0% found this document useful (0 votes)
6 views

cao2021

This document presents an advanced modification to a conventional solar-powered combined cooling, heating, and power (CCHP) system with thermal energy storage, focusing on exergetic and economic assessments. The study employs a genetic algorithm for multi-objective optimization, examining the impact of various design parameters on system performance. Results indicate that the modified system enhances overall exergy efficiency and reduces costs compared to the conventional setup under different solar operational modes.

Uploaded by

allah125cc
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
6 views

cao2021

This document presents an advanced modification to a conventional solar-powered combined cooling, heating, and power (CCHP) system with thermal energy storage, focusing on exergetic and economic assessments. The study employs a genetic algorithm for multi-objective optimization, examining the impact of various design parameters on system performance. Results indicate that the modified system enhances overall exergy efficiency and reduces costs compared to the conventional setup under different solar operational modes.

Uploaded by

allah125cc
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 58

Journal Pre-proof

Exergetic and economic assessments and multi-objective optimization of a modified


solar-powered CCHP system with thermal energy storage

Yan Cao, Hayder A. Dhahad, Hussein Togun, Maghsoud Abdollahi Haghghi, Hassan
Athari, Abdeliazim Mustafa Mohamed
PII: S2352-7102(21)00560-X
DOI: https://doi.org/10.1016/j.jobe.2021.102702
Reference: JOBE 102702

To appear in: Journal of Building Engineering

Received Date: 22 February 2021


Revised Date: 15 April 2021
Accepted Date: 6 May 2021

Please cite this article as: Y. Cao, H.A. Dhahad, H. Togun, M.A. Haghghi, H. Athari, A.M. Mohamed,
Exergetic and economic assessments and multi-objective optimization of a modified solar-powered
CCHP system with thermal energy storage, Journal of Building Engineering, https://doi.org/10.1016/
j.jobe.2021.102702.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Published by Elsevier Ltd.


Authors’ contributions

Yan Cao
Supervision, Conceptualization, Methodology, Formal analysis, Writing - Review & Editing

Hayder A. Dhahad
Conceptualization, Methodology, Software, Data curation, Formal analysis, Writing - Review &
Editing

Hussein Togun
Supervision, Methodology, Investigation, Resources, Writing - Review & Editing

Maghsoud Abdollahi Haghghi

f
oo
Conceptualization, Project administration, Software, Validation, Writing - Original Draft

Hassan Athari

r
-p
Resources, Investigation, Writing - Original Draft
re
Abdeliazim Mustafa Mohamed
Resources, Investigation, Writing - Review & Editing
lP
na
ur
Jo
Exergetic and economic assessments and multi-objective optimization of a

modified solar-powered CCHP system with thermal energy storage

Yan Cao1, Hayder A. Dhahad2, Hussein Togun3, Maghsoud Abdollahi Haghghi4,5*, Hassan

Athari5, Abdeliazim Mustafa Mohamed6

1. School of Mechatronic Engineering, Xi’an Technological University, Xi’an, 710021 China

2. Mechanical Engineering Department, University of Technology, Baghdad, Iraq

3. Department of Biomedical Engineering, University of Thi-Qar, 64001 Nassiriya, Iraq

f
4. Department of Mechanical Engineering, School of Engineering, Urmia University, Urmia, Iran

oo
5. Department of Mechanical Engineering, Elm-o-Fann University College of Science and Technology, Urmia,

r
Iran
-p
6. Department of Civil Engineering, College of Engineering, Prince Sattam bin Abdulaziz University, Alkharj
re
11942, Saudi Arabia
lP

*
Corresponding author:

Maghsoud Abdollahi Haghghi


na

E–mail address: [email protected]


ur

Tel: +98 914 943 8521


Jo

1
Abstract

An advanced modification is applied to a conventional solar-powered trigeneration

application for combined cooling, heating, and power (CCHP) generation. Accordingly, its

optimization is described for various cases employing a genetic algorithm and adopting

exergetic and economic approaches. In this way, the net generated electricity, the overall

exergy efficiency, and total cost per unit exergy are chosen as the objective functions. The

conventional CCHP setup comprises a solar subsystem including high-temperature solar

f
collectors in arrangement with hot and cold storage tanks, an organic Rankine cycle (ORC), a

oo
heating production heat exchanger, and a single-effect absorption chiller. In addition to the

r
-p
solar subsystem, the modified setup embraces three heating production heat exchangers, a
re
double-effect absorption chiller, and two regenerative ORCs. Indeed, the use of regeneration
lP

