based on component maps provided
with the GasTurb software, version 10 [8], which have been
compiled by the developer of GasTurb from data published in the
public domain. The maps are represented by les stating values for
ow, pressure ratio, isentropic efciency and speed for the
complete operating range of the component. A schematic map le
for a compressor is shown in Fig. 1. The beta lines are auxiliary
coordinates which have values between 0 and 1, and in the map

Pout
2.1. Basics of DNA
DNA (dynamic network analysis) is a simulation tool used for
energy systems analyses [5,6]. It is the present result of an ongoing
development at the Department of Mechanical Engineering,
Technical University of Denmark, which began with a Masters
Thesis work in 1990 [7]. Since then the program has been developed to be generally applicable for covering unique features, and
hence supplementing other simulation programs.
In DNA the physical model is formulated by connecting the
relevant component models through nodes and by including
operating conditions for the complete system. The physical model
is converted into a set of mathematical equations to be solved
numerically. The mathematical equations include mass and energy
conservation for all components and nodes, as well as relations for
thermodynamic properties of the uids involved. In addition, the
components include a number of constitutive equations representing their physical properties, e.g. heat transfer coefcients for
heat exchangers and isentropic efciencies for compressors and
turbines. The program includes a component library with models
for a large number of different components existing within energy

Pin

is

N
Tin

Tin
Pin

Fig. 1. Schematic compressor map.

1486

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492

they are represented by parabolic functions [9]. The introduction of


beta lines is a convenient way of avoiding the ambiguity arising
when the speed lines in the map are horizontal or vertical. The
efciencies can be given as contours, and these are illustrated in the
gure by dotted lines. Mathematically, the map provides the relationships among ow, speed and beta lines:

f1

!
p
Tin
N
; b; p 0
Pin
Tin

(1)

Isentropic efciency, speed and beta lines:

!
N
f2 his ; b; p 0
Tin

(2)

And pressure ratio, speed and beta lines:

f3

!
Pout
N
; b; p 0:
Pin
Tin

(3)

For turbines, only the rst two relationships are needed. Suitable
map les have been implemented into DNA. At design, a location
within the map is selected (in terms of beta and speed), and the
map data are linearly scaled (in the case of pressure ratio, it is
pressure ratio - 1 that is scaled) in order to achieve the data chosen
at design. Subsequently, the component data at any off-design
operating point are obtained from the map le using linear interpolation. In DNA the operating point may be specied by any two of
the variables in the equations, giving a high exibility in specifying
the design point and part-load operation according to a given need.
3. Design point performance
In this section, the performance data for the gas turbine studied
are presented, and two different design models are created. The
full-load condition is chosen as design point.

Table 1
Full-load performance of the LM2500 gas turbine
Inlet mass ow [kg/s]
Fuel ow [kg/s]
Exhaust ow [kg/s]
Compressor outlet temperature [ C]
Exhaust temperature [ C]
Relative inlet pressure loss [%]
Relative outlet pressure loss [%]
Generator efciency [%]
Compressor pressure ratio [-]
Electrical power output [kW]
Thermal efciency [%]

88.4
1.934
89.5
495
533.8
1
2.9
97.5
23.54
31 207
37.70

essential parameters for full-load used to create performance


models in DNA are listed in Table 1. The manufacturer data are
estimated average performance data (not guaranteed) obtained
using a performance model. This model is based on performance
data compiled from operation data, and it is adjusted periodically as
the performance data are received from the eld. No component
efciencies, maximum cycle temperature or information on cooling
ows were given. Neither was any variable guide vane schedule
specied as related to the part power performance. Ambient
conditions are 15  C and 1.01325 bar. Based on the ow gures it
can be concluded that the overboard bleed, i.e. ow extracted along
the compression, is 0.83 kg/s or about 1% of the inlet ow. Only the
static conditions out of the compressor were given; hence, the
compressor outlet temperature and pressure ratio are computed
based on an assumption of the outlet Mach number. An outlet Mach
number of 0.25 is assumed, which according to Walsh and Fletcher
[10] is an ideal gure for the compressor outlet.
When applied to marine applications, the LM2500 gas turbine
runs on a light distillate fuel with a lower heating value (LHV) of
18 400 Btu/lb, corresponding to 42.798 MJ/kg. For the modelling,
a fuel composition reecting typical gures for Marine Gas Oil
(MGO) is used.
3.2. Simple model

