Taurus 70

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Case Studies in Thermal Engineering 26 (2021) 101206

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


journal homepage: www.elsevier.com/locate/csite

High ambient temperature effects on the performance of a gas


turbine-based cogeneration system with supplementary fire in a
tropical climate
Daniel Armando Pinilla Fernandez a, Blanca Foliaco a, Ricardo Vasquez Padilla b,
Antonio Bula a, Arturo Gonzalez-Quiroga a, *
a
UREMA Research Unit, Department of Mechanical Engineering, Universidad Del Norte, Barranquilla, Colombia
b
School of Environment, Science and Engineering, Southern Cross University, Lismore, NSW, 2480, Australia

H I G H L I G H T S

• High ambient temperature affects the performance of natural gas-based cogeneration.


• Validated ASPENHYSYS® simulation compare operation with that at ISO conditions.
• Gas turbine efficiency drops 0.06% per ◦ C rise in ambient temperature above 15 ◦ C.
• (Un)fired cogeneration increases power output 5%–17% compared to gas-turbine-alone.
• High ambient humidity affects cogeneration performance and limits inlet air cooling.

A R T I C L E I N F O A B S T R A C T

Keywords: High ambient temperature negatively affects gas turbine performance, especially in a tropical
Power generation climate. Cogeneration improves fuel utilization by taking advantage of the energy discharged as
Heat recovery waste heat in the exhaust gases. This case study assesses the effects of high ambient temperature
Ambient temperature
on the performance of a natural gas-based cogeneration plant in Barranquilla, Colombia, a
Electricity and steam
Overall efficiency
location with a hot and humid tropical climate throughout the year with an annual average
Tropical climates temperature of 27.4 ◦ C. The cogeneration plant encompasses gas and vapor turbine generation,
supplementary fire, waste heat recovery, and process heat exchange. Validated ASPENHYSYS®
simulation allows comparing gas-turbine-alone indicators with those at ISO conditions, i.e., 15 ◦ C
and 101.3 kPa. Ambient temperature reduces gas turbine power output by up to 22% and de­
creases thermal efficiency by around 0.06% for every ◦ C rise above ISO conditions. Cogeneration
with and without supplementary fire increases power output by 17% and 5% compared to gas-
turbine-alone operation. The energy utilization factor increases by 27–37% without supplemen­
tary fire and above 37% with supplementary fire. Results give insight into the challenges of
cogeneration plants in a tropical climate. Further studies should include the effects of high hu­
midity on power plant performance and the potential benefits of cooling inlet air.

* Corresponding author.
E-mail address: arturoq@uninorte.edu.co (A. Gonzalez-Quiroga).

https://doi.org/10.1016/j.csite.2021.101206
Received 5 April 2021; Received in revised form 24 June 2021; Accepted 28 June 2021
Available online 1 July 2021
2214-157X/© 2021 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

Nomenclature

AFR Air-fuel ratio


ap Approach point
Amb Ambient
BD Boiler blowdown
CHP Combined heat and power
EAF Excess air factor
EUF Energy utilization factor
FM Fired mode
GHG Greenhouse gases
GT Gas turbine
HE Heat exchanger
HRSG Heat recovery steam generator
LHV Lower heating value
ṁi Mass flow rate, MJ kg− 1
NG Natural gas
η Thermal efficiency
pp Pinch point
PR Pressure ratio
Q Heat flow rate, W
S Standard deviation
SM Simulation model
stch Stoichiometric
ΔT Temperature difference, ◦ C
T Temperature, ◦ C
UFM Unfired mode
VT Vapor turbine
V̇ Volumetric flow rate, Nm3 s− 1
Ẇ Power output, W

1. Introduction

Current trends indicate significant energy demand increments, especially in emerging economies, where the preferred option is fossil fuel-
based energy. The growing interest in cleaner energy sources brings new challenges and opportunities for fossil fuel-based power generation
development. Natural gas is the less polluting fossil fuel concerning greenhouse gas (GHG) emissions and the most flexible for operating under
variable loads. Thermal efficiency improvement could be responsible for more than 40% of GHG emissions reduction related to power gen­
eration in the next 20 years [1,2]. Gas turbines are versatile equipment capable of supplying varying energy requirements. Turbine design
information is essential for a judicious equipment selection. Manufacturers typically provide turbine capacity and design information for full
load operation under ISO conditions, i.e., 15 ◦ C and 101.3 kPa. However, changes in environmental conditions affect turbine power generation
capacity and efficiency. Studies show that ambient temperature is an influential factor in gas turbine performance [3].
An ambient temperature of 37 ◦ C caused an average power loss of 17%, accompanied by an efficiency drop of 2.2% compared to the
gas turbine design value [3]. Actual data shows that the gas turbine lost 0.1% in thermal efficiency and 1.47 MW of its power output for
every ◦ C rise in ambient temperature above ISO conditions [4]. Likewise, a gas turbine modeling study shows reductions of around
0.06% in thermal efficiency and 0.12 MW of power output for every ◦ C rise in ambient temperature above ISO conditions [5]. A gas
turbine simulation model indicates that efficiency decreased by 0.03–0.07% for every ◦ C rise in ambient temperature [6]. Additionally,
power output decreases by 5–21% and thermal efficiency by more than 1.5% during hot seasons compared to ISO conditions operation
[7]. The adverse effects of high ambient temperature are more evident for power generation facilities located in hot regions, such as the
north Caribbean region of Colombia, where the humid tropical climate predominates [8].
Mitigating the harmful effects of high ambient temperature is imperative in hot regions. Reducing the inlet air temperature or
recovering residual heat in the exhaust gases improves power plant performance [7,9]. Inlet air cooling technologies such as evap­
orative cooling, high pressure fogging, and absorption chiller cooling; in a power plant produce 1–15% higher power output than that
without inlet air cooling [10–13]. Combined heat and power (CHP) or cogeneration is a well-known solution to counteract ambient
temperature effects. These systems use thermal energy in the turbine exhaust gases to produce, for instance, superheated steam. Steam
represents an energy vector that can supply both thermal and mechanical energy. Thermal energy transfer through a heat exchanger
can supply process energy requirements. A steam turbine can supply mechanical power for driving rotational equipment or electric
generators [9,14]. The addition of combined cycles in power plants provides backup power during peak demands. A cogeneration unit
consumes less fuel than separate conventional technologies to generate a certain amount of electrical and thermal energy, thus
improving fuel utilization and reducing GHG emissions per power generation unit [2,9].