in the ORC provides the possibility of establishing two regenerative ORCs (RORCs) along

whit a multi-heat recovery situation at the defined framework. Three different solar
na

operational modes, i.e., high, low, and zero radiation statuses during a day are considered to
ur

analyze and compare the conventional and modified systems under identical conditions
Jo

through a comprehensive parametric study. Taking into account the parametric study results,

the oil mass ratio of the solar subsystem has a remarkable impact on the exergy and cost

criteria of the system throughout the day. Moreover, the modification process provides the

best exergetic and economic performances. This capability ascends the overall exergy

efficiency by 0.8, 0.6, and 0.4 percent-points, and the exergoeconomic factor by 0.4, 0.9, and

2.1 percent-points for the aforementioned solar operation modes, respectively.

Keywords: Modified solar system; Energy storage; High-temperature solar collectors; Multi-

objective optimization; Exergoeconomic

1. Introduction

2
The development of conventional energy conversion cycles driven by solar energy for city

planning, which are of lower exergy efficiency in comparison with the majority of other

energy conversion techniques, is directly related to economic aspects of the energy sector [1].

Thus, the modification of the configuration of such systems can bring about improvements in

not only their thermodynamic performance, but also their economics [2]. In this regard, the

evaluation and modification of energy conversion systems composed of parabolic trough

solar collectors (PTSCs), which are widely employed in plants with medium and high

production capacities, are of paramount importance [3]. The performance of such systems has

f
oo
been recently scrutinized by researchers of which some are elucidated below.

r
The energy and exergy performances of parabolic trough solar collectors (PTSCs) in a solar-
-p
based plant were investigated by Al-Zahrani and Dincer [4] for various design and operating
re
conditions. They reported that the energy and exergy efficiencies of the designed system were
lP

found to be 66.35% and 38.51%, respectively. A power production system producing 57 kW


na

of electricity using solar radiation and fuel combustion was assessed from the

exergoeconomic standpoint by Cavalcanti and Motta [5]. The electrical cost per unit exergy
ur

(CPUE) of the system was determined to be 10.42 $/GJ. Habibi et al. [6] analyzed and
Jo

optimized a system constituted of a solar-driven ammonia-water regenerative Rankine cycle

and a LNG cold energy recovery system from the thermoeconomic viewpoint. The total cost

rate was found to be 1181 $/h. Oyekale et al. [7] conducted an optimization of a hybrid

system based on solar and biomass energies, with cost between 10.5 and 12.1 cents/kWh.

Likewise, the exergy efficiency of the system was obtained to be 10.7%.

Ashouri et al. [8] examined and optimized a PTSC-based system utilizing a thermal storage

tank integrated with an ORC from the exergoeconomic viewpoints. The system was capable

of reaching an exergy efficiency of 22.7% and a product cost rate of 2.66 M$/year. Calise et

al. [9] evaluated a solar-geothermal hybrid electricity production system comprised of an

3
ORC and PTSCs. The exergy efficiency and electrical cost rate of this system were evaluated

as 45.44% and 737×103 €/year, respectively. The thermoeconomic performance of a system

consisting of a two-stage regenerative ORC with PTSCs was examined by Mehrpooya et al.

[10]. To absorb the heat duty of the condenser, they used the cold energy of liquefied natural

gas (LNG) and found that the system can reach an energy efficiency of 19.59% and a product

cost rate of 3.88 M$/year. Zare and Moalemian [11] assessed the thermodynamic and

exergoeconomic aspects of an integrated system composed of a Kalina cycle and PTSCs. The

findings demonstrated that the overall exergy efficiency of the system and generated

f
oo
electricity were individually calculated as 14% and almost 450 kW at a solar radiation density

r
of 1000 W/m2 with a CPUE of 0.118 $/kWh.
-p
An electricity, heating, and hydrogen production application using PTSCs was assessed by
re
Mohammadi and Mehrpooya [12]. Their study aimed to utilize the electrical power generated
lP

in the ORC for driving the electrolyzer. Hence, the exergy efficiency of the system was found
na

to be 26.81%. A gas turbine trigeneration plant was modified through integration with PTSCs

by Dabwan et al. [13], producing 45461 m3/day of freshwater, 2300 kg/s of chilled water, and
ur

360 MW of electrical power. The levelized electrical cost of the modified setup (MS) was
Jo

calculated between 5.7 and 7.6 cents/kWh. Calise et al. [14] evaluated a solar-geothermal

multi-generation plant, consisting of an organic Rankine cycle (ORC) driven by geothermal

energy and a PTSC field, which produces electricity, heating, cooling, and freshwater.