3.1. Manufacturer data


The gas turbine selected for study is the LM2500 manufactured by General Electric. This machine, being an upgraded version
of the LM2500, is commonly used for marine applications. It is an
aero-derivative gas turbine originally derived from the CF6 family
of aircraft engines used on wide body aircraft. The LM2500 gas
turbine is a two-shaft design, with gas generator mechanically
uncoupled from the power turbine, enabling the power turbine to
operate at a continuous speed of either 3600 or 3000 rpm,
regardless of the speed of the gas generator. It features a 17-stage
axial compressor with the rst six stages provided with variable
guide vanes. The compressor turbine and power turbine feature
two and six stages, respectively; see Fig. 2.
Performance data for design and off-design conditions for the
LM2500 gas turbine were provided by General Electric. The

In the simple model, the gas turbine is assumed to consist of


a compressor, combustion chamber, compressor turbine and power
turbine; see Fig. 3. No inlet or exhaust pressure losses, cooling
ows, or bleed are modelled. A combustion chamber pressure loss
of 3% and a mechanical efciency of 99% are assumed. Inlet mass
ow, pressure ratio and generator efciency are dened according
to the specications (Table 1), and the compressor and turbine
isentropic efciencies and combustor outlet temperature are
adjusted to meet the specications for thermal efciency and
exhaust temperature. The results are given in Table 2.
3.3. Complex model
In the complex model, the gas turbine is assumed to consist of
an inlet, compressor, combustion chamber, compressor turbine,
power turbine, and exhaust. Similarly as for the simple model,
a combustion chamber pressure loss of 3% and a mechanical efciency of 99% are assumed. Inlet mass ow, pressure ratio,

CC
C

Fig. 2. The LM2500 gas turbine (by courtesy of General Electric).

CT

PT

Fig. 3. Schematic gure of the simple gas turbine model.

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492


Table 2
Comparison of full-load performance for the LM2500 gas turbine

hCis [%]
hTis [%]
COT [ C]
TIT [ C]
PE [kW]
Texh [ C]
hth [%]

GE data

Simple model

Complex model

31 207
533.8
37.70

83.8
86.5
1236
1236
31 207
532.2
37.70

85
88
1310
1250
31 207
531.8
37.75

generator efciency, and inlet and outlet pressure drops are dened
according to the specications. Bleed is simulated by extracting 1%
of the ow after the compressor.
Cooling is used to enable higher turbine entry temperatures
than the maximum allowable metal temperature. This is accomplished by bleeding off a part of the compressed air which then
passes through cooling passages inside the blades. Also air needs to
be taken off for sealing, in order to hinder the expanding gases from
penetrating the disk system. Effects of cooling are simulated by
bleeding off 10% of the air ow after the compressor. Depending on
where expansion work takes place, this air is mixed with the hot
gases at different points along the expansion. To simulate nozzle
guide vane cooling, 8% of the bleed is injected before the
compressor turbine, and the remaining 2% is injected before the
power turbine in order to simulate the use of rotor cooling and
seals. From a modelling point of view, the cooling air ows are
mixed with the exhaust gases before the turbines. When running in
part-load the percentage of air taken off for cooling is retained.
A schematic gure of the engine model is shown in Fig. 4. In
practice, for modern gas turbines, the turbine cooling system
generally is more complex than modelled here; air is usually
extracted from the compressor at a number of places, matching the
pressure condition where it should be injected in the turbine.
Inlet mass ow, pressure ratio, generator efciency, and inlet
and exhaust pressure losses are dened according to the specications, and the compressor and turbine isentropic efciencies and
combustor outlet temperature are adjusted to meet the specications for thermal efciency and exhaust temperature. Due to the
losses in the turbine associated with cooling, in practice, the isentropic efciency of the compressor turbine would be lower than
that of the power turbine. Quantifying the detrimental effect of
cooling on the turbine isentropic efciency is, however, beyond the
scope of the current work. Since the overall performance is of
primary interest here, for simplicity, the same efciency gure,
representing an average value during the whole expansion process,
is selected. The results are given in Table 2.
3.4. Comparison of model results
From Table 2 it can be concluded that both the simple and
complex models agree very well with manufacturer data, i.e. for
a given power output the exhaust temperature and the thermal