2
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

Several CHP cycle studies show performance improvement when compared with the gas-turbine-alone operation. A design study
developed models of different power generation cycle configurations, including compressor-gas turbine systems and two CHP cycle
variants [15]. The compressor-gas turbine system reached a fuel utilization efficiency of 38%. On the other hand, CHP shows a two-fold
increase in fuel utilization. A CHP simulation shows power output increments of more than 50% when cogeneration is included [16].
Results indicate that gas and steam turbines reach power outputs of 43.6 and 28.6 MW, respectively, with variations of less than 10% of
theoretical values. The study of a small combined gas-steam power generation cycle [17] shows results similar to those in Ref. [16]. Power
output increases up to 30%, and cycle efficiency increases from 32 to 42% when the power plant includes gas and steam turbines. A
simulation study on the influence of operating conditions in a CHP cycle showed that the combined cycle output power decreases by 0.26
MW, and efficiency drops by 0.015–0.04% for every ◦ C rise of ambient temperature above ISO conditions [6]. Ambient temperature
effects are lower on a cogeneration power plant compared to a gas-turbine alone power plant. That makes the CHP cycles an attractive
technology for mitigating ambient temperature impact on gas turbine-based power generation in tropical climates.
Some authors carried out parametric studies and modifications to the Brayton cycle through ASPENHYSYS®. A study of the performance of
a microturbine shows the efficiency improved up to 83.3% by implementing technologies such as exhaust gas recirculated, steam injection, and
humid air turbine [18]. Two papers study a cogeneration system in tropical climates: a heat recovery steam generator (HRSG) and a new
operating strategy, exhaust gas recycle plus inlet guide vanes technology, in a 393 MW plant [19,20]. Another paper analyzes the gas
composition, the air intake cooling, and steam injection in a combined cycle [21]. However, there was no discussion about the plant’s capacity
regarding cogeneration and heat production for industrial processes. In dry climates [22], a flare gas recovery system in a combined
Brayton-Rankine cycle was evaluated through four different ways to recover the energy of the flare gas with an ASPENHYSYS® v10 Simulation.
Their work focused on economic analysis. A simple gas turbine cycle simulation with variation in fuel type achieved an efficiency of 31.9%
without validation [23]. Exergy analysis of a thermal plant in Nigeria was recently performed using ASPENHYSYS® 8.8; however, there is no
detailed information about any improvement [24]. A paper related to applying a cogeneration system in the Kraft industry studied an
ASPENPLUS® simulation [25]. The study focused on the cost optimization of the system employing a genetic algorithm method.
Table 1 summarizes recent works regarding power generation simulations using ASPENHYSYS®. Efficiency and inlet air temperature
were used to compare the achieved system performance. None of the published investigations addressed the simulation in ASPENHYSYS®
of the gas turbine and cogeneration, with and without supplementary fire for the paper industry in a tropical climate. The validation
against operation data and the simulation setup are valuable contributions to improving gas turbine-based power plants.
The new gas fields in the north Caribbean and continuous growing energy demand increased natural gas importance for Colombia
[26]. The currently installed natural gas-based electricity generation capacity in Colombia amounts to around 1.7 GW, corresponding
to 10% of the total power generation available. In Barranquilla, the thermal power plant installed capacity amounts to about 1.2 GW,
with a significant gas turbine contribution [27]. These plants withstand the influence of the demanding environmental conditions in
Barranquilla, with 32 ◦ C average ambient temperature and variations in the range 27–37 ◦ C [8]. This paper focuses on the effect of
ambient temperature in Barranquilla on a cogeneration power plant. A Solar Turbines Taurus™ T70 gas turbine, a waste heat recovery
boiler, and a vapor turbine are the main stages. We generated a simulation model of the cogeneration plant in the process software
ASPENHYSYS®. Both manufacturer information and experimental data were used for setting the simulation up and for validation. The
analysis starts by comparing ISO and actual performance indicators caused by local ambient conditions for the conventional gas
turbine model. Then we evaluate performance improvement after adding the cogeneration system with and without supplementary
fire. Finally, we develop an outlook to discuss the potential use of cogeneration systems in future research or industrial applications.

2. Methodology

2.1. The cogeneration plant

Here we use a cogeneration plant model created in ASPENHYSYS®. The plant locates in Barranquilla, Colombia, a humid tropical
zone of saline environment. The plant consists of three main pieces of equipment: a gas turbine, a waste heat recovery boiler, and a
steam turbine. Oil & Gas leader companies use this software to model, design, and optimize refining and petrochemicals processes.
Fig. 1 shows a schematic diagram of the cogeneration plant. The diagram illustrates the materials and energy streams connections

Table 1
Comparative table for recent simulation studies in gas turbines and power generation plants using ASPENHYSYS®. ηthermal is the thermal efficiency,
and Tair the turbine inlet temperature for every system.
Reference Year Country System Validation ηthermal Tair, ◦ C

[24] 2021 Nigeria Rankine modified – Exergetic analysis Yes 33.19% 30


[23] 2020 Iraq Gas turbine simple cycle – Type of fuel No 0.05%/◦ C 15
[22] 2020 Iran Combined cycle + Flare Gas Recovery No – 25
[21] 2020 Indonesia Combined cycle + HRSG – Effect of operative parameters Yes 0.07%/◦ C − 15–15
[20] 2018 Singapore Combined cycle + HRSG No 56.14% 15
[19] 2018 Singapore Combined cycle + Exhaust gas recycled + Inlet guide vanes No 0.098%/% plant 15
load
[18] 2015 United Microturbine + Exhaust Gas Recirculated/Steam Injection/Humid Air Yes 81.2% 15
Kingdom Turbine
[25] 2013 Iran Combined Cooling, Heating and Power + HRSG + optimal design in a No – 25
Kraft process

3
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

between the three main components: the gas turbine, the waste heat recovery boiler, and the process heat and steam expansion section.
Compressed ambient air and natural gas enter the gas turbine for combustion and power generation. High-temperature exhaust gases
flow through the waste heat boiler to generate superheated steam. The boiler includes supplementary fire for cases in which the plant
requires additional steam generation to supply increased power or heat demand. When the supplementary fire is on, the hot gases
deviate to the supplementary burner. Inside the burner, the remaining oxygen in the exhaust gases reacts with the supplementary
natural gas, which increases the available energy and, therefore, the steam mass flow rate in the boiler. Finally, the superheated steam
supplies either process heat for paper drying or produces shaft power in a steam turbine. The outlet streams consist of medium-pressure
steam at the exit of the heating process and wet steam from the steam turbine.
The plant operator provided historical operating data of each subsystem for setting the simulation up. This data is essential to
identify the expected relation between operational variables and to validate the model. Figs. S1, S2, and S3 in the Supplementary
Material show the observed behavior of power output in the gas turbine and steam production in the waste heat boiler. The gas turbine
operation data shows that excessive ambient temperature negatively affects power output. Operation data also shows a direct rela­
tionship between steam production and the ambient temperature. Besides, there is a direct proportion between steam production and
supplementary fuel consumption.