Alirahmi et al. [15] studied a geothermal and PTSC-based poly-generation plant producing

electricity, cooling, freshwater, hydrogen, and heating. The system’s optimum exergy

efficiency and CPUE were reported to be 29.95% and 129.7 $/GJ, individually. Delpisheh et

al. [16,17] designed, studied and optimized a trigeneration application embracing a PTSC

field and solar storage tanks incorporated with an ORC, a desalination system, and a low-

temperature electrolyzer for simultaneous production of electricity, freshwater, and hydrogen

4
from energy and exergy [16] and exergoeconomic [17] viewpoints. Their application was

projected in three solar radiation modes of operation during a day based on global solar data.

The performance [18], exergy [19], and exergoeconomic [20] facets of a combined cooling,

heating, and power (CCHP) system embracing PTSCs coupled with an ORC and a single-

effect absorption chiller (SEACH) were investigated by Al-Sulaiman et al. [18-20] under low,

high and zero radiation modes. The exergy efficiency of the system for the aforementioned

radiation modes was 20%, 8%, and 7%, respectively, while the CPUE of the system for the

considered modes was correspondingly 21 $/GJ, 15 $/GJ, and 23 $/GJ. Haghghi et al. [21,22]

f
oo
carried out a series of case studies to analyze the thermodynamic [21] and exergoeconomic

r
[22] aspects of a PTSC-based CCHP system that meets the demands of a university building
-p
in Iran under the three radiation modes stipulated above. For these respective radiation
re
modes, the system’s exergy efficiencies were 17%, 8.3%, and 17% and the investment cost
lP

rates (ICRs) were 147.7 $/h, 210.8 $/h, and 66.4 $/h. The exergoeconomic aspect of
na

integrating PTSCs into a CCHP system was optimized by Baghernejad et al. [23]. Their

findings revealed that such integration can result in a significant reduction, of 11.5%, in the
ur

products unit costs and considerable improvement, of 11.7%, in the system’s exergy
Jo

efficiency. A trigeneration system composed of two ORC, a gas turbine, an absorption heat

pump, and PTSCs was assessed by Khalid et al. [24]. The levelized electricity cost of the

system was 0.12 €/kWh. Wang et al. [25] performed thermodynamic, exergoeconomic, and

environmental evaluations of a CCHP system comprised of a Brayton cycle, an absorption

chiller/heater, a heat storage tank, and PTSCs. The findings unveiled that the exergy

efficiency for the cooling and heating operation modes was equal to 24.9% and 25.7%,

respectively. A solar-based trigeneration system in which PTSCs were integrated with a

storage tank so as to feed an ORC rejecting heat to an absorption heat pump was analyzed

and optimized from energetic, exergetic, and financial viewpoints by Bellos et al. [26]. A

5
thermodynamic-economic-environmental investigation of a solar-based CCHP system using

PTSCs was performed by Dabwan and Pei [27]. Their results showed that the integration of

126 hectares of PTSCs with the trigeneration plant bring about the most optimal system,

resulting in a levelized electricity cost of 5.75 cents/kWh.

The application of PTSCs as environmentally-friendly technologies that are accompanied by

no or little detrimental emissions, and subsequent deleterious environmental impacts, can be

an important option in the future of the energy industry throughout the world. Nonetheless, to

reach higher performance and more desirable economics for PTSC-based systems, it is

f
oo
important to modify their base systems. The crucial novelties and originalities of the current

r
reach is as follows:


-p
This study makes modifications on a conventional system (CS), which has recently been
re
studied by numerous researchers, using PTSCs.
lP

 The configured modified CCHP system uses a multi-heat recovery technique to


na

ameliorate the capability of the CS.

 Comprehensive analysis of the role of circulated oil mass ratio of the solar subsystem on
ur

the performance criteria.


Jo

The CS consists of a PTSC field in combination with thermal storage tanks, an ORC for

electricity generation, a heating process heat exchanger (HPHX), and a SEACH for cooling

that is driven by the waste heat from the ORC. The MS embraces the same solar subsystem

coupled with two RORCs, three HPHXs, and a double-effect absorption chiller (DEACH). In

this regard, the utilization of the regeneration through employing an open feed heater section

in the RORC provides an opportunity to define an innovative heat recovery process for the

whole CCHP system at the suggested framework. Subsequently, the use of hot and cold

storage tanks, creating three different solar operation modes to cover the energy demand

6
throughout a day. These modes including low, high, and zero radiation status, which are

corresponded to early and late of the day, middle of the day, and night hours, respectively.

