Nothing Special   »   [go: up one dir, main page]

A Review On Efficiency Improvement Methods in Organic Rankine Cycle System: An Exergy Approach

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

International Journal of Advances in Applied Sciences (IJAAS)

Vol. 11, No. 1, March 2022, pp. 1~10


ISSN: 2252-8814, DOI: 10.11591/ijaas.v11.i1.pp1-10  1

A review on efficiency improvement methods in organic


Rankine cycle system: an exergy approach

Gollangi Raju1, Nagamalleswara Rao Kanidarapu2


1
School of Mechanical Engineering, Vellore Institute of Technology, Vellore, India
2
Centre for Disaster Mitigation and Management, School of Chemical engineering, Vellore Institute of Technology, Vellore, India

Article Info ABSTRACT


Article history: Exergy, one of the handed-down energy conservation techniques, which can
obtain from thermodynamic laws (first and second), will disclose the work
Received Oct 11, 2020 presented within the system, the amount of irreversibility as well as what are
Revised Jul 14, 2021 the possible ways to reduce inefficiencies in the system. This discourse
Accepted Dec 9, 2021 mainly highlighted various techniques and possible methods for efficiency
improvement in the organic Rankine cycle (ORC). That means mainly
concentrated on following key parameters like the selection of working
Keywords: fluid, suitable expander, the different heat sources of an evaporator, and
modifications in heat exchanger based on the application of ORC system
Exergy analysis through an exergy approach for better performance, decrease energy losses,
Exergy destruction and destruction rate. This review can help to pontificate for better-
Heat source temperature summarized results that were done before and suggest some ideas for how to
Organic Rankine cycle select an optimized parameter for better efficiency and to decrease the
Working fluid destruction rate in the ORC system.
This is an open access article under the CC BY-SA license.

Corresponding Author:
Nagamalleswara Rao Kanidarapu
Centre for Disaster Mitigation and Management, Vellore Institute of Technology
Vellore, Tamilnadu, India
Email: aspenmodels@gmail.com

NOMENCLATURE
Nomenclature Subscript
ORC Organic Rankine cycle d Destruction
FLT The first law of thermodynamics Cr critical
SLT The second law of thermodynamics hi Hot stream inlet
I Irreversibility ho Hot stream outlet
s Entropy ci Cold stream inlet
T Temperature in K co Cold steam outlet
m Mass in kg/mol out out
Ex Physical Exergy 0 Reference state
e Specific exergy Eva Evaporator
h Enthalpy ex Expander
W Work done Cond Condenser

1. INTRODUCTION
From past decades people are facing a lot of issues to successful utilization of energy and also to
recover energy from waste heat. Primarily vapor Rankine cycle with water as a working fluid is a way to
convert a large amount of thermal energy into power [1]–[7]. Water has phenomenal properties
(thermal/chemical), is easy to pump from one place to another also has some disadvantages like erosion of
blades and it is not suitable for low-temperature applications [3], [6], [8].

Journal homepage: http://ijaas.iaescore.com


2  ISSN: 2252-8814

To fill this gap, the need for a conventional or organic working fluid that has a better property at
low-temperature ranges to recover lower heat into work. This ORC was established to produce work output
from low-grade waste heat. ORC system can use for power generation in a wide range of applications likely
power plants, geothermal plants, solar applications, and waste heat in industries as shown in Figure 1 [9],
[10]. Heat driven from geothermal ORC has performed with low-pressure vapor generator will give 38.11%
exergy efficiency and 29.98% for high-pressure generator [11]–[14]. Solar ORC coupled with internal feed
liquid heater and internal heat exchanger will give better exergy results and also ORC combined with
trilateral flash cycle increases up to 15.94% of exergy efficiency [15]–[18]. Introducing an internal heat
exchanger, effective usage of turbine bleeding/regeneration in basic ORC helps to get higher efficiency of
nearly 38.82% [19]–[21]. The researchers provided a simplified way to select an optimum operating
conditions ORC and produced 264.14 kW power at 1300 °C temperature heat source [22]–[24]. New
modified ORC suggested that energy, exergy, environmental aspects are considered for better performance
[25]–[30] while recovering heat from gas turbine exhaust, the exergy destruction rate of ORC depends on
heat source temperature [31] that means if it is increased, the rate of exergy destruction also increases.
Energy recovers from medium ORC employed with liquefied natural gas (LNG); transcritical CO2 cycle
raised the exergetic efficiency from 12.3 to 13.08% [5]. Work recovery from LNG cold energy is possible
with ORC by using efficient working fluid R22 [32]. The main aim of this review paper is to notify various
exergy-based efficiency improvement methods in the ORC system. The paper contains three chapters: first to
select the ORC system and its working also equation employed with exergy presented. The second chapter
talks about the role of various equipment in ORC, also how to select an optimized one based on the literature.
The third chapter with some conclusions and suggestions in calculating exergy destruction and efficiency
improvement methods.

Figure 1. ORC system sharing with the different heat source

2. METHODOLOGY STRUCTURE OF ORGANIC RANKINE CYCLE


The most efficient way to produce power from waste heat (lower grade) is the ORC, it has the
structure of a condenser, pump, evaporator, expansion device as shown in Figure 2. An effective working
fluid will play a major role to generate power by observing heat from the heat source and converting it into
work with the help of an expander.

Figure 2. Schematic line diagram of an ORC system

Int J Adv Appl Sci, Vol. 11, No. 1, March 2022: 1-10
Int J Adv Appl Sci ISSN: 2252-8814  3

Processes are done on this cycle: i) Isentropic pumping process to raise the pressure of working
fluid; ii) Isobaric evaporation to raise the temperature of the fluid (phase change) at constant pressure with
the help heat source, fluid will get more potentiality; iii) Isentropic expansion to produce work output with
the help of turbine or other expanders, at this stage fluid will lose its internal energy; and iv) Isobaric
condensation to get back its original phase for pumping purpose.

3. ENERGY AND EXERGY ANALYSIS


From FLT in the steady-state condition, the mass and energy pass through a system is always
constant, also potential and kinetic energy are neglected in the control volume. Mathematically FLT:
i) equation for mass balance as (1); and ii) energy balance in the control system as (2) and (3).
∑ 𝑚𝑖𝑛 = ∑ 𝑚𝑜𝑢𝑡 (1)

𝑒𝑖𝑛 = 𝑒𝑜𝑢𝑡 (2)

𝑄 + 𝑊 = ∑ 𝑚𝑜𝑢𝑡 ℎ𝑜𝑢𝑡 − ∑ 𝑚𝑖𝑛 ℎ𝑖𝑛 (3)

SLT always says there is an entropy generation called irreversibility in any process, which means total
energy at the inlet doesn’t match with the outlet. Based on this there is new technique exergy (availability of
work) was introduced to reduce irreversibility within the system. The maximum possible work generated
from a system to its reference (temperature 298 K, pressure 1 bar) or atmosphere condition. In ORC there are
no chemical changes along with all processes. So chemical exergy is considered as zero. Figure 3 shows the
physical exergy module.