1487

efciency for both models are very close to the manufacturer data.
As a consequence of cooling ows, the turbine inlet temperature
(TIT) is lower than the combustor outlet temperature (COT) for the
complex model, whereas these temperatures are equal for the
simple model. Since various losses are excluded from the simple
model, the isentropic efciencies for the compressor and the
turbines need to be lower.
4. Part-load performance
In this section the models for part-load performance estimations are described.
4.1. Turbine constant
The mass ow coefcient, 4, is the index of total mass ow
entering the nozzle throat of an expansion, according to the relationship for a compressible, isentropic ow in a single nozzle, and it
is dened as [11]:

f q
P
V

(4)

In a multistage turbine the expansion process can be divided into


a number of segments. For each segment a nozzle analogy may be
developed which treats each segment as if it were a single nozzle
[11]. This analogy is known as Stodolas Ellipse, and it states that for
each nozzle the following relationship is valid [11,12]:

s

2
P
fin f 1  out
Pin

(5)

This proportionality is mathematically valid only for an innite


number of stages, but it has been shown to be empirically valid
down to eight 50% reaction stages in a steam turbine [11]. It has
been assumed that all nozzle ow areas remain constant.
By applying Eq. (5) to a whole turbine and assuming ideal gas, an
expression for the turbine constant (i.e. the constant that governs
the relationship between the mass ow coefcient and the
expression on the right-hand side of the proportionality sign of Eq.
(5)) can be obtained:

p
W Tin

CT q
2  P2
Pin
out

(6)

This expression for the turbine constant is used in the subsequent


calculations. It is recognised from Eq. (6) that for outlet pressure
approaching zero, the turbine constant becomes equal to the mass
ow coefcient (transformed using the perfect gas law). That the
ow coefcient is equal to a constant when the back pressure
becomes close to zero, is empirically known to be true for uncontrolled multistage expansion to high vacuum [11].
4.2. Component maps

CC
I

CT

1% 10%
8%
Bleed
Cooling flows

PT

2%

Fig. 4. Schematic gure of the complex gas turbine model.

In order to be successful with the prediction of the off-design


performance using component maps, it is essential that suitable
maps are used. The best option is of course to use the component
maps which reect the actual components of the machine under
consideration. However, manufacturers rarely share such information. The analyses conducted by the research community,
therefore, often need to be based on general map characteristics
available in the public domain. Ideally a map should be selected

1488

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492

35

Pressure ratio [-]

30
25
20
n=0.10

15

n=0.9

10

n=0.8

5
n=0.4
0
0

10

20

n=0.7

n=0.5 n=0.6

30

40

Corrected mass flow [kg/s]

10

n=0.4, 0.5, 0.6, 0.7, 0.8, 0.9, 1.0

6
50

60

70

80

90

100

Pressure ratio [-]

Corrected mass flow [kg/s]


Fig. 5. The map for the compressor (scaled with respect to the design data used here)
showing pressure ratio and corrected mass ow. The rotational speed, n, is given as
a relative corrected gure, i.e. the ratio of the corrected speed to the corrected speed at
design.

Fig. 7. The map for the compressor turbine (scaled with respect to the design data
used here) showing corrected mass ow and pressure ratio. The rotational speed, n, is
given as a relative corrected gure, i.e. the ratio of the corrected speed to the corrected
speed at design.

with characteristics similar to those of the component requiring the


off-design behaviour estimate.
The compressor map used here, compiled from data in Stevenson and Saravanamuttoo [13], is intended to reect the compressor
of the LM2500 gas turbine. The map used for the compressor
turbine reects a high work, low aspect ratio axial turbine [14], and
the power turbine is modelled using a map of an axial turbine
derived from Serovy [15]. The component maps used are depicted
in Figs. 510, where the corrected mass ow and corrected speed
given in the maps are dened as follows:

When the stators are rotated away from the axial direction, the
axial velocity and mass ow are decreased for a given rotational
speed. At low rotational speeds, this delays stalling of the rst few
stages and choking in the last stage, hence improving the surge
margin. Whether the compressor map used here includes any
variable geometry within its characteristics is unknown.