2.2. Flow streams

For simulation purposes, ambient air consists of a mixture of N2, O2, Ar, CO2, and water vapor at a pressure of 101.3 kPa. The
absolute air humidity remains approximately constant, around 22.7 g of H2O per kg of dry air, corresponding to 30 ◦ C and 65% relative
humidity [8]. Table 2 shows the ambient air composition on a dry basis. The supply conditions of natural gas are 35 ◦ C and 2213 kPa.
Table 3 shows the monthly-updated natural gas composition supplied by the local natural gas transport company [28].
Liquid water at the waste heat boiler inlet is available at 116 ◦ C and 2103 kPa. We used the Peng-Robinson equation of state to
calculate the properties of ambient air, natural gas, and gas mixtures. The Peng-Robinson equation is suitable for hydrocarbons (e.g.,
methane, propane, ethane), light gases (e.g., N2, H2), and mixtures thereof [29]. Likewise, the STEAM NBS model was used to calculate
liquid water and steam properties [30].

2.3. Combustion reactions

Natural gas combustion reactions were specified in the ASPENHYSYS® reactions module as 100% conversion reactions [31].

2.4. Gas turbine

The Taurus™ T70 gas turbine is an axial flow turbomachine with two shafts. First, the multi-stage compressor draws ambient air
into the gas turbine and pressurizes it. Compressed air enters the combustion chamber while fuel is injected, mixed, and ignited during
the start cycle. Combustion will continue whenever both pressurized air and fuel flows are appropriate. The expansion of hot-
pressurized gas produced in the combustor drives the turbine. An electric generator converts shaft power into electric power [32].
Fig. 2a shows the gas turbine flowsheet in ASPENHYSYS®, consisting of a series of subsystems for compression, combustion,
expansion, and heat exchange. First, there is a two-stage compression zone comprising two compressors and an intercooler. Next, an
adiabatic reactor and a heat exchanger represent the combustion chamber and its heat losses. Then there is a two-stage expansion zone
encompassing two turbines for driving the compressors. Finally, a turbine and a heat exchanger represent power generation and heat
losses during expansion.
Ambient air flow in stream 3 passes through the compression zone with inter cooling, Qe, undergoing a pressure increase from
atmospheric conditions, 101.3 kPa, to 1313 kPa. Then, compressed air and fuel flow mix in the turbine combustor. At the exit of the
turbine combustor, exhaust gases in stream 8 flow through a heat exchanger, transferring Qp1 to the surroundings. Next, exhaust gases

Fig. 1. Schematic block diagram of the cogeneration power plant.

4
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

Table 2
Ambient air composition for simulating the cogene­
ration power plant in ASPENHYSYS®.
Component Mole fraction
1
N2 7.8 × 10−
1
O2 2.1 × 10−
3
Ar 9.0 × 10−
3
CO2 1.0 × 10−

Table 3
Natural gas composition for simulating the cogeneration power plant in ASPENHYSYS®.
Component Component Name Mole fraction
1
CH4 Methane 9.80 × 10−
3
C2H6 Ethane 2.53 × 10−
4
C3H8 Propane 6.00 × 10−
4
i-C4H10 i–Butane 2.00 × 10−
5
n-C4H10 n-Butane 8.00 × 10−
5
i-C5H12 i–Pentane 7.00 × 10−
5
n-C5H12 n–Pentane 2.00 × 10−
4
n-C6H14 n–Hexane 2.00 × 10−
2
N2 Nitrogen 1.50 × 10−
3
CO2 Carbon Dioxide 1.30 × 10−

flow into the two-stage expansion zone from stream 9 to stream 11. The energy extracted during expansion drives the compression
stage. Stream 11 represents the power turbine inlet point in which pressure decreases until 103.8 kPa. Finally, gases in stream 12 enter
into a heat exchanger representing the heat losses in the expansion zone, Qp2.
The gas turbine performance depends on the following input parameters: total pressure ratio, inter-cooling heat transfer, global
compression adiabatic efficiency, combustion heat losses, global expansion adiabatic efficiency, and expansion heat losses. Com­
parisons between model predictions and experimental data allow estimating the parameters listed in Table 4.

2.5. Waste heat boiler

The waste heat boiler works as a heat recovery steam generator and complying with the ASME Section I code [33]. The boiler
recovers energy from the exhaust gases and produces superheated steam. Fig. 2b shows the ASPENHYSYS® flowsheet used to simulate
the waste heat boiler, where a conversion reactor represents the supplementary fire. A set of three heat exchangers represent the
superheater, the evaporator, and the economizer. Finally, the two-phase separator corresponds to the steam drum. Energy balances for
the superheater and evaporator provide the exhaust gases temperature at the boiler outlet.
The boiler brings into thermal contact exhaust gases and pressurized water. The exhaust gases in stream 13 flow through the
combustion chamber, in which they mix with stream 14, an additional fuel inlet. Stream 14 sets the operation mode in the waste heat
boiler, i.e., fired or unfired. The combustion chamber and transport system are not thermally insulated, and Qp3 represents the cor­
responding heat losses. After leaving the combustion chamber, gases enter the multi-stage heat exchangers circuit. First, exhaust gases
flow through the superheater to generate superheated steam. Then, gases pass through the evaporator, turning saturated liquid water
into saturated steam. Subsequently, exhaust gases preheat pressurized liquid water in the economizer. The pressure drop of the exhaust
gases is negligible. The supplementary fire activates in case of increments in the plant steam requirement.
On the cold side, the operation starts when pressurized liquid water enters the economizer, in which pressure drops by about 35 kPa
according to experimental data provided by the plant operator. Pressurized liquid stream mixes with recycled saturated steam in a
steam drum separator at 2068 kPa. Around 3% wt. of saturated steam leaves the system as the boiler blowdown (BD). The remaining
fraction flows back to the steam drum. Saturated steam exiting the drum enters into the superheater, in which the final temperature
increase occurs until the stream becomes superheated at around 216 ◦ C.
The pressure drop in the evaporator, the superheater, and the pipes is negligible. The HRSG functioning depends on combustion
and transport heat loss, boiler blowdown, pinch, and approach points. Comparing model predictions and experimental data allows
estimating the parameters. Combustion and transport heat losses depend on the HRSG operation mode. For unfired operation, gas
transport heat losses amount to around 400 kW. The heat losses increase due to higher temperatures when the supplementary fire is on.
The increase in heat losses amounts to 1% of the supplementary natural gas lower heating value (LHV). Equation (1) gives combustion
and transport heat losses in the supplementary combustion chamber based on actual plant data. In Equation (1), ṁ14 represents the
supplementary natural gas mass flow rate.
Qp3 = 400 + 0.01 ṁ14 LHV (1)