Figure 3. Definition of physical exergy [33]

Mathematical expression: i) Exergy can express as (4); ii) Specific exergy as (5) and (6); and
iii) Steady-state condition as (7).

𝐸𝑥̇ = 𝑚𝑒𝑥 (4)


𝑒𝑥 = ℎ − ℎ0 − 𝑇0 (𝑠 − 𝑠0 ) (5)

∑ 𝐸𝑥̇𝑖𝑛 − ∑ 𝐸𝑜𝑢𝑡 − 𝐸𝑥̇ 𝑑 = ∆𝐸𝑥̇ (6)

𝑄 + 𝑊 = ∑ 𝐸𝑥̇𝑜𝑢𝑡 − ∑ 𝐸𝑥̇𝑖𝑛 + 𝐼 (7)

Where, Q is total heat input, W is work output from the system, and I is irreversibility in the system rate.
Heat and exergy equations for an evaporator and condenser:
𝑄 = 𝑚ℎ(ℎ5 − ℎ6 ) = 𝑚𝑓 (ℎ3 − ℎ2 ) (8)
𝐸𝑣𝑎

𝐼𝐸𝑣𝑎 = (𝐸5 − 𝐸6) − (𝐸3 − 𝐸2) (9)


𝑄𝑐𝑜𝑛𝑑 = 𝑚𝑐 (ℎ8 − ℎ7 ) = 𝑚𝑓 (ℎ4 − ℎ1 ) (10)

𝐼𝑐𝑜𝑛𝑑 = (𝐸7 − 𝐸8) − (𝐸1 − 𝐸4) (11)


A review on efficiency improvement methods in organic Rankine … (Nagamalleswara Rao Kanidarapu )
4  ISSN: 2252-8814

where, I represent irreversibility in the respective process. The exergetic efficiency of condenser and
evaporator can be written as (12) and (13); for expander as (14), (15), and (16); for the pump as (17), (18),
and (19).
ɳ𝐸𝑣𝑎 = 1 − (𝐼𝐸𝑣𝑎 /(𝐸5 − 𝐸6 )) (12)

ɳ 𝑐𝑜𝑛𝑑 = 1 − (𝐼𝑐𝑜𝑛𝑑 /(𝐸4 − 𝐸1 )) (13)

𝑊𝑒𝑥 = 𝑚 𝑓 (ℎ3 − ℎ4 ) (14)

𝐼𝑒𝑥 = 𝐸3 − (𝐸4 + 𝑊𝑒𝑥 ) (15)

ɳ𝑒𝑥 = 𝑊𝑒𝑥 /(𝐸3 − 𝐸4 ) (16)

𝑊𝑝𝑢𝑚𝑝 = 𝑚𝑓 (ℎ2 − ℎ1 ) (17)

𝐼𝑝𝑢𝑚𝑝 = (𝐸1 + 𝑊𝑝𝑢𝑚𝑝 ) − 𝐸2 (18)

ɳ𝑝𝑢𝑚𝑝 = (𝐸2 − 𝐸1 ) /𝑊𝑝𝑢𝑚𝑝 (19)

Total work done by ORC system given as (20), thermal and exergy efficiency of ORC system as (21) and (22).

𝑊𝑛𝑒𝑡 = 𝑊𝑒𝑥 − 𝑊𝑝𝑢𝑚𝑝 (20)

ɳ𝑡ℎ = 𝑊𝑛𝑒𝑡 / 𝑄𝐸𝑣𝑎 (21)

ɳ𝐸𝑥 = 𝑊𝑛𝑒𝑡 / 𝐸𝑖𝑛 (22)

Where, Ein refers the total exergy input to the system by heat source can be written as (23).

𝐸𝑖𝑛 = (𝐸5 − 𝐸6 ) = 𝑚ℎ (ℎ5 − ℎ6 ) − 𝑇0 (𝑠5 − 𝑠6 ) (23)

4. MODIFICATIONS IN THE ORC SYSTEM


4.1. Role of pumps in ORC
Generally, raising the pressure of working fluid in ORC system pumps are plays a vital role without
a temperature change. Fluid pumps are devices with a negligible exergy destruction rate in ORC [14]. Some
researchers have developed a pumpless ORC system to recover lower grade waste heat. investigated an ORC,
control valves are replaced instead of a fluid pump, also two heat exchangers for a pre-expansion process, a
generator, and an expender aligned coaxially for the process of power generation [34]. Power output is
mainly depending on the inlet temperature of the evaporator, if it increases correspondingly evaporation
pressure will increase and it leads to higher power output. Exergy calculated to gravity-type pumpless ORC
system with three heat exchangers with a heat source at 90 °C, power fluctuation is mainly due to liquid
working fluid has more enthalpy at condensers compared to internal energy inside evaporator [35]. For
pumpless ORC have to follow some boundary conditions: heat leakage from heat exchangers to surrounding
should be neglected and the rate of flow of hot water insides the evaporator constant through the cycle.

4.2. Heat exchanger/evaporator


Heat exchangers are energy-efficient devices because it delivers a large amount of heat output
compared with electrical input [36]. It has a wide range of applications such as water heating, drying,
desalination, and space heating [37]. Integration heat exchangers performance is possible with renewable
resources like solar and geothermal [38]. During an exchange of heat between hot to cold fluids, according to
SLT, there is lots of chance to generate irreversibility, to overcome that an exergetic criterion was proposed
by optimum parameter and heat transfer units [39]. The efficiency of an optimum layout of a phase change
evaporator for stationary application with DWF (based on exergy) increases from 67 to 72.3% [40]–[44].
Prediction of exergy destruction (around 1.36 kW) for solar-based–direct expansion heat exchangers was
done at Calicut (India) climate conditions with R22 as working fluid [45]–[47]. Geometrical changes in the
heat exchanger can improve exergy efficiency, which means using a different types of heat exchangers (shell
and tube, double tube), using nanofluids to exchange heat [33], [48], [49]. Maddah et al. [50] were
experimented to prepare modified twisted type tapes for heat exchangers by using turbulent nanofluid
(SiO2-water) [51].