Wcorr

s
R T
W
Rstd Tstd

(7)

P
Pstd

N
Ncorr s
R T
Rstd Tstd

(8)

The parameter R is the gas constant and std refers to standard


atmosphere (T 15  C, P 101.325 kPa).
As pointed out in Section 3.1, the compressor of the LM2500
gas turbine is provided with a number of stages with variable
geometry. In general, compressors often are provided with several
rows of variable stators at the front of the compressor in order to
permit large pressure ratios to be achieved in a single-shaft [10,16].

4.3. Generator efciency


According to The ngstrom Laboratory, Sweden (J. Santiago, The
ngstrom Laboratory, Uppsala, Sweden, 2007, private communication) and theory about generator losses [17], one fraction of the
generator loss (i.e. the copper losses, which are produced in the
winding of the stator) is dependent on the load squared, whereas
the rest of the loss is independent of the load. The efciency of the
generator is usually dened as the ratio of the electrical power
output to the mechanical power input. Assuming that the copper
losses constitute a fraction, FCU, of the total losses in design, the
following equation for the efciency in part-load as a function of
this fraction, the efciency at design, hD, and the load can be
derived:

hPL

(9)

In order to nd a reasonable gure for FCU, results for the generator


of the LM2500 gas turbine are considered. This is a BRUSH DAX
2-pole turbogenerator (BDAX 71-193ER 60 Hz, 13.8 kV, 0.8PF),

100

100

90

90

80
n=0.7

70
n=0.4

60

n=0.8

Isentropic efficiency [%]

Isentropic efficiency [%]

LoadhD
i
h
LoadhD 1  hD 1  FCU FCU Load2

n=0.9 n=1.0

n=0.6
n=0.5

50
40
30
20
10

80
70

n=0.4

60

n=0.5 n=0.6

n=0.7

n=1.0
n=0.9
n=0.8

50
40
30
20
10

0
0

20

40

60

80

100

Corrected mass flow [kg/s]


Fig. 6. The map for the compressor (scaled with respect to the design data used here)
showing isentropic efciency and corrected mass ow. The rotational speed, n, is given
as a relative corrected gure, i.e. the ratio of the corrected speed to the corrected speed
at design.

Pressure ratio [-]


Fig. 8. The map for the compressor turbine (scaled with respect to the design data
used here) showing isentropic efciency and pressure ratio. The rotational speed, n, is
given as a relative corrected gure, i.e. the ratio of the corrected speed to the corrected
speed at design.

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492


100

n=0.7 n=0.8 n=0.9 n=1.0

n=0.4 n=0.5 n=0.6

35

96

30

Efficiency [%]

Corrected mass flow [kg/s]

40

1489

25
20
15
10
5

92

GE
Formula

88

84

0
0

80

Pressure ratio [-]


Fig. 9. The map for the power turbine (scaled with respect to the design data used
here) showing corrected mass ow and pressure ratio. The rotational speed, n, is given
as a relative corrected gure, i.e. the ratio of the corrected speed to the corrected speed
at design.

which can run on either 3000 rpm, corresponding to a frequency of


50 Hz, or 3600 rpm, corresponding to a frequency of 60 Hz. Efciency gures for this generator for the load ranging from 10% to
100% are shown in Fig. 11. For FCU equal to 0.43 very good agreement between Eq. (9) and the data is found. Eq. (9) with FCU equal to
0.43 is used in this paper for part-load modelling of the generator.

5. Results and discussion


Using DNA, performance results for the models are obtained for
the load ranging from 10% to 100%. For each design model, both
turbine constants and maps are employed, resulting in four cases in
total. The model results are compared with the manufacturer data;
see Figs. 1215. For the models using maps, the agreement with
manufacturer data is very good except for the lightest loads. No
results are displayed for loads below 20%, because for these
conditions the operating points of the power turbine become
located outside the data range included in the map. A contributing
source for the deviation at low powers might be the uncertainty
about operation and modelling of variable geometry for the
compressor (see Sections 3.1 and 4.2). In general, the results
suggest that the complex model gives slightly better agreement
with manufacturer data at part-load, but the difference is indeed
small.