A judicious selection of pinch and approach points prevents cross temperatures in the different heat exchangers and determines the
gas–steam temperature profiles. Equation (2) gives the pinch point, ΔTpp, where T18 is the gas temperature at the evaporator exit and

5
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

Fig. 2. (a) Gas turbine and (b) heat recovery steam generator (HRSG) flowsheets programmed in ASPENHYSYS® and its subsystems are as follows.
Fig. 2a includes low-pressure compression (Compressor LP), compression inter cooling (Qe), high-pressure compression (Compressor HP), com­
bustion chamber (Turbine Combustor), combustion chamber heat losses (Qp1), expansions for driving high and low-pressure compression (Expander
HP and Expander LP), gas turbine (Power Turbine), and expansion heat losses (Qp2). Fig. 2b includes supplementary combustion chamber (Sup­
plementary Fire), supplementary combustion and transport heat losses (Qp3), pressurized water heating (Economizer), pressurized water evapo­
ration phase (Evaporator), split flow to recirculation, and blowdown (TEE-100), liquid and vapor phase separator (Steam drum), and superheated
steam generation phase (Superheaters).

Ts the steam saturation temperature. The approach point, ΔTap, is calculated via Equation (3), where T21 is the temperature of
pressurized water exiting the economizer [14]. Based on this, pinch and approach points correspond to 13.9 ◦ C and 8.3 ◦ C, respec­
tively, for the HRSG operation.
ΔTpp = T18 − Ts (2)

ΔTap = Ts − T21 (3)

2.6. Process heat and steam expansion

The cogeneration system’s final stage encompasses heat exchangers and a steam turbine. The transferred heat covers plant re­
quirements, and the steam turbine transfers shaft work to a roller press.
Figure S4 in the Supplementary Material shows the ASPENHYSYS® flowsheet used to model heat transfer and vapor expansion.
This stage consists of two heat exchangers representing the heat losses to the environment during transport and expansion, and a vapor
turbine, producing shaft power. The steam turbine requires saturated steam at around 203 ◦ C. The outlet pressure amounts to

6
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

Table 4
Gas turbine input parameters for the ASPEN­
HYSYS® simulation. PR represents the total
pressure ratio in the compression zone, Qe
compression intercooler heat transfer, ηcompr
compression adiabatic efficiency, Qp1 combus­
tion chamber heat losses, ηexpan expansion
adiabatic efficiency, and Qp2 expansion heat
losses.
Parameter Value

PR 12.96
Qe, kW 240
ηcompr, % 70
Qp1, kW 320
ηexpan, % 89.30
Qp2, kW 1117

approximately 653 kPa, which results in wet steam.


During transport in the “process heat and steam expansion” section, heat losses depend on the heat exchanger temperature dif­
ference. Since the vapor turbine inlet temperature and boiler outlet temperature are fixed values, the temperature difference through
E− 104 and E− 105, ΔTHE, is calculated with Equation (4).
T28 − T31
ΔTHE = (4)
2

3. Results and discussion

3.1. Validation of the simulation model

Actual plant data allowed validating the simulation results presented here. The data gathered in 2014 contains daily average values
of operational parameters for 24 days of the gas turbine operation. The ambient air temperature, Tamb , and the expansion inlet
temperature, T11 , were used for validation. Table 5 shows the comparison between the developed simulation and gas turbine

Table 5
Daily average values of the actual running (AR) gas turbine and results of the ASPENHYSYS® simulation model (SM) for the cogeneration plant
located in Barranquilla, Colombia. Tamb represents the ambient temperature, T 11 the expansion inlet temperature, ST11 the standard deviation of the
expansion inlet temperature daily average, Ẇ GT gas turbine power output, ηGT the gas turbine termal efficiency, and V̇ NG GT the turbine fuel con­
sumption rate. It shows the differences (D) between actual data and simulation results as percentage values.
Tamb , ◦ C T11 , ◦ C ST11 , ◦ C ẆGT , MW ηGT , % V̇ NG GT , Nm3 s− 1

AR SM D, % AR SM D AR SM D, %

28.0 689 54.5 4.2 4.6 9.4 22.3 25.6 14.8 0.48 0.49 3.3
29.1 727 10.4 4.7 5.0 6.1 25.9 27.8 7.4 0.46 0.52 13.7
28.4 726 6.8 4.6 5.0 7.6 25.2 27.8 10.3 0.47 0.52 12.0
27.4 716 8.8 4.4 4.9 11.0 24.7 27.3 10.4 0.45 0.52 13.9
28.8 721 6.7 4.5 4.9 9.3 25.0 27.5 10.1 0.46 0.52 13.2
28.8 728 5.2 4.7 5.0 6.7 25.5 27.9 9.6 0.47 0.52 12.2
28.7 733 5.9 4.8 5.0 5.5 25.7 28.3 9.9 0.47 0.53 11.5
28.4 737 6.3 4.8 5.1 6.2 25.7 28.5 10.8 0.48 0.53 12.1
28.8 718 8.4 4.2 4.9 16.2 24.2 27.3 12.9 0.44 0.52 16.8
28.9 711 9.8 4.2 4.8 15.1 24.1 26.8 11.4 0.44 0.51 16.0
29.7 716 9.5 4.3 4.8 11.6 24.5 27.1 10.4 0.45 0.51 14.2
28.8 734 10.6 4.8 5.0 4.7 25.8 28.3 9.6 0.48 0.53 11.1
28.6 735 4.6 4.8 5.1 5.8 25.7 28.4 10.3 0.47 0.53 11.8
28.9 734 8.0 4.6 5.0 9.6 25.3 28.3 12.1 0.47 0.53 13.8
28.4 738 5.8 4.8 5.1 6.1 25.7 28.6 11.3 0.48 0.53 11.6
28.7 735 9.1 4.8 5.1 5.0 25.7 28.4 10.3 0.48 0.53 11.0
29.2 742 6.4 4.9 5.1 4.8 25.8 28.8 11.3 0.48 0.54 10.8
28.2 735 6.9 4.8 5.1 6.7 25.5 28.4 11.4 0.48 0.53 11.3
29.1 740 9.3 4.9 5.1 4.5 25.8 28.7 11.1 0.48 0.53 10.3
28.8 741 5.9 4.9 5.1 5.2 25.8 28.7 11.4 0.48 0.54 10.9
28.8 738 11.7 4.8 5.1 7.3 25.7 28.7 11.8 0.47 0.53 12.7
28.8 725 9.1 4.4 4.9 11.4 24.8 27.7 11.6 0.46 0.52 14.5
27.8 716 8.7 4.3 4.9 13.5 24.4 27.3 11.7 0.45 0.52 15.1
30.6 729 14.5 4.6 4.9 7.7 25.2 27.8 10.3 0.46 0.52 12.5

7
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

experimental data from the gas turbine. Table 5 shows the difference between the power output, ẆGT , first law efficiency, ηGT , and fuel
mass flow rate, V̇ NG GT , of the actual gas turbine and results from the simulation. Daily standard deviations of expansion inlet tem­
perature, ST11 , are lower than 10%, which indicates some level of data variability throughout the day. The maximum difference be­
tween simulations and the actual gas turbine data amounts to 17%.
The gathered data for the validation of the HRSG simulation corresponds to 20 days of operation. The daily average of the sup­
plementary fuel flow rate, V̇ Sup NG , is the inlet value of this system. Table 6 shows the comparison between the simulation and the
experimental data from the HRSG. Table 6 includes the difference between the flow rate of steam production in the HRSG, ṁ28 . The
comparison between the HRSG simulation and the actual system results in differences up to 12%.