Int J Adv Appl Sci, Vol. 11, No. 1, March 2022: 1-10
Int J Adv Appl Sci ISSN: 2252-8814  5

4.3. The heat source for the evaporator


A clear observation is done that the source of heat temperature was varied between 80 to 330 °C,
according to temperature the power output from ORC also increases as [22]. The exhaust of the gas turbine
has contained a heavier amount of heat, and it was considered as a source of Waste heat to convert as
work/power output, this temperature ranged from 375 to 600 °C. The max exergy was found at 600 °C
because the mass of the fluid is a larger amount than the required heat load [31]. The maximum generated
power is 3.85 MW through the SCO2-ORC system with a heat source of coal-fired flue gases 200 to 300 °C
temperature range at evaporator [32]. By decreasing mass flow rate through a solar pond, the temperature
average (47 to 59.78 °C) at the exit of the pond increased for effective utilization as a heat source and got
15.94% of exergetic efficiency [18], [52]. Various heat sources namely geothermal (100-130 °C), solar low
heat (100-225 °C), engine exhaust waste heat (150-300 °C), high-temperature energy from solar(250-350 °C)
was carried on ORC and calculated 28% of expander efficiency at higher temperatures [53], [54]. An
investigation with two heat sources, lubricant oil at 120 °C and engine exhaust at 350 °C was done on the
ORC system for effective results [40], [55]. Three kinds of heat sources: saturated steam, hot water,
combinations of both in an ORC system with different pinch ranges was performed at a constant condenser
temperature of 40 °C and it is found that combined heat source will give lower exergy destruction and higher
efficiency [5], [56], [57]. The 63% of exergy efficiency recorded through ORC, which has a heat source of
hot water 150-300 °C and delivers 90 °C [58], [59]. Heat recovered from smelting furnace exhaust gas; ORC
worked at 38% of efficiency [60]–[62].

4.4. Expansion devices


In ORC system expansion devices are the key equipment, these are helpful to produce electrical
energy by using the kinetic energy of fluids. Here can classify as dynamic devices (turbines), positive-
displacement devices, and ejectors as shown in Figure 4. In the turbine, the pressurized working fluid is
converted as velocity energy (kinetic energy) through a nozzle. After that, it was transferred to turbine blades
to convert into electric power with the help of generator equipment [63]. Axial, radial type turbines are
available for this operation, generally axial have more than 7 capacities compared with radial turbines. A
method was proposed for radial type turbine for mobile ORC [37] with pentane and R245fa as working fluid
and found a 7.3% of efficiency difference [64]. Using R123 working fluid in an axial turbine with a thermal
efficiency of 10.5% and clinch 6.3 kW power output and 88% of isentropic efficiency [65]. Working of scroll
expander follows a similar principle of compressors, for orc the basic way is a better design to improve
efficiency. Few researchers explored a general tool for how to design a scroll expander for ORC [66] and
achieved 40% of isentropic efficiency. An investigation was done on the performance of ORC through 88
ml/r displacements of scroll expander and its influences on isentropic efficiency decrement from 0.72 to
0.41% [67], [68]. It can be observed that R123, R245fa working fluids have a high frequency to get a power
maximum of 3.75 kW [69], [70]. Similarly, a simulated model of screw expander by considering leakage,
heat transfer, and friction loss was developed [71]. Rotary type vane expander contains vane, rotor, and stator
in a closed volume and expansion process done when the rotor starts rotating. Power output varies from
1-10 kW in an ORC system because of the inlet port location of the working chamber, type of working fluid,
and expander dimensions [72], [73]. In the case of piston expander power output varied between 250-1150 W
in a micro-ORC system [74]. There is the absence of moving parts in the ejector, and ejector type ORC has
more exergetic efficiency compared with traditional ORC [75]. From survey turbines are suggested to large-
scale ORCs, screw expanders are better to use in medium-scale, ORCs in small and micro-size are enough
with scroll expanders (mainly for laboratories) [63], [76], [77].

4.5. Selection of working fluid


Selecting the best working fluid is a crucial task in ORC because heat duty from the heat source was
carried out throughout the cycle operation. So, it should have better performance and thermal props in the
expansion and condensation process. Also, withstand a wide range of temperature-pressure changes in
between the process.
Recovered heat from gas turbine waste with the help of recuperated ORC, siloxanes as working
fluid for effective energy efficiency, and high-temperature ranges (375 to 600 °C) [31], [78]. By using three
hydrofluorocarbons (HFC) refrigerants (R245a, R1234yf, R1234ze) with temperature difference 120 to
170 °C, nonetheless higher exergetic efficiency of 38.92% was found with R245a as a working fluid [21],
[79]. ORC-CO2 cycle operated with help of different working fluids, among those all for the optimized
process showed that ORC working with R152a as fluid extracted maximum exergetic efficiency of 13.08%
[80], [81]. Coal-fired flue gases converted as work by the SCO2-ORC system, working fluid heptane/R601
yields higher efficiency of 45.54% and 25.65% of destruction rate [32], [78]. R22 shows greater performance
among the eight working fluids in ORC to get cold energy in LNG [2], [82]. The R123 working fluid used in
the ORC system along with the internal heat exchanger (IHE) and feed liquid heater (FLH), obtained better
A review on efficiency improvement methods in organic Rankine … (Nagamalleswara Rao Kanidarapu )
6  ISSN: 2252-8814

energy and exergy values of 52.28 and 20.44% respectively [4], [83]. ORC operated using hydrocarbons for
cost analysis, cost of cyclohexane is very less among all working fluids (chlorofluorocarbons (CFCs),
hydrochlorofluorocarbons (HCFCs), HFCs, and natural fluids) [53], [84]. R245fa working fluid raises exergy
efficiency from 67 to 72.3% in ORC for stationary applications and also toluene working fluid 64.5 to 73%
for the same application of heat recovery [27]. Achieved 38.11%, 29.98%, and 15.93% exergy destruction in
low-pressure vapor generator (LPVG), high-pressure vapor generator (HPVG), and condenser (COND) in
ORC system with geothermal dual working fluid [14], [85]. ORC with saturated steam and hot water
combination heat source along with R245fa fluid shows better cycle efficiency of 9.4%. and also, the same
working fluid was carried on the pumpless ORC system and showed 232 W of power at output [5], [34], [86].
An exergy analysis was done on ORC by using various working fluids and got maximum power generation
of 1227 kW with help of R600a fluid [1], [87]. A statement found that mixing of working (70% n-
octane+30% n-pentane) results and the highest exergetic records [88]. In the heat recovering process from
smelting furnace exhaust among a wide variety of working fluids m-xylene has the most efficient results
[50], [89], [90]. Table 1 is showing general parameters while selecting working fluid and Table 2 is showing
basic thermodynamic properties of various working fluids.