The main objective and methodology of this study is to comparatively investigate the

conventional and modified setups from exergy and exergoeconomic perspectives. In this

regard, a parametric study is carried out to determine the influence of variations in design

parameters such as inlet temperature of the ORC pump, inlet pressure of the ORC turbine,

and the oil mass ratio during charging time of the solar subsystem on the main exergy and

exergoeconomic variables of the system, including exergy destruction rate (EDR), exergy

f
oo
efficiency, CPUE, ICR, and exergoeconomic factor, for the mentioned solar radiation modes.

r
Then, a multi-objective optimization is performed using a genetic algorithm (GA) based on
-p
attaining the optimal values of the CCHP exergy efficiency, CPUE of the CCHP, and net
re
output electricity as objective functions. The current study is carried out using a code
lP

developed in Engineering Equation Solver (EES) software. Nevertheless, the description of


na

the suggested systems, modeling process, system validation, results and discussion, and

conclusion are presented in the following.


ur
Jo

2. Description of systems

The plot of the conventional and modified systems are illustrated in Figs. 1a and 1b,

respectively. The CS in Fig. 1a is a CCHP system encompassing a solar subsystem (a PTSC

field and storage tanks) and an ORC that generates electricity. The waste heat leaving the

ORC is used to provide heating in a HPHX and cooling via a SEACH. The MS (Fig. 1b) is a

combination of the mentioned solar subsystem with three HPHXs, two RORCs, and a

DEACH. The modifications enhance production rates and the system efficiencies when both

systems work under identical operating conditions. In general, a day can be divided into three

different radiation operation modes: a solar mode corresponding to the times 6:00-8:00 and

7
16:00-18:00, in which solar radiation levels are low; a solar and storage mode corresponding

to 8:00-16:00, when there is a high level of solar radiation; and a storage mode corresponding

to 18:00 and 6:00, during which there is no solar radiation and consequently no opportunity

to collect solar energy. According to the modified setup, during the solar mode, the absorbed

solar energy is totally used to operate the RORCs, whilst in the solar and storage mode, apart

from operating RORCs, a portion of the collected energy is stored. In fact, in the solar and

storage mode, Therminol-66 oil (the working fluid in the solar subsystem) at the outlet of the

PTSCs (state 1) is separated into two parts. The first part (state 2) flows into the evaporators

f
oo
of the ORCs, while the second portion (state 8) enters a heat exchanger acting as hot flow to

r
heat a cold oil flow from a cold storage tank (CST), leading to the storage of energy in a hot
-p
storage tank (HST). Eventually, in the storage mode, the system is driven utilizing the stored
re
hot oil (state 12). Note that n-octane, as the working fluid, is used to generate electricity by
lP

RORCs, and that the recovered energy of streams enhances the temperature of the flows
na

exiting the pumps and before entering the evaporators of the ORC (states 19 and 32). In

comparison with the oil flow, which leaves the CS’s ORC evaporator, the Therminol-66
ur

temperature leaving the MS first evaporator (state 4) experiences a lower reduction. Hence,
Jo

the high-temperature Therminol-66 is still usable to operate another RORC at a lower

pressure. Thus, the second RORC, encompassing a turbine working at a low pressure, is

employed in the MS so as to harness this surplus energy. Furthermore, the second RORC’s

waste heat is used to provide one more stage of heating. Lithium bromide-water (LiBr-H2O)

is used as the working fluid in the DEACH, whose function is described in Ref. [28].

3. Modeling

Table 1 lists the input data used in the current study. To simplify the simulations, the

following assumptions are employed:

8
 Both proposed systems work at steady-state.

 The friction effect and pressure drops in the heat exchangers and piping network are minor

and neglected.

 The kinetic and potential energies and exergies are neglected.

 The refrigerant flows exiting the evaporator and the condenser are saturated.

 Both weak and strong solutions leaving the absorber and desorbers are saturated at

equilibrium conditions.

3.1. Parabolic trough solar collectors

f
oo
Using the sun’s beam radiation rate (Gb), the absorbed solar radiation rate can be expressed as

r
follows [18]:

S = ρc  γ  τ  α r  k θ  Gb
-p (1)
re
Her, ρc denotes the mirror reflectance coefficient, γ is the intercept factor, τ is the glass cover
lP

transmittance, αr is the absorbance of the receiver, and kθ is the incidence angle modifier,
na

which can be written as [26]:


ur

k  1  2.2307 10-4  1.110-4 2  3.18596 10-6 3  4.85509 10-8 4 (2)
Jo

Here, θ is the incident angle, for which the following equation is valid [26]:

cos(θ)= cos2(θz )+cos2(δ)  sin2(ω) (3)

Here, zenith angle, declination angle, and the hour angle are denoted by θz, δ, and ω,

respectively.