3.2. Analysis of simulation results

This study combines performance curves and experimental data. The gas turbine manufacturer reports that the power output at full
load operation and ISO condition amounts to 6.35 MW and the thermal efficiency to 34%. The gas turbine operational curves show that
during ISO conditions, exhaust gas temperature falls in the range of 480–490 ◦ C, the mass flow rate of exhaust gas in the range of
25.2–26.5 kg s− 1, and turbine inlet temperature in the range of 720–740 ◦ C [34].
The ambient temperature in Barranquilla ranges from 27 to 37 ◦ C. As temperature increases above nominal conditions (15 ◦ C), air
density and specific compression work negatively affect the turbine power output. This section deals with the air intake temperature
effect in turbine output power, fuel consumption, flue gas temperature, and turbine thermal efficiency.
The HRSG system and the vapor turbine improve plant performance, thus counterbalancing the ambient temperature adverse
effects. We now evaluate how steam generation varies with ambient temperature, turbine inlet temperature, and excess air factor for
fired mode operation. The discussion then extends to the steam mass flow through vapor turbine and excess air factor in global output
power, thermal efficiency, and utilization factor for cogeneration plant when the system works for fired and unfired operation.
Equation (5), derived from Ref. [14], describes the excess air factor (EAF) estimation at different operation conditions. In this equation
ṁ3 corresponds to the air mass flow rate of the turbine, ṁ7 the fuel mass flow rate, and AFRstch the stoichiometric value of the air-fuel
ratio. The AFRstch equals 17.1 for the complete combustion reactions involved in this process.
ṁ3
EAF = − 1 (5)
AFRstch ṁ7
In these cases, the ambient temperature remains at 32 ◦ C with maximum supplementary fuel consumption. The Supplementary
Material shows how ambient temperature affects the mass flow rate of flue gases and the excess air factor. Additionally, Supplementary
Material includes the steam generation variation in the HRSG with ambient temperature, turbine inlet temperature, and excess air
factor during unfired operation. Supplementary Material also explores the effect of steam mass flow rate on the vapor turbine.
Fig. 3a shows the turbine output power dependence on ambient air temperature and turbine inlet temperature. Increments in
ambient air temperature negatively affect the turbine output power as expected. Higher temperatures result in lower air density and a

Table 6
Daily average values of the actual running (AR) HRSG and ASPENHYSYS® simulation model (SM) results for
the cogeneration plant in Barranquilla, Colombia. V̇ Sup NG represents the supplementary fuel flow rate, and ṁ28
steam production in the HRSG mass flow rate. D stands for the differences between actual data and simulation
results as percentage values.
1
V̇Sup NG , Nm3 h− 1 ṁ28 , kg s−

AR SM D, %

436 5.0 5.1 2.4


626 5.8 6.4 8.9
593 5.8 6.2 6.6
570 5.5 6.0 8.2
559 5.6 6.0 7.0
548 5.6 6.0 7.6
559 5.6 6.1 8.4
593 5.6 6.3 11.1
671 6.1 6.6 9.1
660 6.1 6.6 7.2
705 6.3 6.8 9.1
682 6.3 6.7 7.0
682 6.1 6.8 10.5
682 6.1 6.7 9.4
682 6.2 6.7 9.4
716 6.3 6.9 9.7
638 5.9 6.5 9.7
548 5.4 6.0 10.5
570 5.5 6.0 9.2
582 5.6 6.1 9.2

8
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

consequent reduction of the air mass flow rate, which leads to a reduction of fuel mass flow rate entering the combustion chamber.
Likewise, power output is directly related to the increase in the turbine inlet gas temperature. Increasing fuel mass flow rate leads to
higher gas temperatures at the inlet of the turbine, thus alleviating the negative effect of higher air temperatures. However, the gas
turbine control system limits gases temperature at the inlet to 750 ◦ C, ultimately determining the maximum turbine output power.
During the hottest days of the year, when the ambient temperature reaches 36.7 ◦ C, the maximum turbine output power reaches 4.95
MW, i.e., decreases by around 22% compared with ISO conditions. According to the literature, the power output drops expected at
36.7 ◦ C of ambient temperature should range 10–17% concerning ISO conditions [3,4,7]. Note that the local air humidity for the
current study is higher than in previous literature.
Fig. 3b shows how ambient air and gas temperature changes at the turbine inlet affect fuel consumption. Ambient air temperature is
inversely related to fuel consumption while keeping the turbine inlet gas temperature constant. The increase of ambient air tem­
perature causes a decrease in the air mass flow rate because the air density increases. Because of this, fuel flow drops to meet turbine
inlet temperature without surpassing the operational upper limit. The reduction of fuel consumption relates to Fig. 3a, in which the
increase in ambient air temperature causes power output to decrease. According to experimental data, the fuel volumetric flow rate
ranges from 0.506 to 0.514 Nm3s− 1, which results in power output between 4.7 and 4.8 MW.
Fig. 4a and Fig. S5 (see Supplementary Material) illustrate the effect of ambient temperature and inlet temperature in the turbine
outlet temperature and gases mass flow rate. Fig. 4a shows that the turbine outlet gas temperature is directly related to ambient air
temperature and turbine inlet temperature changes. However, turbine inlet temperature changes at constant ambient air temperature
have a more decisive influence on outlet gas temperature than the opposite. On the other hand, Fig. S5 in the Supplementary Material
shows that ambient air temperature negatively affects the gases mass flow rate at constant turbine inlet temperature. Both outlet gas
temperature and gas mass flow rate are relevant variables that affect steam production in the HRSG unit. Nevertheless, the broad range
of exhaust gas temperature, i.e., 490–530 ◦ C, makes it the most influential variable compared to the exhaust gases mass flow rate range,
i.e., 22.05–22.7 kg s− 1, as shown below. Results suggest a variation of excess air factor (EAF) during gas turbine operation since fuel
consumption and air mass flow rate featured different behaviors regarding ambient temperature changes.
Fig. 4b illustrates the gas turbine efficiency variation caused by changes in inlet gas temperature and ambient air temperature.
During the warmest days of the year, gas turbine efficiency drops to 26%. On the contrary, when the ambient air temperature is at the
lower limit, the turbine efficiency reaches 27.5%. Increasing ambient air temperature negatively affects turbine performance because
lower gas mass flow decreases power output. Gas turbine efficiency decreases 0.06% for every ◦ C rise of ambient temperature above
ISO conditions. The drop in thermal efficiency found in this work agrees with other studies’ results, which report gas turbine thermal
efficiency drops from 0.03 to 0.3% for every ◦ C rise in ambient temperature [3–5,7]. Also, variations in the turbine inlet gas tem­
perature with constant ambient temperature are directly proportional to gas turbine efficiency. In this case, air mass flow and
compression work remain constant, but the available energy in combustion gases increases. Considering the turbine operational limits
and historical data, the maximum efficiency that the system can reach during the warmest days is 26.7%.
On the other hand, Fig. 5a shows the HRSG steam production variation with turbine inlet temperature and supplementary fuel
consumption in the furnace. Increasing the turbine inlet temperature improves the HSRG performance by reducing up to 1.2 Nm3 hr− 1
of supplementary gas per degree ◦ C. On this basis, increasing turbine inlet temperature from 716 to 740 ◦ C at constant steam mass flow
production represents a total decrease of 40 Nm3 hr− 1 in supplementary fuel consumption. Steam production increases from 5.1 to 6.5
kg s− 1 as supplementary fuel consumption increases from 365 to 640 Nm3 hr− 1 at an expansion inlet temperature of 740 ◦ C.
Fig. 5b shows the HRSG steam production as a function of supplementary fuel consumption for different EAF values and an ambient