Figure 4. Various types of expansion devices

Table 1. General parameters while selecting working fluid [91], [92]


Parameters Details
Environmental issues ODP, GWP, chemical stability, corrosion, flammability
Performance Good thermal range, heat-carrying for cycle optimization
Features The critical temperature, density, surface tension, specific heat capacity
Economic Easy availability and low cost

Table 2. Basic thermodynamic properties of various working fluids [35], [92], [93]
Substance Mass kg/kmol Tcr (k) Pcr (bar) ODP
R22 86.46 369.3 49.71 0.05
R134a 102.03 380 36.9 0.055
R152a 66.05 386.6 44.99 0
Propane 44.09 396.82 42.49 -
Isobutane 58.123 408.14 36.48 0
R245fa 134.048 427.2 36.4 0
R123 136.467 456.9 36.74 0.02
Isopentane 72.150 460.43 33.81 0
CO2 44.01 303.98 73.77 0
R227ea 170.03 374.75 29.25 0
R124 136.48 395.28 36.24 0.022
R141b 116.95 477.35 42.12 0.12
R143a - 345.857 37.6 0
R600 - 425.125 37.96 0
R601 - 469.7 33.7 0
Cyclohexane 84.12 554 40.8 0
Toluene 89.3 592 41.3 0

Int J Adv Appl Sci, Vol. 11, No. 1, March 2022: 1-10
Int J Adv Appl Sci ISSN: 2252-8814  7

5. CONCLUSION
Based on the above literature, this paper has an informative collection regarding how to perform an
exergy analysis, various methodologies that are available in the present market to improve each process or
entire ORC system. Some strong points like heat source temperature, evaporator inlet pressure, and selection
of working fluid play major roles for a better-optimized ORC system. It is found that a large amount of
exergy destruction at evaporator, expander, condenser, and pump respectively. For further research, this may
continue with different hear sources like solar energy, geothermal, industrial waste to recover waste heat
towards a better sustainable environment.

REFERENCES
[1] H. Y. Lee and K. H. Kim, “Energy and exergy analyses of a combined power cycle using the organic rankine cycle and the cold
energy of liquefied natural gas,” Entropy, vol. 17, no. 9, pp. 6412–6432, Sep. 2015, doi: 10.3390/e17096412.
[2] Ö. Kaşka, “Energy and exergy analysis of an organic Rankine for power generation from waste heat recovery in steel industry,”
Energy Convers. Manag., vol. 77, pp. 108–117, Jan. 2014, doi: 10.1016/j.enconman.2013.09.026.
[3] A. Sadreddini, M. A. Ashjari, M. Fani, and A. Mohammadi, “Thermodynamic analysis of a new cascade ORC and transcritical
CO2 cycle to recover energy from medium temperature heat source and liquefied natural gas,” Energy Convers. Manag., vol. 167,
pp. 9–20, Jul. 2018, doi: 10.1016/j.enconman.2018.04.093.
[4] F. I. Abam, E. B. Ekwe, S. O. Effiom, M. C. Ndukwu, T. A. Briggs, and C. H. Kadurumba, “Optimum exergetic performance
parameters and thermo-sustainability indicators of low-temperature modified organic Rankine cycles (ORCs),” Sustain. Energy
Technol. Assessments, vol. 30, pp. 91–104, Dec. 2018, doi: 10.1016/j.seta.2018.09.001.
[5] L. Jiang et al., “Investigation on a small-scale pumpless Organic Rankine Cycle (ORC) system driven by the low temperature heat
source,” Appl. Energy, vol. 195, pp. 478–486, Jun. 2017, doi: 10.1016/j.apenergy.2017.03.082.
[6] A. Nemati, H. Nami, F. Ranjbar, and M. Yari, “A comparative thermodynamic analysis of ORC and Kalina cycles for waste heat
recovery: A case study for CGAM cogeneration system,” Case Stud. Therm. Eng., vol. 9, pp. 1–13, Mar. 2017, doi:
10.1016/j.csite.2016.11.003.
[7] M. A. C. Munaaim, N. Hamidin, A. Ayob, N. M. Sani, and A. M. Salleh, “Application of micro hydro electric system attached to
effluent discharge point of sewerage treatment plant application of micro hydro electric system attached to effluent discharge
point of sewerage treatment plant,” in 5th Annual Conference on Sustainable Energy and Environmental Sciences (SEES 2016),
2016, doi: 10.5176/2251-189X_SEES16.33.
[8] A. Zeynali, A. Akbari, and M. Khalilian, “Investigation of the performance of modified organic Rankine cycles (ORCs) and
modified trilateral flash cycles (TFCs) assisted by a solar pond,” Sol. Energy, vol. 182, pp. 361–381, Apr. 2019, doi:
10.1016/j.solener.2019.03.001.
[9] L. Paradeshi, M. Mohanraj, M. Srinivas, and S. Jayaraj, “Exergy analysis of direct-expansion solar-assisted heat pumps working
with R22 and R433A,” J. Therm. Anal. Calorim., vol. 134, no. 3, pp. 2223–2237, Dec. 2018, doi: 10.1007/s10973-018-7424-3.
[10] B. Dai, K. Zhu, Y. Wang, Z. Sun, and Z. Liu, “Evaluation of organic Rankine cycle by using hydrocarbons as working fluids:
Advanced exergy and advanced exergoeconomic analyses,” Energy Convers. Manag., vol. 197, p. 111876, Oct. 2019, doi:
10.1016/j.enconman.2019.111876.
[11] G. Liao, J. E, F. Zhang, J. Chen, and E. Leng, “Advanced exergy analysis for organic Rankine cycle-based layout to recover waste
heat of flue gas,” Appl. Energy, vol. 266, p. 114891, May 2020, doi: 10.1016/j.apenergy.2020.114891.
[12] Y. Lee, J. Kim, U. Ahmed, C. Kim, and Y. W. Lee, “Multi-objective optimization of Organic Rankine Cycle (ORC) design
considering exergy efficiency and inherent safety for LNG cold energy utilization,” J. Loss Prev. Process Ind., vol. 58, pp. 90–
101, Mar. 2019, doi: 10.1016/j.jlp.2019.01.006.
[13] C. H. Han and K. H. Kim, “Entransy and exergy analyses for optimizations of heat-work conversion with carnot cycle,” J. Therm.
Sci., vol. 25, no. 3, pp. 242–249, Jun. 2016, doi: 10.1007/s11630-016-0856-9.
[14] C. Yan, A. Yang, I. L. Chien, S. Wei, W. Shen, and J. Ren, “Advanced exergy analysis of organic Rankine Cycles for Fischer-
Tropsch syngas production with parallel dry and steam methane reforming,” Energy Convers. Manag., vol. 199, p. 111963, Nov.
2019, doi: 10.1016/j.enconman.2019.111963.
[15] S. Karellas and K. Braimakis, “Energy-exergy analysis and economic investigation of a cogeneration and trigeneration ORC-VCC
hybrid system utilizing biomass fuel and solar power,” Energy Convers. Manag., vol. 107, pp. 103–113, Jan. 2016, doi:
10.1016/j.enconman.2015.06.080.
[16] Y. Zhao, G. Liu, L. Li, Q. Yang, B. Tang, and Y. Liu, “Expansion devices for organic Rankine cycle (ORC) using in low
temperature heat recovery: A review,” Energy Convers. Manag., vol. 199, p. 111944, Nov. 2019, doi:
10.1016/j.enconman.2019.111944.
[17] P. Gao, Z. X. Wang, L. W. Wang, and H. T. LU, “Technical feasibility of a gravity-type pumpless ORC system with one
evaporator and two condensers,” Appl. Therm. Eng., vol. 145, pp. 569–575, Dec. 2018, doi:
10.1016/j.applthermaleng.2018.09.049.
[18] M. Leveni, G. Manfrida, R. Cozzolino, and B. Mendecka, “Energy and exergy analysis of cold and power production from the
geothermal reservoir of Torre Alfina,” Energy, vol. 180, pp. 807–818, Aug. 2019, doi: 10.1016/j.energy.2019.05.102.
[19] D. Roy and S. Ghosh, “Energy and exergy analyses of an integrated biomass gasification combined cycle employing solid oxide
fuel cell and organic Rankine cycle,” Clean Technol. Environ. Policy, vol. 19, no. 6, pp. 1693–1709, Aug. 2017, doi:
10.1007/s10098-017-1358-5.
[20] S. Wu and Y. Li, “Exergy-economic criteria for evaluating heat exchanger performance,” J. Therm. Sci., vol. 10, no. 3, pp. 218–
222, Jul. 2001, doi: 10.1007/s11630-001-0022-9.
[21] L. Grosu, A. Marin, A. Dobrovicescu, and D. Queiros-Conde, “Exergy analysis of a solar combined cycle: organic Rankine cycle
and absorption cooling system,” Int. J. Energy Environ. Eng., vol. 7, no. 4, pp. 449–459, Dec. 2016, doi: 10.1007/s40095-015-
0168-y.
[22] R. Scaccabarozzi, M. Tavano, C. M. Invernizzi, and E. Martelli, “Comparison of working fluids and cycle optimization for heat
recovery ORCs from large internal combustion engines,” Energy, vol. 158, pp. 396–416, Sep. 2018, doi:
10.1016/j.energy.2018.06.017.
[23] B. Ma et al., “Exergy loss analysis on diesel methanol dual fuel engine under different operating parameters,” Appl. Energy, vol.
261, p. 114483, Mar. 2020, doi: 10.1016/j.apenergy.2019.114483.