The PTSCs net heat is obtained as follows [3]:

̇ ̇ [ ] (4)

̇ [ ( )] (5)

where ̇ denotes the oil mass flow rate in the receiver, Cp specific the heat at constant

pressure, Ar stands the PTSC receiver area, Aap shows the PTSC aperture area, T represents

9
the temperature, UL indicates the PTSC overall heat loss coefficient, and FR illustrates the

heat removal factor. The subscripts “in”, “out”, “r”, and “0 “represent inlet, outlet, receiver,

and dead state, respectively.

The aforementioned areas can be expressed as [3]:

Aap = W  Dc,out   L (6)

A r = π  Dr,out  L (7)

where W and L respectively denote the width and length of the PTSC, and Dc,out and Dr,out are

f
the outlet diameters of the cover and receiver, respectively.

oo
The heat removal factor can be expressed as [3]:

r
̇
[ (
̇
)]
-p (8)
re
Here F1 is the PTSC efficiency factor, defined as [3]:
lP

UO
F1 = (9)
UL
na

where UO denotes the overall heat transfer coefficient of the PTSC and UL is the heat loss
ur

coefficient of the PTSC. These terms can be defined as [3]:


Jo

-1
 Ar 1 
UL =  +  (10)
  hc,ca +hr,ca   Ac hr,cr 

-1
1 Dr,out D D  
UO =  + +  r,out  ln  r,out    (12)
 UL hc,r,in  Dr,in  2  k r  Dr,in   

The PTSC heat loss coefficient is constituted of three heat transfer coefficient types (see

Table 2). Also, as formulated in Table 2, Kr denotes the receiver thermal conductivity

coefficient, Dr,in is the inlet diameter of the receiver, and hc,r,in is the receiver internal

convection heat transfer coefficient.

The cover area (Ac) of the PTSC can be written as follows [3]:

10
A c = π  Dc,o  L (11)

The solar power, which is collected by the oil flowing in the receiver, can be expressed as

follows [21]:

̇ (13)

Finally, the PTSC total operating time during the solar mode (storing time) is [21]:

Δt h,solar = 24  Δt chst  Δt dhst (14)

where ∆tchst is the HST total charging time and ∆tdhst is the HST total discharge time.

f
The oil mass ratio (OMR) during the charging is defined as the ratio of the hot oil mass flow

oo
rate entering the solar heat exchanger to the total mass flow rate leaving the PTSCs. That is

r
[16], -p
re
̇ ⁄ ̇ (15)
{
̇ ⁄ ̇
lP

3.2. Thermal storage tanks


na

Table 3 shows the equations applied to simulate the thermal storage tanks during the

charging, storing, and discharging periods.


ur

3.3. Energy and exergy analyses


Jo

Mass and energy balances of a control volume at steady-state conditions can be written as

[29]:

∑ ̇ ∑ ̇ (16)

∑ ̇ ∑ ̇ ̇ ̇ (17)

where ̇ and ̇ denotes heat transfer and work rates, respectively .

The steady-state exergy rate balance can be written as follows for each component (18)

[29]:

11
̇ ∑( ) ̇ ̇ ∑ ̇ ∑ ̇

where ̇ denotes the EDR in the control volume, denotes the temperature at the

component boundaries, and ̇ and ̇ are the outlet and inlet exergy flow rates.

Additionally, subscripts “d”, “j,” and “cv” respectively represent destruction, boundary, and

control volume.

In this paper, only the physical exergy rate ( ̇ ) is effective and computed by [29]: (19)

f
̇ ∑ ̇ [ ]

oo
where h and s are the specific enthalpy and specific entropy, respectively .

r
Also, the overall EDR of the system is [29]: -p (20)
re
̇ ∑ ̇
lP

The energy and exergy balances of the system components are presented in Refs. [18,21,30],
na

while the energy and exergy equations and efficiencies are provided in Table 4.
ur

3.4. Exergoeconomic analysis


Jo

A general cost rate balance can be defined for each component as [29]: (21)

̇ ∑ ̇ ̇ ̇ ∑ ̇

̇ ̇ (22)

Here, denotes CPUE, ̇ stands the cost rate, and ̇ represents the ICR of component k.