20

40

60

80

100

Load [%]
Fig. 11. Generator efciency versus load for manufacturer data and results of formula.

In order to understand the part-load behaviour of the aeroderivative gas turbine, the effect of operating two turbines in series
is considered, following the outline of Cohen et al. [16]. The nondimensional ow at the exit of the gas generator (index 4) is
a function of the non-dimensional ow at the inlet to the
compressor turbine (index 3), and the compressor turbine pressure
ratio and temperature ratio; see Eq. (10). In turn, the temperature
ratio is a function of the efciency. The efciency could be obtained
from the turbine characteristics, but, in practice, the variation with
rotational speed for a given pressure ratio is small, particularly over
the restricted range of operation of the compressor turbine.
Moreover, the effect on the non-dimensional speed becomes even
smaller as it is dependent on the square root of the temperature
ratio. If a mean value of efciency at any given pressure ratio is
used, the non-dimensional ow out becomes a function of the nondimensional ow in and the pressure ratio. Assuming this, a single
curve representing the compressor turbine outlet ow characteristics can readily be obtained for points on the single curve of the
inlet ow characteristics; see dotted curve in Fig. 16.

p
p s
T4
W T3 P3 T4

P4
P3 P4 T3

(10)

25

Pressure ratio [-]

20

100

Isentropic efficiency [%]

90

n=1.0

80
n=0.4

70

n=0.5 n=0.6

n=0.7

n=0.8 n=0.9

60

15

10

50
40

30
20
0

10

0
0

20

Pressure ratio [-]


GE data

Fig. 10. The map for the power turbine (scaled with respect to the design data used
here) showing isentropic efciency and pressure ratio. The rotational speed, n, is given
as a relative corrected gure, i.e. the ratio of the corrected speed to the corrected speed
at design.

40

60

80

Load [%]

Complex
& maps
Simple & CT

Complex
& CT
Simple & maps

Fig. 12. Compressor pressure ratio versus load.

100

1490

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492


100

600

90

Exhaust temperature [gr C]

Inlet mas flow [kg/s]

80
70
60
50
40
30
20

500

400

300

200

100

10
0

0
0

20

40

60

80

100

20

Load [%]

60

80

100

Load [%]
Complex
& CT

Complex
& maps

GE data

40

Simple & CT

GE data

Simple & maps

Complex
& maps
Simple & CT

Complex
& CT
Simple & maps

Fig. 13. Inlet mass ow versus load.

Fig. 15. Exhaust temperature versus load.

From Fig. 16 it can be seen that the requirement for ow compatibility between the compressor turbine and the power turbine
places a major restriction on the operation of the compressor
turbine. The maximum pressure ratio across the compressor
turbine is controlled by the choking of the power turbine (point a),
and at all times the pressure ratio is controlled by the swallowing
capacity of the power turbine (e.g. point b). A further consequence
of the xed relationship between the turbine pressure ratios is that
it is possible to plot the compressor turbine pressure ratio versus
the compressor pressure ratio. From this gure it follows that the
compressor turbine pressure ratio increases with the compressor
pressure ratio until the power turbine becomes choked, at which it
becomes constant.
At full-load, often both the compressor turbine rst-stage nozzle
and the power turbine nozzle operate at or near choked ow
condition [18]. In the case studied here, the power turbine is
operated close to choked condition, and the non-dimensional mass
ow decreases rst slowly and then with increasing rate with

decreasing pressure ratio. The compressor turbine is operated


within a narrow range of pressure ratios, and at all times it stays
choked. Considering the model with maps, the turbine ow characteristics (i.e. whether the ows are choked or not) at full-load are
determined by the location of the design points within the turbine
maps. That is, by choosing other design point locations within the
turbine maps, giving other ow characteristics, a different partload performance would be obtained. With respect to the model
with turbine constants, the turbine ow characteristics at the
design point are governed by the pressure ratio and the value of the
turbine constant. For illustration, the actual operating points for the
turbines for the simple model with turbine constants are shown in
Fig. 17.
The results shown in Figs. 12 and 13 suggest that turbine
constants are sufcient to predict the mass ow and pressure ratio
characteristics during part-load conditions. In terms of exhaust
temperature and thermal efciency, on the contrary, the results
suggest that the model with turbine constants does not predict
these performances adequately at light loads (see Figs. 14 and 15).
This is because the isentropic efciencies are assumed constant.
Neglecting the inuence of load on the isentropic efciency for the
compressor has no signicant effect, because even though the
pressure ratio and non-dimensional mass ow have decreased
signicantly for the lowest loads, the operating point is still located
in a region of the map with high efciency.