A review on efficiency improvement methods in organic Rankine … (Nagamalleswara Rao Kanidarapu )


8  ISSN: 2252-8814

[24] J. Sun, Q. Liu, and Y. Duan, “Effects of reinjection temperature on thermodynamic performance of dual-pressure and single-
pressure geothermal ORCs,” Energy Procedia, vol. 158, pp. 6016–6023, Feb. 2019, doi: 10.1016/j.egypro.2019.01.517.
[25] A. Behzadi, E. Gholamian, E. Houshfar, and A. Habibollahzade, “Multi-objective optimization and exergoeconomic analysis of
waste heat recovery from Tehran’s waste-to-energy plant integrated with an ORC unit,” Energy, vol. 160, pp. 1055–1068, Oct.
2018, doi: 10.1016/j.energy.2018.07.074.
[26] N. F. Tumen Ozdil, M. R. Segmen, and A. Tantekin, “Thermodynamic analysis of an organic Rankine cycle (ORC) based on
industrial data,” Appl. Therm. Eng., vol. 91, pp. 43–52, Dec. 2015, doi: 10.1016/j.applthermaleng.2015.07.079.
[27] T. B. Nur and Sunoto, “Exergy analysis of biomass organic Rankine cycle for power generation,” IOP Conf. Ser. Mater. Sci.
Eng., vol. 309, no. 1, p. 12057, Feb. 2018, doi: 10.1088/1757-899X/309/1/012057.
[28] J. Zhao et al., “How to rapidly predict the performance of ORC: Optimal empirical correlation based on cycle separation,” Energy
Convers. Manag., vol. 188, pp. 86–93, May 2019, doi: 10.1016/j.enconman.2019.02.095.
[29] N. Chaiyat, S. Chaongew, P. Ondokmai, and P. Makarkard, “Levelized energy and exergy costings per life cycle assessment of a
combined cooling, heating, power and tourism system of the San Kamphaeng hot spring, Thailand,” Renew. Energy, vol. 146, pp.
828–842, Feb. 2020, doi: 10.1016/j.renene.2019.07.028.
[30] H. Khosravi, G. R. Salehi, and M. T. Azad, “Design of structure and optimization of organic Rankine cycle for heat recovery from
gas turbine: The use of 4E, advanced exergy and advanced exergoeconomic analysis,” Appl. Therm. Eng., vol. 147, pp. 272–290,
Jan. 2019, doi: 10.1016/j.applthermaleng.2018.09.128.
[31] A. Marin, A. Dobrovicescu, L. Grosu, and A. Gheorghian, “Energy and exergy analysis of an Organic Rankine Cycle,” UPB Sci.
Bull. Ser. D Mech. Eng., vol. 76, no. 4, pp. 127–136, 2014, doi: 136. https://doi.org/10.9790/1684-12128592.
[32] H. Nami, A. Nemati, and F. Jabbari Fard, “Conventional and advanced exergy analyses of a geothermal driven dual fluid organic
Rankine cycle (ORC),” Appl. Therm. Eng., vol. 122, pp. 59–70, Jul. 2017, doi: 10.1016/j.applthermaleng.2017.05.011.
[33] H. Nami, I. S. Ertesvåg, R. Agromayor, L. Riboldi, and L. O. Nord, “Gas turbine exhaust gas heat recovery by organic Rankine
cycles (ORC) for offshore combined heat and power applications - Energy and exergy analysis,” Energy, vol. 165, pp. 1060–
1071, Dec. 2018, doi: 10.1016/j.energy.2018.10.034.
[34] J. C. Gutierrez, G. V. Ochoa, and J. Duarte-Forero, “A comparative study of the energy, exergetic and thermo-economic
performance of a novelty combined Brayton S-CO2-ORC configurations as bottoming cycles,” Heliyon, vol. 6, no. 7, 2020, doi:
https://doi.org/10.1016/j.heliyon.2020.e04459.
[35] H. Kianfard, S. Khalilarya, and S. Jafarmadar, “Exergy and exergoeconomic evaluation of hydrogen and distilled water
production via combination of PEM electrolyzer, RO desalination unit and geothermal driven dual fluid ORC,” Energy Convers.
Manag., vol. 177, pp. 339–349, Dec. 2018, doi: 10.1016/j.enconman.2018.09.057.
[36] O. Ozgener and A. Hepbasli, “Modeling and performance evaluation of ground source (geothermal) heat pump systems,” Energy
Build., vol. 39, no. 1, pp. 66–75, Jan. 2007, doi: 10.1016/j.enbuild.2006.04.019.
[37] O. Ozgener and A. Hepbasli, “Performance analysis of a solar-assisted ground-source heat pump system for greenhouse heating:
An experimental study,” Build. Environ., vol. 40, no. 8, pp. 1040–1050, Aug. 2005, doi: 10.1016/j.buildenv.2004.08.030.
[38] K. Rahbar, S. Mahmoud, R. K. Al-Dadah, N. Moazami, and A. S. B. Ennil, “Preliminary mean-line design and optimization of a
radial turbo-expander for waste heat recovery using organic Rankine cycle,” Energy Procedia, vol. 75, pp. 860–866, Aug. 2015,
doi: 10.1016/j.egypro.2015.07.188.
[39] A. A. Shayesteh, O. Koohshekan, A. Ghasemi, M. Nemati, and H. Mokhtari, “Determination of the ORC-RO system optimum
parameters based on 4E analysis; Water–Energy-Environment nexus,” Energy Convers. Manag., vol. 183, pp. 772–790, Mar.
2019, doi: 10.1016/j.enconman.2018.12.119.
[40] H. You, J. Han, and Y. Liu, “Performance assessment of a CCHP and multi-effect desalination system based on GT/ORC with
inlet air precooling,” Energy, vol. 185, pp. 286–298, Oct. 2019, doi: 10.1016/j.energy.2019.06.177.
[41] A. Sadreddini, M. Fani, M. Ashjari Aghdam, and A. Mohammadi, “Exergy analysis and optimization of a CCHP system
composed of compressed air energy storage system and ORC cycle,” Energy Convers. Manag., vol. 157, pp. 111–122, Feb. 2018,
doi: 10.1016/j.enconman.2017.11.055.
[42] F. Mohammadkhani, M. Yari, and F. Ranjbar, “A zero-dimensional model for simulation of a Diesel engine and exergoeconomic
analysis of waste heat recovery from its exhaust and coolant employing a high-temperature Kalina cycle,” Energy Convers.
Manag., vol. 198, p. 111782, Oct. 2019, doi: 10.1016/j.enconman.2019.111782.
[43] O. A. Oyewunmi, C. J. W. Kirmse, A. M. Pantaleo, and C. N. Markides, “Performance of working-fluid mixtures in ORC-CHP
systems for different heat-demand segments and heat-recovery temperature levels,” Energy Convers. Manag., vol. 148, pp. 1508–
1524, Sep. 2017, doi: 10.1016/j.enconman.2017.05.078.
[44] E. Sevinchan, I. Dincer, and H. Lang, “Energy and exergy analyses of a biogas driven multigenerational system,” Energy, vol.
166, pp. 715–723, Jan. 2019, doi: 10.1016/j.energy.2018.10.085.
[45] P. Chen, G. He, Y. Gao, X. Zhao, and D. Cai, “Conventional and advanced exergy analysis of an air-cooled type of absorption-
ejection refrigeration cycle with R290-mineral oil as the working pair,” Energy Convers. Manag., vol. 210, p. 112703, Apr. 2020,
doi: 10.1016/j.enconman.2020.112703.
[46] R. Kong, T. Deethayat, A. Asanakham, N. Vorayos, and T. Kiatsiriroat, “Thermodynamic performance analysis of a R245fa
organic Rankine cycle (ORC) with different kinds of heat sources at evaporator,” Case Stud. Therm. Eng., vol. 13, p. 100385,
Mar. 2019, doi: 10.1016/j.csite.2018.100385.
[47] S. Marandi, F. Mohammadkhani, and M. Yari, “An efficient auxiliary power generation system for exploiting hydrogen boil-off
gas (BOG) cold exergy based on PEM fuel cell and two-stage ORC: Thermodynamic and exergoeconomic viewpoints,” Energy
Convers. Manag., vol. 195, pp. 502–518, Sep. 2019, doi: 10.1016/j.enconman.2019.05.018.
[48] S. O. Mert and A. Reis, “Experimental performance investigation of a shell and tube heat exchanger by exergy based sensitivity
analysis,” Heat Mass Transf. und Stoffuebertragung, vol. 52, no. 6, pp. 1117–1123, Jun. 2016, doi: 10.1007/s00231-015-1636-x.
[49] A. Atiz, H. Karakilcik, M. Erden, and M. Karakilcik, “Investigation energy, exergy and electricity production performance of an
integrated system based on a low-temperature geothermal resource and solar energy,” Energy Convers. Manag., vol. 195, pp.
798–809, Sep. 2019, doi: 10.1016/j.enconman.2019.05.056.
[50] H. Maddah, N. Ghasemi, B. Keyvani, and R. Cheraghali, “Experimental and numerical study of nanofluid in heat exchanger fitted
by modified twisted tape: exergy analysis and ANN prediction model,” Heat Mass Transf. und Stoffuebertragung, vol. 53, no. 4,
pp. 1413–1423, Apr. 2017, doi: 10.1007/s00231-016-1906-2.
[51] J. M. Montelongo-Luna, B. R. Young, and W. Y. Svrcek, “An open source exergy calculator tool,” Proc. Can. Eng. Educ. Assoc.,
Aug. 2011, doi: 10.24908/pceea.v0i0.3960.