Besides, Table 5 presents the equations connected with the balance of component cost rates

together with auxiliary relations.

The total ICR is computed as follows for each component [29]: (23)

12
where Zk, CRF, φ, and N denote the purchase cost of component k, capital recovery factor,

maintenance factor, annual operation time of the plant, respectively.

The chemical engineering plant cost index is used to convert the system (24)

component costs, available in a given year, to the present year (PY) [29]:

̇ ̇

(25)
̇ ∑ ̇

f
oo
The year 2019 has been selected as PY with the cost index of CI@2019 since this was the closet

time for which the information is available.

r
-p
The heat transfer surface area required for the applied heat exchangers for the (26)
re
satisfactory operation can be determined as follows [29]:
lP

̇
na

Here, ΔTLMTD is the logarithmic mean temperature difference, written as [29]: (27)
ur

( )
Jo

The capital recovery factor is [29]: (28)

Here, n and ir respectively denote life time of the component and annual interest rate.

Utilizing relations derived from the cost balances, a clearer understanding of the

exergoeconomic evaluation can be attained as follows [29]:

̇
(29)
̇

̇ ̇ (30)

13
̇ ∑ ̇ (31)

where cF,k is the fuel CPUEs for component k and ̇ is the exergy destruction cost rate

(EDCR).

The electrical and CPUE of the CCHP system are [29]:

̇ ̇ (32)
̇ ̇
̇

̇ ̇ ̇ ̇

f
̇ ̇ ̇

oo
̇ ̇ ̇ ̇ ̇ ̇
(33)
{ ̇ ̇ ̇ ̇ ̇

r
-p
re
Additionally, the electrical and CCHP cost rates are [29]:
lP

̇ ̇ (34)

[ ̇ ̇ ̇
na

]
̇ { (35)
[ ̇ ̇ ̇ ̇ ̇ ]
ur

Lastly, the exergoeconomic factor of the system can be expressed as [29]:


Jo

̇ (36)
̇ ̇

3.5. Optimization

GA is an optimization technique commencing with an initial estimation. The best results are

selected during each stage and transferred to the next stage based on their performance in

objective functions and capabilities in fulfilling criterions and addressing restrictions. This

method has been widely utilized by numerous researchers and is convenient to apply of in

EES. GA is used as an optimization method in the present study [30].

14
Table 1 contains the GA input data. The net output electricity, CCHP exergy efficiency as

well as the CPUE of the CCHP system are the objective functions. The decision variable

ranges follow:

RORC turbine 1 inlet pressure 1000  P20  2500 (37)

Pressure ratio between high-pressure and P20 (38)


2 3
P33
low pressure RORC turbines

P20 (39)
4 7
P21

f
oo
Pressure ratio of RORC turbine 1 P33 (40)
7  10
P34

r
RORC pumps 1 and 3 inlet temperature -p 360  T25 and T36  395 (41)
re
Pinch point temperature difference of the 30  Tpp-ORC eva  50 (42)
lP

RORC evaporator

PTCS outlet temperature 570  T1  650 (43)


na

Oil mass ratio 62  OMR  72 (44)


ur

For all operational modes of the modified setup, the objective functions are weighted so as to
Jo

carry out a multi-objective optimization. Accordingly, the multi-objective function (MOF) is

expressed as [17]:

̇ (45)

ω1 +ω2 +ω3 =1 (46)

0  ω1 ,ω2 , and ω3  1 (47)

where ω1, ω2, and ω3 denote weight coefficients of the CCHP exergy efficiency, net output

electricity, and CPUE of the CCHP system, respectively.

4. Validation

15
To validate the accuracy of the simulation results, the findings are compared with those of

Al-Sulaiman et al. [18] at the operating conditions introduced in Table 6. Additionally, to

verify the simulation of the PTSCs, a comparison is made between the results of the current

study and the numerical results presented by Al-Sulaiman et al. [18], together with the Sandia

National Laboratory’s experimental findings, presented by Dudley et al. [31]. The latter are

based on the heat loss against the average temperature above ambient around the fluid inside

the absorber, as shown in Fig. 2.