Fig. 3. (a) Gas turbine power output and (b) fuel consumption as a function of ambient air temperature and turbine inlet temperature. Ambient air
temperature ranges from 27 ◦ C to 37 ◦ C and turbine inlet temperature from 720 ◦ C to 750 ◦ C.

9
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

Fig. 4. (a) Gas turbine outlet temperature and (b) efficiency change with ambient air temperature and turbine inlet temperature. Ambient air
temperature varies from 27 ◦ C to 37 ◦ C and turbine inlet temperature from 720 ◦ C to 750 ◦ C.

temperature of 32 ◦ C. As expected, increasing EAF reduces steam production in the HRSG at constant supplementary fuel consumption.
The mass flow rate of steam decreases by around 0.03 kg s− 1 for every 0.02 rise in EAF. On the other hand, steam production directly
relates to supplementary fuel consumption at constant EAF. The extra fuel required to obtain a constant HRSG steam production is
lower when EAF decreases since the turbine exhaust gas temperature is inversely related to EAF. Hence, supplementary fuel con­
sumption decreases 7.0 Nm3 hr− 1 every 0.02 reduction in EAF to maintain a constant steam production.
Fig. 6 shows power output, thermal efficiency, and the cogeneration plant utilization factor. These parameters illustrate the
advantage of adding an HRGS and a vapor turbine to a natural gas-based power plant. Fig. 6a shows the variation of global power
output as a function of steam mass flow through the vapor turbine for different values of EAF in the gas turbine. The ambient tem­
perature remains constant at 32 ◦ C in this analysis. Global output power is directly related to steam mass flow through the vapor
turbine for a constant value of EAF. Implementing the cogeneration system with a vapor turbine improves power generation without
supplementary fire. For an EAF constant value, global power output consists of two parts of energy. The vapor turbine generates up to
5% of additional power than the gas turbine power output without supplementary fire. When the cogeneration system operates with
supplementary fire, the vapor turbine generates up to 17% additional power than the gas-turbine-alone operation. Most of the time, the
cogeneration plant power output ranges from 4.9 to 5.1 MW without supplementary fire. On the other hand, with supplementary fire,

Fig. 5. Steam production in the HRSG depends on supplementary fuel consumption and (a) turbine inlet temperature and (b) excess air factor, EAF.
Ambient air temperature remains at 32 ◦ C, exhaust gas mass flow 22.37 kg s− 1, and fired operation. Fuel consumption varies from 320 to 700 Nm3
hr− 1. Turbine inlet temperature, T11, ranges from 716 to 740 ◦ C and EAF from 2.40 to 2.48.

10
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

the cogeneration plant power output varies from 5.0 to 5.6 MW.
Fig. 6b shows the global thermal efficiency variation with steam mass flow through vapor turbine for different EAF values. The
global efficiency rises to 28% at the maximum steam mass flow rate through the vapor turbine for unfired operation. The addition of an
HRSG and a vapor turbine improves global efficiency. On the other hand, efficiency drops down to 20.2% with the minimum steam
mass flow rate through the vapor turbine with supplementary fire in the HRSG. Heat losses during steam generation, transport, and
expansion decrease the exploitable energy from supplementary fire.
The energy utilization factor is an essential indicator for comparing the cogeneration plant with the gas-based conventional power
plant. Equation (6) indicates how to estimate the energy utilization factor (EUF) for the cogeneration plant. ẆGT represents the gas
turbine power output, ẆVT the vapor turbine shaft power, Q̇steam the heat flow to produce the superheated steam for plant re­
quirements, ṁ7 and ṁ14 are the natural gas mass flow rate in the gas turbine and supplementary combustion, respectively. LHV
represents the fuel lower heating value. Note that the utilization factor is equal to the thermal efficiency when the gas turbine operates
alone.