Int J Adv Appl Sci, Vol. 11, No. 1, March 2022: 1-10
Int J Adv Appl Sci ISSN: 2252-8814  9

[52] Wang Yanan, W. Jiekang, and M. Xiaoming, “Exergy Analysis of Cogeneration System for the Wind–Solar–Gas Turbine
Combined Supply,” Appl. Sol. Energy (English Transl. Geliotekhnika), vol. 54, no. 5, pp. 369–375, Nov. 2018, doi:
10.3103/S0003701X18050213.
[53] J. Galindo, S. Ruiz, V. Dolz, and L. Royo-Pascual, “Advanced exergy analysis for a bottoming organic rankine cycle coupled to
an internal combustion engine,” Energy Convers. Manag., vol. 126, pp. 217–227, Oct. 2016, doi:
10.1016/j.enconman.2016.07.080.
[54] Z. Wu, Y. Zhang, and Y. Sheng, “Energy, exergy, economic (3E) analysis and multi-objective optimization of a novel dual
functional integration system,” Energy Convers. Manag., vol. 199, p. 111962, Nov. 2019, doi: 10.1016/j.enconman.2019.111962.
[55] S. Mohammadzadeh Bina, S. Jalilinasrabady, and H. Fujii, “Energy, economic and environmental (3E) aspects of internal heat
exchanger for ORC geothermal power plants,” Energy, vol. 140, pp. 1096–1106, Dec. 2017, doi: 10.1016/j.energy.2017.09.045.
[56] Ł. Tomków and M. Cholewiński, “Improvement of the LNG (liquid natural gas) regasification efficiency by utilizing the cold
exergy with a coupled absorption - ORC (organic Rankine cycle),” Energy, vol. 87, pp. 645–653, Jul. 2015, doi:
10.1016/j.energy.2015.05.041.
[57] T. C. Roumpedakis, T. Christou, E. Monokrousou, K. Braimakis, and S. Karellas, “Integrated ORC-Adsorption cycle: A first and
second law analysis of potential configurations,” Energy, vol. 179, pp. 46–58, Jul. 2019, doi: 10.1016/j.energy.2019.04.069.
[58] E. Akrami, A. Chitsaz, P. Ghamari, and S. M. S. Mahmoudi, “Energy and exergy evaluation of a tri-generation system driven by
the geothermal energy,” J. Mech. Sci. Technol., vol. 31, no. 1, pp. 401–408, Jan. 2017, doi: 10.1007/s12206-016-1242-y.
[59] W. Sun, X. Yue, and Y. Wang, “Exergy efficiency analysis of ORC (Organic Rankine Cycle) and ORC-based combined cycles
driven by low-temperature waste heat,” Energy Convers. Manag., vol. 135, pp. 63–73, Mar. 2017, doi:
10.1016/j.enconman.2016.12.042.
[60] M. Ebrahimi and K. Ahookhosh, “Integrated energy-exergy optimization of a novel micro-CCHP cycle based on MGT-ORC and
steam ejector refrigerator,” Appl. Therm. Eng., vol. 102, pp. 1206–1218, Jun. 2016, doi: 10.1016/j.applthermaleng.2016.04.015.
[61] M. Mirzaei, M. H. Ahmadi, M. Mobin, M. A. Nazari, and R. Alayi, “Energy, exergy and economics analysis of an ORC working
with several fluids and utilizes smelting furnace gases as heat source,” Therm. Sci. Eng. Prog., vol. 5, pp. 230–237, Mar. 2018,
doi: 10.1016/j.tsep.2017.11.011.
[62] S. Wang, C. Liu, L. Liu, X. Xu, and C. Zhang, “Ecological cumulative exergy consumption analysis of organic Rankine cycle for
waste heat power generation,” J. Clean. Prod., vol. 218, pp. 543–554, May 2019, doi: 10.1016/j.jclepro.2019.02.003.
[63] Z. Li, R. Huang, Y. Lu, A. P. Roskilly, and X. Yu, “Analysis of a combined trilateral cycle - Organic Rankine cycle (TLC-ORC)
system for waste heat recovery,” Energy Procedia, vol. 158, pp. 1786–1791, Feb. 2019, doi: 10.1016/j.egypro.2019.01.421.
[64] A. W. Costall, A. Gonzalez Hernandez, P. J. Newton, and R. F. Martinez-Botas, “Design methodology for radial turbo expanders
in mobile organic Rankine cycle applications,” Appl. Energy, vol. 157, pp. 729–743, Nov. 2015, doi:
10.1016/j.apenergy.2015.02.072.
[65] S. Y. Cho, C. H. Cho, and S. K. Choi, “Experiment and cycle analysis on a partially admitted axial-type turbine used in the
organic Rankine cycle,” Energy, vol. 90, pp. 643–651, Oct. 2015, doi: 10.1016/j.energy.2015.07.092.
[66] P. Garg, G. M. Karthik, P. Kumar, and P. Kumar, “Development of a generic tool to design scroll expanders for ORC
applications,” Appl. Therm. Eng., vol. 109, pp. 878–888, Oct. 2016, doi: 10.1016/j.applthermaleng.2016.06.047.
[67] S. C. Kaushik and O. K. Singh, “Estimation of chemical exergy of solid, liquid and gaseous fuels used in thermal power plants,”