It is observed that there exists good agreement between the results of the current study and

f
oo
previously published works.

r
5. Results and discussion -p
re
5.1 Overall results
lP

Table 7 represents the results attained from comparing the systems at the baseline design
na

conditions (reported in Table 1) for all three radiation modes based on thermodynamic and

exergoeconomic variables. The difference between the two systems is also shown. Regarding
ur

this table, in all radiation modes, the net output electricity, and energy and exergy efficiencies
Jo

of the MS are significantly higher compared to the conventional one. Besides, the electricity

to heating ratio improves for the MS, indicating a reduction in the heating load due to the

increase in the generated electricity and decline in the waste heat, which is considered as a

positive point. Accordingly, decreasing the waste heat, the electricity to cooling decreases,

unveiling an enhancement in the produced cooling capacity. According to the calculations,

the exergy destruction rate of the MS is very close to that of the CS, and even it is lower in

the first radiation mode. However, it was expected to be considerably higher for all the

radiation modes since the MS enjoys higher production rates and more components.

Although the electrical CPUE enhances in solar and solar and storage modes, it reduces in the

16
storage mode. Additionally, the CPUE of the CCHP declines for all products, implying the

fact that the MS is cost-effective. Furthermore, although the ICR of the MS is greater, its

exergoeconomic factor grows, indicating the fact that compared with the CS, the modified

one is more efficient.

Sankey diagrams (see Fig. 3) are shown to provide a better perception of the modified setup’s

exergoeconomic performance at the baseline design conditions. Referring to these diagrams,

first, the contribution of all the subsystems is demonstrated based on ICR and EDCR, and

then the final value of the exergoeconomic factor is reported. From the information gathered

f
oo
in this figure, evident is the fact that the exergoeconomic factor of the MS in the solar mode,

r
which is equal to 69.7% is higher compared to the other two modes; 62.1% for the solar and
-p
storage mode, and 32.5% for the storage mode.
re
5.2. ORC pump inlet temperature
lP

Figs. 4a, 4b, and 4c correspondingly present the impact of the ORC pump inlet temperature
na

on the total EDR, electrical exergy efficiency, and CCHP exergy efficiency. The EDR rises

as a consequence of increasing the irreversibility due to the increased ORC pump inlet
ur

temperature. But, since EDR has an inverse relation with exergy efficiency, the exergy
Jo

efficiencies of the CCHP system and the electricity generation decline with increasing the

ORC pump inlet temperature. Another highlighted point is the closeness of the systems’

EDR. Thus, as the amounts of products generated by the modified setup are greater, its

produced exergy rate is higher as well, bringing about enhancements in its exergy

efficiencies. The electrical exergy efficiency of all the radiation modes improves, but in the

storage mode, this variable is of lower value than the MS at high temperatures. Referring to

Fig. 4a, the exergy destruction rate of the MS varies from 1685 kW to 1717 kW, 4768 kW to

4811 kW, and 892.5 kW to 956.4 kW for solar, solar and storage, and storage modes,

respectively. For the previously mentioned radiation operation modes, the electrical exergy

17
efficiency is 8.0%, 3.4%, and 5.1% at the temperature of 360 K, respectively, which are

correspondingly 0.8, 0.5, and 0.2 percent-points higher than those of the conventional one. At

a temperature of 390 K, this efficiency is equal to 6.3%, 2.7%, and 3.9% for the mentioned

radiation modes, respectively. Consequently, this efficiency witnesses an increase by 0.4 and

0.2 percent-points for the first two operation modes and a reduction by 0.2 percent-points for

the storage mode. Also, the MS’s CCHP exergy efficiency declines from 12.2%, 5.2%, and

7.6% to 10.7%, 4.5%, and 6.6% for the abovementioned radiation modes, respectively.

Likewise, for the solar, solar and storage, and storage operation modes, the CCHP exergy

f
oo
efficiency of the MS improves by 0.6, 0.6, and 0.3 percent-points at a temperature of 360 K,

r
and by 0.4, 0.2, and 0.1 percent-points at a temperature of 390 K compared to the

conventional CCHP system, respectively.


-p
re
Variations in the CPUE of the electricity and CCHP with inlet temperature of the ORC pump
lP

are depicted in Figs. 5a and 5b respectively. Taking into account these figures and solar
na

radiation modes, increasing the evaluated temperature here declines the electrical CPUE and

increases the CPUE of the CCHP. These reduction are owing to the decline in the electrical
ur

and CCHP exergy efficiencies, which previously discussed. Also, the modified setup’s
Jo

electrical CPUE, in solar mode and storage mode, is higher than that of the conventional one,

whilst for the MS, in solar and storage mode, this parameter is lower at low temperatures and

higher at high temperatures. Nonetheless, the CS’s CPUE is lower than the MS at all

radiation modes except for high temperatures in the storage mode. For the solar, solar and

storage and storage operation modes, the CPUE of the MS is found to be 46.86 $/GJ, 28.93

$/GJ, and 45.93 $/GJ, respectively, whilst at 390 K it is calculated as 52.19 $/GJ, 31.66 $/GJ,

and 51.07 $/GJ, respectively. At 360 K, the differences in the CPUE of the CCHP between

the systems for solar, solar and storage, and storage operation modes are 5.5%, 5.2%, and

2.8% respectively, while at 390 K, the differences are -4.0%, -0.6%, and 0.9%, respectively.