40

Thermal efficiency [%]

35
30
25
20
15
10
5
0
0

20

40

60

80

100

Load [%]
GE data

Complex
& maps
Simple & CT

Complex
& CT
Simple & maps

Fig. 14. Thermal efciency versus load.

Fig. 16. Operation of turbines in series.

800
600
400
200
0
1

Non-dimensional flow

Non-dimensional flow

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492

1491

800
600
400
200
0
1

P3/P4

P4/Pa

Fig. 17. Actual operating points for the two turbines for the simple model with turbine constants. Marks indicate operating points between 10% and 100% load.

Since the non-dimensional ow for the compressor turbine is


constant and the pressure ratio is only slightly decreased, the
isentropic efciency for the compressor turbine also is more or less
unaffected by the load. What gives rise to the deviations at light
loads when using turbine constants, is simply that the inuence on
the isentropic efciency for the power turbine is neglected. As the
load is decreased, the pressure ratio for the power turbine is
decreased signicantly (since the pressure ratio across the
compressor turbine stays nearly constant), affecting the efciency
essentially.
In the models with turbine constants and constant efciencies
for the turbomachinery, the unreasonably high efciency for the
power turbine will result in a too low exhaust temperature for light
loads, which explains the deviation shown in Fig. 15. Moreover,
neglecting the degradation in component efciencies in part-load,
results in an over-estimation of the thermal efciency (see Fig. 14,
a consequence which is partly counter-acted by a small over-estimation of the turbine inlet temperature).

An essential advantage with the approach of using maps is that


agreement always can be found, though it may sometimes involve
substantial effort. If the part-load performance achieved is not
satisfactory with one set of maps, these can be replaced by others.
Another parameter affecting the part-load performance which can
be varied, is the location of the design point within the maps. The
procedure is discussed in Kurzke [19]. If these measures would not
be sufcient, the map characteristics can also be changed manually
in order to attain satisfactory agreement. With the use of turbine
constants no actions can be taken to change the part-load
performance.
The methodologies discussed here for predicting the part-load
performance of aero-derivative gas turbines are of course general
and can be applied to any gas turbine type. However, it is not sure
that the conclusions drawn about the applicability of using turbine
constants would be the same for other gas turbine types.

6. Conclusions

The authors would like to express their gratitude to Joseph


Peters and Danny Hutchison at General Electric, USA, for providing
information and performance data on the LM2500 gas turbine,
and Joachim Kurzke (developer of GasTurb), Germany, for
providing information on the component maps provided with
GasTurb 10. Furthermore, Juan de Santiago and Mats Lejon at The
ngstrom Laboratory, Sweden, are thanked for their guidance on
generator performance. The funding from the Danish Center for
Maritime Technology (DCMT) is acknowledged.

In this paper different methodologies for predicting the partload performance of an aero-derivative gas turbine have been
investigated. Model results have been compared with manufacturer data for the LM2500 gas turbine.
With respect to the design models, a simple model, featuring
a compressor, combustor and turbines, results in equally good
performance prediction (in terms of thermal efciency and exhaust
temperature) as a more complex model including also bleed,
cooling ows and pressure losses. The major drawbacks of the
simple model are that unrealistically poor component efciencies
need to be used, and the lack of cooling modelling, results in
unrealistic temperature levels within the cycle. Having a reasonable
estimation of the maximum cycle temperature can, for example, be
important if the formation of pollutant emissions should be
considered. Moreover, the complex model shows slightly better
agreement with manufacturer data in part-load; however, the
difference is indeed small.
Considering the two different methodologies for predicting the
part-load performance maps and turbine constants the results
suggest that the mass ow and pressure ratio characteristics can be
well predicted with both methods. Down to about 6070% load also
the thermal efciency and exhaust temperature can be well predicted with both methods, but below that point the use of turbine
constants and assuming constant efciencies for turbomachinery,
results in an under-prediction of the exhaust temperature and overprediction of the thermal efciency; the lower the load, the higher
the deviations. For instance, at 30% load, the model with turbine
constants gives a thermal efciency 8.5% higher and an exhaust
temperature 12% lower than the manufacturer data. For the same
operating point, the model with maps provides agreement in
thermal efciency and an exhaust temperature 1.5% lower than
manufacturer data.