J. Therm. Anal. Calorim., vol. 115, no. 1, pp. 903–908, Jan. 2014, doi: 10.1007/s10973-013-3323-9.
[68] R. Kumar, “A critical review on energy, exergy, exergoeconomic and economic (4-E) analysis of thermal power plants,” Eng. Sci.
Technol. an Int. J., vol. 20, no. 1, pp. 283–292, Feb. 2017, doi: 10.1016/j.jestch.2016.08.018.
[69] B. Fajar TK, D. Panca N, and G. Wicaksono, “Unsteady exergy analysis for a recycle and recovery machine,” Int. J. Refrig., vol.
87, pp. 1–9, Mar. 2018, doi: 10.1016/j.ijrefrig.2017.10.002.
[70] M. S. Orosz, A. V Mueller, B. J. Dechesne, and H. F. Hemond, “Geometric design of scroll expanders optimized for small
organic rankine cycles,” J. Eng. Gas Turbines Power, vol. 135, no. 4, Apr. 2013, doi: 10.1115/1.4023112.
[71] L. Shen et al., “A study of clearance height on the performance of single-screw expanders in small-scale organic Rankine cycles,”
Energy, vol. 153, pp. 45–55, Jun. 2018, doi: 10.1016/j.energy.2018.02.004.
[72] M. Mohanraj, Y. Belyayev, S. Jayaraj, and A. Kaltayev, “Research and developments on solar assisted compression heat pump
systems – A comprehensive review (Part A: Modeling and modifications),” Renew. Sustain. Energy Rev., vol. 83, pp. 90–123,
Mar. 2018, doi: 10.1016/j.rser.2017.08.022.
[73] J. Mascuch, V. Novotny, V. Vodicka, and Z. Zeleny, “Towards development of 1-10 kW pilot ORC units operating with
hexamethyldisiloxane and using rotary vane expander,” Energy Procedia, vol. 129, pp. 826–833, Sep. 2017, doi:
10.1016/j.egypro.2017.09.196.
[74] M. Bianchi et al., “Experimental analysis of a micro-ORC driven by piston expander for low-grade heat recovery,” Appl. Therm.
Eng., vol. 148, pp. 1278–1291, Feb. 2019, doi: 10.1016/j.applthermaleng.2018.12.019.
[75] X. Li, X. Li, and Q. Zhang, “The first and second law analysis on an organic Rankine cycle with ejector,” Sol. Energy, vol. 93, pp.
100–108, Jul. 2013, doi: 10.1016/j.solener.2013.04.003.
[76] A. Razmi, M. Soltani, and M. Torabi, “Investigation of an efficient and environmentally-friendly CCHP system based on CAES,
ORC and compression-absorption refrigeration cycle: Energy and exergy analysis,” Energy Convers. Manag., vol. 195, pp. 1199–
1211, Sep. 2019, doi: 10.1016/j.enconman.2019.05.065.
[77] P. Gao, L. Jiang, L. W. Wang, R. Z. Wang, and F. P. Song, “Simulation and experiments on an ORC system with different scroll
expanders based on energy and exergy analysis,” Appl. Therm. Eng., vol. 75, pp. 880–888, Jan. 2015, doi:
10.1016/j.applthermaleng.2014.10.044.
[78] L. Shi, G. Shu, H. Tian, and S. Deng, “A review of modified organic Rankine cycles (ORCs) for internal combustion engine
waste heat recovery (ICE-WHR),” Renew. Sustain. Energy Rev., vol. 92, pp. 95–110, Sep. 2018, doi: 10.1016/j.rser.2018.04.023.
[79] G. Han, B. Hua, Q. Chen, and Q. Yin, “Generalized expression of exergy in the thermodynamics,” Sci. China Ser. A Math., vol.
45, no. 1, pp. 70–75, 2002, doi: 10.1360/02ys9007.
[80] F. Ünal, G. Temir, and H. Köten, “Energy, exergy and exergoeconomic analysis of solar-assisted vertical ground source heat
pump system for heating season,” J. Mech. Sci. Technol., vol. 32, no. 8, pp. 3929–3942, Aug. 2018, doi: 10.1007/s12206-018-
0744-1.
[81] H. Yağli, Y. Koç, A. Koç, A. Görgülü, and A. Tandiroğlu, “Parametric optimization and exergetic analysis comparison of
subcritical and supercritical organic Rankine cycle (ORC) for biogas fuelled combined heat and power (CHP) engine exhaust gas
waste heat,” Energy, vol. 111, pp. 923–932, Sep. 2016, doi: 10.1016/j.energy.2016.05.119.
[82] V. M. V Padmanabhan and S. K. Palanisamy, “Exergy efficiency and irreversibility comparison of R22, R134a, R290 and R407C
to replace R22 in an air conditioning system,” J. Mech. Sci. Technol., vol. 27, no. 3, pp. 917–926, Mar. 2013, doi:
10.1007/s12206-013-0103-1.
A review on efficiency improvement methods in organic Rankine … (Nagamalleswara Rao Kanidarapu )
10  ISSN: 2252-8814