Ẇ GT + Ẇ VT + Q̇steam
EUF = (6)
(ṁ7 + ṁ14 )LHV
Fig. 6c shows how the overall energy utilization factor varies with the steam mass flow rate through the vapor turbine for different
EAF values. The cogeneration system with HRSG and the vapor turbine improves the utilization factor. The energetic exploitation of
exhaust gases raises the utilization factor up to 37% for an EAF of 2.40. However, most of the time, the global utilization factor ranges
from 28 to 36%. This improvement represents an increase of about 5–35% compared to the gas turbine maximum value for the unfired
operation. Our results are consistent with the study developed by Ref. [17], with improvements in the range of 18–34%.
On the other hand, utilization factor ranges 25–49% with supplementary fire. When most steam flows through the vapor turbine,
shaft transmission losses rise far higher than transport heat losses. Consequently, increasing the steam mass flow rate through the vapor
turbine above 6 kg s− 1 causes the utilization factor to drop below the gas turbine reference value of 26.4%. The utilization factor is
inversely related to steam mass flow through the vapor turbine and EAF. Nevertheless, the effect of changing EAF from 2.40 to 2.48 in
the utilization factor is less than 5% for both unfired and fired modes.
Figure S11 in the Supplementary Material compares the simulated energy utilization factor of the cogeneration plant and the
efficiency values from other combined cycle plants in the literature reviewed [6,20,21]. The results obtained in the simulation are
consistent with other studies. The EUF values from the fired mode operation are closer concerning the reported efficiencies from Refs.
[6,20], with differences up to 10%. Also, both unfired and fired operations present better performance than the efficiency values from
Ref. [21], although their power generation system works below the ISO temperature. The implementation of the supplementary fire
system is a distinctive feature of this work. The cogeneration plant operating under fired mode increases up to 15% in the EUF relative
to the unfired mode. This advantage contributes to stabilizing plant performance with variable loads. Likewise, the supplementary fire
system facilitates the supply of additional energy to meet demand peaks during the highest temperature conditions.

3.3. Outlook

The previous sections showed the adverse effects on the gas turbine performance caused by the hot and humid climate with
temperatures well above ISO conditions. The addition of a heat and power cogeneration system enhanced output power and thermal
efficiency. The waste heat recovery system was a convenient way to further exploit the remaining energy in the flue gases from the gas
turbine.
Installing an HRSG contributed to saving energy and decreased the overall plant fuel consumption. A fraction of the thermal energy
in the turbine exhaust gases generated steam, which resulted in additional energy to perform other tasks in the plant, such as paper
drying or air heating. The addition of a steam turbine after the HRSG resulted in a considerable increase in shaft power. These actions
demonstrate a significant reduction of the impact of ambient temperature on a gas turbine-based power generation facility in a hot
climate. An open possibility is connecting the steam turbine to a second electric generator for stabilization during the year’s warmest
days when the gas turbine does not meet the power demand.
The supplementary fire system enhanced the performance of the cogeneration power plant when the unfired operation does not
meet the total energy demand. The main advantage of the fire system is that no additional compressed air is necessary because the
extra fuel reacts with the remaining oxygen in the exhaust gases. Results in this paper revealed a considerable increment of steam that
could increase either the process heat or the steam turbine power output. The potential increment of global power output with lower
fuel consumption converts the supplementary fire into a practical addition to the cogeneration power plant.
The constant volume operation in the air compressor limits the gas turbine performance as ambient temperature increases. Still, the
results indicate that the power output and efficiency of the gas turbine improved as long as the ambient temperature remained at their
lower values. Because of this, the incorporation of an inlet air cooling system could mitigate the negative influence of high temper­
atures in tropical locations. Though this work does not include an air-cooling system, further research should assess the effect of inlet
air cooling on the cogeneration plant performance.

4. Conclusions

This case study presented a validated cogeneration plant model developed in ASPENHYSYS® software. The model gave insight into

11
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

Fig. 6. Cogeneration plant (a) power output, (b) global efficiency, and (c) utilization factor as functions of steam mass flow through the vapor
turbine for different values of excess air factor (EAF). Ambient air temperature sets at 32 ◦ C. The supplementary fuel consumption is constant at 700
Nm3 hr− 1. Left side lines represent the performance parameter value under unfired mode (UFM) and right side under fired mode (FM).

the effects of high ambient temperature on a cogeneration plant located in Barranquilla (Colombia), a location with a hot and humid
tropical climate characterized by high temperatures and humidity throughout the year.
High ambient temperature decreases air density and consequently the air mass flow rate of the gas turbine. The consequence was a
drop in both power output and thermal efficiency for gas-turbine-alone operation. The gas turbine power output featured a maximum
reduction of 22%, and the thermal efficiency decreased by 26% during the hottest days of the year compared to ISO conditions. Heat,
recovered from the gas turbine exhaust gases, produced superheated steam and additional power. The cogeneration increased global
power output and the energy utilization factor of the plant. The global power output increased up to 17% compared to gas-turbine-
alone operation, and the energy utilization factor raised between 25 and 49%.
Cogeneration mitigates the negative effect of high ambient temperatures, such as those in the Colombian Caribbean. Cooling inlet
air is an attractive alternative for improving power plant performance in a tropical climate. Further studies should assess the effects of
high humidity on the power plant performance and the implementation of inlet air cooling.

CRediT authorship contribution statement

Daniel Pinilla Fernandez: Writing- Original draft preparation, software. Blanca Foliaco: Methodology, Writing- Original draft
preparation. Ricardo Vasquez Padilla: Conceptualization; Writing - review & editing. Antonio Bula: Project administration; Writing
- review & editing. Arturo Gonzalez-Quiroga: Supervision; Writing - review & editing.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgments

The authors acknowledge the support from UNIVERSIDAD DEL NORTE. BF acknowledges the support provided by Ministerio de
Ciencia Tecnología e Innovación MINCIENCIAS through the “Convocatoria Nacional para Estudios de Doctorado en Colombia, con­
vocatoria 727”.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.csite.2021.101206.

References

[1] International Energy Agency - Iea, Energy efficiency 2020, Energy Effic (2020) 2020, https://doi.org/10.1787/dfd85134-en.
[2] M.N. Khan, I. Tlili, New advancement of high performance for a combined cycle power plant: thermodynamic analysis, Case Stud. Therm. Eng. 12 (2018)
166–175, https://doi.org/10.1016/j.csite.2018.04.001. April.
[3] O. Agwu, C. Eleghasim, Mechanical drive gas turbine selection for service in two natural gas pipelines in Nigeria, Case Stud. Therm. Eng. 10 (November 2016)
19–27, https://doi.org/10.1016/j.csite.2017.02.003, 2017.