Acknowledgements

References
[1] Zhu P, Saravanamuttoo HIH. Simulation of an advanced twin-spool Industrial
gas turbine. J Eng Gas Turbines Power 1992;114:1805.
[2] Pathak M, Suresh Kumar P, Saha UK. Prediction of off-design performance
characteristics of a gas turbine cycle using matching technique. Int J Turbo Jet
Engines 2005;22:10319.
[3] Najjar YSH, Lamfon NJ, Akyurt M. Modelling and simulation of gas turbine
engines, part 1: design point performance at ISO conditions. Int J Power
Energy Syst 1995;15(2):617.
[4] Deidewig F, Dopelheuer A. Studies on NOx-emissions of SST engine concepts.
In: AGARD PEP symposium on advanced aero-engine concepts and controls,
Seattle, USA, AGARD-CP-572, 1995.
[5] Elmegaard B, Houbak N. DNA a general energy system simulation tool. In:
Proceedings of SIMS 2005, Trondheim, Norway, 2005.
[6] Elmegaard B. Simulation of boiler dynamics development, evaluation and
application of general energy system simulation tool. Ph.D. thesis, Technical
University of Denmark, Department of Mechanical Engineering, Denmark,
1999.
[7] Perstrup C. Analysis of power plant installation based on network theory (in
Danish). M.Sc. thesis, Technical University of Denmark, Laboratory of Energetics, Denmark; 1991.
[8] Kurzke J. Component map collection 2, Compressor and turbine maps for gas
turbine performance computer programs. Germany; 2004, http://www.
gasturb.de/.
[9] Kurzke J. Preparing compressor maps for gas turbine performance modelling
with SmoothC, version 8. Germany; 2002, http://www.gasturb.de/.
[10] Walsh PP, Fletcher P. Gas turbine performance. Malden, USA: Blackwell
Science Ltd.; 1998.
[11] Cooke DH. On prediction of off-design multistage turbine pressures by Stodolas ellipse. J Eng Gas Turbines Power 1985;107:596606.

1492

F. Haglind, B. Elmegaard / Energy 34 (2009) 14841492

[12] Csanady GT. Theory of turbomachines. New York, USA: McGraw-Hill; 1964.
[13] Stevenson JD, Saravanamuttoo HIH. Simulating indirect thrust measurement
methods for high bypass turbofans. ASME Paper No. 93-GT-335, 1993.
[14] Stabe RG, Ehitney WJ, Moft TP. Performance of high-work low aspect ratio
turbine tested with a realistic inlet radial temperature prole. NASA TM
83665, NASA Headquarters, Suite 5K39, Washington, DC 20546-0001, USA;
1984.
[15] Serovy GK. Compressor and turbine prediction system development lessons
from thirty years of history. AGARD LS 83; 1976.

[16] Cohen H, Rogers GFC, Saravanamutto HIH. Gas turbine theory. 4th ed. Essex,
UK: Longman Group Limited; 1996.
[17] Wildi T. Electrical machines, drives, and power systems. 6th ed. New Jersey,
USA: Pearson Education, Inc.; 2006.
[18] Kurz R. Gas turbine performance. Solar turbines incorporated, San Diego,
California, USA; 2005, http://turbolab.tamu.edu/pubs/Turbo34/T34pg131.pdf.
[19] Kurzke J. How to create a performance model of a gas turbine from a limited
amount of information. In: Proceedings of ASME Turbo Expo 2005, RenoTahoe, Nevada, USA, vol. 1, 2005, pp. 14553.

S-ar putea să vă placă și