[83] S. K. Som and A. Datta, “Thermodynamic irreversibilities and exergy balance in combustion processes,” Prog. Energy Combust.
Sci., vol. 34, no. 3, pp. 351–376, Jun. 2008, doi: 10.1016/j.pecs.2007.09.001.
[84] B. K. Kanabar and B. M. Ramani, “Energy and exergy analysis of vapour absorption refrigeration cycle—a review,” J. Inst. Eng.
Ser. C, vol. 97, no. 3, pp. 479–491, Jul. 2016, doi: 10.1007/s40032-015-0216-7.
[85] M. Ameri, H. Mokhtari, and M. Bahrami, “Energy, exergy, exergoeconomic and environmental (4E) optimization of a large steam
power plant: A case study,” Iran. J. Sci. Technol. - Trans. Mech. Eng., vol. 40, no. 1, pp. 11–20, Mar. 2016, doi: 10.1007/s40997-
016-0002-z.
[86] D. Tiwari, “Energy and exergy analysis of organic Rankine cycle using alternative working fluids,” IOSR J. Mech. Civ. Eng., vol.
12, no. 01, pp. 85–92, Mar. 2017, doi: 10.9790/1684-1201028592.
[87] K. Manjunath and S. C. Kaushik, “Second law thermodynamic study of heat exchangers: A review,” Renew. Sustain. Energy Rev.,
vol. 40, pp. 348–374, Dec. 2014, doi: 10.1016/j.rser.2014.07.186.
[88] S. Baral, D. Kim, E. Yun, and K. C. Kim, “Application of micro hydro electric system attached to effluent aplication of micro
hydro electric system attached,” 2016.
[89] N. Akbari, “Introducing and 3E (energy, exergy, economic) analysis of an integrated transcritical CO2 Rankine cycle, Stirling
power cycle and LNG regasification process,” Appl. Therm. Eng., vol. 140, pp. 442–454, Jul. 2018, doi:
10.1016/j.applthermaleng.2018.05.073.
[90] R. Prabakaran and D. Mohan Lal, “A novel exergy based charge optimisation for a mobile air conditioning system: An
experimental study,” J. Therm. Anal. Calorim., vol. 132, no. 2, pp. 1241–1252, May 2018, doi: 10.1007/s10973-018-6998-0.
[91] S. Seyedkavoosi, S. Javan, and K. Kota, “Exergy-based optimization of an organic Rankine cycle (ORC) for waste heat recovery
from an internal combustion engine (ICE),” Appl. Therm. Eng., vol. 126, pp. 447–457, Nov. 2017, doi:
10.1016/j.applthermaleng.2017.07.124.
[92] K. Thurairaja, A. Wijewardane, S. Jayasekara, and C. Ranasinghe, “Working Fluid Selection and Performance Evaluation of
ORC,” Energy Procedia, vol. 156, pp. 244–248, Jan. 2019, doi: 10.1016/j.egypro.2018.11.136.
[93] B. Navongxay and N. Chaiyat, “Energy and exergy costings of organic Rankine cycle integrated with absorption system,” Appl.
Therm. Eng., vol. 152, pp. 67–78, Apr. 2019, doi: 10.1016/j.applthermaleng.2019.02.018.

BIOGRAPHIES OF AUTHORS

Gollangi Raju is a research scholar in the School of Mechanical Engineering,


Vellore Institute of Technology University, Vellore, Tamilnadu, India. His research topics
include energy, exergy analysis and process safety of industrial processes using ASPEN
HYSYS software. He can be contacted at email: gollangi.raju2019@vitstudent.ac.in.

Nagamalleswara Rao Kanidarapu is an associate professor in the Centre for


Disaster Mitigation and Management, Vellore Institute of Technology University, Vellore,
Tamilnadu, India. His research topics include techno-economic analysis, energy analysis
and process safety of industrial processes using ASPEN PLUS software. He can be
contacted at email: aspenmodels@gmail.com, nagamalleswara.rao@vit.ac.in.

Int J Adv Appl Sci, Vol. 11, No. 1, March 2022: 1-10

You might also like