12
D.A. Pinilla Fernandez et al. Case Studies in Thermal Engineering 26 (2021) 101206

[4] A. De Sa, S. Al Zubaidy, Gas turbine performance at varying ambient temperature, Appl. Therm. Eng. 31 (14–15) (2011) 2735–2739, https://doi.org/10.1016/j.
applthermaleng.2011.04.045.
[5] M.M. Rahman, T.K. Ibrahim, A.N. Abdalla, Thermodynamic performance analysis of gas-turbine power-plant, Int. J. Phys. Sci. 6 (14) (2011) 3539–3550,
https://doi.org/10.5897/IJPS11.272.
[6] T.K. Ibrahim, M.M. Rahman, Thermal impact of operating conditions on the performance of a combined cycle gas turbine, J. Appl. Res. Technol. 10 (2012)
567–577, scielomx.
[7] S.S. Baakeem, J. Orfi, S. Alaqel, H. Al-Ansary, Impact of ambient conditions of arab Gulf countries on the performance of gas turbines using energy and exergy
analysis, Entropy 19 (1) (2017), https://doi.org/10.3390/e19010032.
[8] CIOH Oceanografía Operacional, Climatología del Caribe [Accessed: 10-Jan-2021], https://www.cioh.org.co/meteorologia/Climatologia/
ResumenBarranquilla3.php.
[9] G.V. Pradeep Varma, T. Srinivas, Design and analysis of a cogeneration plant using heat recovery of a cement factory, Case Stud. Therm. Eng. 5 (2015) 24–31,
https://doi.org/10.1016/j.csite.2014.12.002.
[10] Z. Liu, I.A. Karimi, T. He, A novel inlet air cooling system based on liquefied natural gas cold energy utilization for improving power plant performance, Energy
Convers. Manag. 187 (November 2018) 41–52, https://doi.org/10.1016/j.enconman.2019.03.015, 2019.
[11] G. Şen, et al., The effect of ambient temperature on electric power generation in natural gas combined cycle power plant—a case study, Energy Rep. 4 (2018)
682–690, https://doi.org/10.1016/j.egyr.2018.10.009.
[12] A.M. Al-Ibrahim, A. Varnham, A review of inlet air-cooling technologies for enhancing the performance of combustion turbines in Saudi Arabia, Appl. Therm.
Eng. 30 (14–15) (2010) 1879–1888, https://doi.org/10.1016/j.applthermaleng.2010.04.025.
[13] M.A. Ehyaei, S. Hakimzadeh, N. Enadi, P. Ahmadi, Exergy, economic and environment (3E) analysis of absorption chiller inlet air cooler used in gas turbine
power plants, Int. J. Energy Res. 36 (4) (Mar. 2012) 486–498, https://doi.org/10.1002/er.1814.
[14] V. Ganapathy, Steam Generators and Waste Heat Boilers: for Process and Plant Engineers, CRC Press, 2014.
[15] E. Bilgen, Exergetic and engineering analyses of gas turbine based cogeneration systems, Energy 25 (12) (2000) 1215–1229, https://doi.org/10.1016/S0360-
5442(00)00041-4.
[16] L. Zheng, E. Furimsky, ASPEN simulation of cogeneration plants, Energy Convers. Manag. 44 (11) (2003) 1845–1851, https://doi.org/10.1016/S0196-8904(02)
00190-5.
[17] M.N. Khan, I. Tlili, W.A. Khan, Thermodynamic optimization of new combined gas/steam power cycles with HRSG and heat exchanger, Arabian J. Sci. Eng. 42
(11) (2017) 4547–4558, https://doi.org/10.1007/s13369-017-2549-4.
[18] U. Ali, C.F. Palma, K.J. Hughes, D.B. Ingham, L. Ma, M. Pourkashanian, Thermodynamic analysis and process system comparison of the exhaust gas recirculated,
steam injected and humidified micro gas turbine, Proc. ASME Turbo Expo 3 (2015), https://doi.org/10.1115/GT2015-42688.
[19] Z. Liu, I.A. Karimi, New operating strategy for a combined cycle gas turbine power plant, Energy Convers. Manag. 171 (2018) 1675–1684, https://doi.org/
10.1016/j.enconman.2018.06.110. July.
[20] Z. Liu, I.A. Karimi, Simulating combined cycle gas turbine power plants in Aspen HYSYS, Energy Convers. Manag. 171 (2018) 1213–1225, https://doi.org/
10.1016/j.enconman.2018.06.049. June.
[21] A. Wiguno, R. Tetrisyanda, G. Wibawa, The effect of gas composition, air intake cooling, and steam injection on combined cycle power plant performance, AIP
Conf. Proc. 2248 (2020), https://doi.org/10.1063/5.0013325. July.
[22] M. Shayan, V. Pirouzfar, H. Sakhaeinia, Technological and economical analysis of flare recovery methods, and comparison of different steam and power
generation systems, J. Therm. Anal. Calorim. 139 (4) (2020) 2399–2411, https://doi.org/10.1007/s10973-019-08429-9.
[23] H.J. Kadhim, T.J. Kadhim, M.H. Alhwayzee, A comparative study of performance of Al-khairat gas turbine power plant for different types of fuel, IOP Conf. Ser.
Mater. Sci. Eng. 671 (1) (2020), https://doi.org/10.1088/1757-899X/671/1/012015, 11.
[24] D.E. Babatunde, A.N. Anozie, J.A. Omoleye, O.J. Odejobi, Performance evaluation of a major thermal power plant in Nigeria, IOP Conf. Ser. Earth Environ. Sci.
655 (2021), https://doi.org/10.1088/1755-1315/655/1/012059.
[25] B. Jabbari, N. Tahouni, A. Ataei, M.H. Panjeshahi, Design and optimization of CCHP system incorporated into kraft process, using Pinch Analysis with pressure
drop consideration, Appl. Therm. Eng. 61 (1) (2013) 88–97, https://doi.org/10.1016/j.applthermaleng.2013.01.050.
[26] U.P.M.E. Unidad de Planeación Minero energética -, Estudio técnico para el plan de abastecimiento de gas natural, 2020.
[27] U.P.M.E. Unidad de Planeación Minero Energética -, Informe mensual de variables de generación y del mercado eléctrico colombiano, ” Bogotá, 2018.
[28] S.A.E.S.P. Promigas, BEO Información Operacional/Composición del Gas Vigente [Accessed: 01-Jun-2020], http://www.promigas.com/Es/BEO/Paginas/
ComposicionGasVigente.aspx.
[29] M.A. Adnan, M.M. Azis, M.R. Quddus, M.M. Hossain, Integrated liquid fuel based chemical looping combustion – parametric study for efficient power
generation and CO2 capture, Appl. Energy 228 (2018) 2398–2406, https://doi.org/10.1016/j.apenergy.2018.07.072.
[30] G. Di Marcoberardino, F. Gallucci, G. Manzolini, M. van Sint Annaland, Definition of validated membrane reactor model for 5 kW power output CHP system for
different natural gas compositions, Int. J. Hydrogen Energy 41 (42) (2016) 19141–19153, https://doi.org/10.1016/j.ijhydene.2016.07.102.
[31] AspenTech, Aspen HYSYS Help, V10, Bedford, 2017.
[32] Solar Turbines Incorporated, Turbomachinery Package Specification: TaurusTM 70 Compressor Set and Mechanical Drive, Solar Turbines Incorporated, San Diego,
2009.
[33] The American Society of Mechanical Engineers, ASME Boiler and Pressure Vessel Code. Section I: Rules for Construction of Power Boilers, 2010. New York, NY.
[34] Solar Turbines Incorporated, TaurusTM 70 - Instruction Manual, 1995.

13

You might also like