18
The impact of increasing the inlet temperature of the ORC pump on the systems’ ICRs and

exergoeconomic factors is illustrated in Figs. 6a and 6b, respectively. It is important

mentioning that as each component must supply the highest capacity in each radiation mode,

all the components employed in both systems have been selected with the highest achievable

capacity. Hence, the ICR associated with the selected mode is used for other modes.

According to these figures, both parameters descend as the temperature ascends. Considering

the performance deterioration as a consequence of enhancing the temperature and reducing

the amount of products, the system’s size declines, leading to a reduction in the ICR. But, due

f
oo
to the increase in EDR, EDCR also increases as the temperature rises, resulting in a drop in

r
the exergoeconomic factor. Due to the increase in the amount of products and the application
-p
of more components in the MS, the size or production rate of its components increases,
re
causing a higher ICR for the MS compared to the conventional one. Yet, since the
lP

exergoeconomic factor of the MS is greater, it can be claimed that the MS exhibits better
na

economic performance than the CS. As demonstrated in Fig. 6a, the MS’s ICR drops from

260.1 $/h, 276.5 $/h, and 36.24 $/h to 256.2 $/h, 272.6 $/h, and 32.37 $/h, for solar, solar and
ur

storage, and storage modes, correspondingly, which averagely are 2.0%, 2.5%, and 15.0%
Jo

higher than those of the CS, respectively. For the considered radiation modes, the

exergoeconomic factor of the MS declines from 69.92% to 69.19%, 62.21% to 61.61%, and

33.41% to 29.49% respectively. At an ORC inlet temperature of 360 K, in comparison with

the CS, the exergoeconomic factor of the MS is 0.47, 0.94, and 2.26 percent-points greater

for the aforementioned radiation operation modes, respectively. Subsequently, the MS’s

exergoeconomic factor of theses modes at 390 K, is individually 0.38, 0.40, and 2.38 percent-

points higher than the CS.

5.3 ORC turbine inlet pressure

19
Figs. 7a, 7b, and 7c exhibit variations in the EDR, electrical exergy efficiency, and CCHP

exergy efficiency with ORC turbine inlet pressure. With increasing the ORC turbine inlet

pressure, EDR reduces slightly. As a consequence of the reduction in the EDR, the

performance of the system improves. The variation range for the EDR of the CS at various

radiation modes are slightly greater than those of the modified one. Specifically speaking, as

the evaluated pressure rises from 1000 kPa to 2500 kPa, the exergy destruction rate of the MS

declines from 1713 kW to 1683 kW for the solar mode, 4816 kW to 4767 kW for the solar

and storage mode, and 916 kW to 899 kW for the storage mode. The modified setup’s

f
oo
electrical exergy rises from 6.4% to 8.1%, 2.8% to 3.4%, and 4.4% to 5.0% for the radiation

r
operation modes, respectively. Furthermore, at 1000 kPa and for the aforementioned radiation
-p
operation modes, the MS’s electrical exergy efficiency is 0.3, 0.1, and 0.1 percent-points
re
greater than the CS, whilst at 2500 kPa, this parameter is correspondingly 0.9, 0.6, and 0.3
lP

percent-points greater. In addition, the CCHP exergy efficiency of the MS is seen to improve
na

from 10.7% to 12.2% for the solar mode, from 4.6% to 5.2% for the solar and storage mode,

and from 7.1% to 7.5% for the storage mode, as ORC turbine inlet pressure rises. Compared
ur

to the CS, this efficiency is enhanced by 0.2-0.7, 0.2-0.7, and 0.2-0.3 percent-points for the
Jo

mentioned operation modes in the MS, respectively.

The effect of increasing the inlet pressure of the ORC turbine on the electrical CPUE and the

CPUE of the CCHP is illustrated in Figs. 8a and 8b, respectively. As seen in Fig. 8a, with an

increase in this pressure, the CS’s CPUE experiences a continuous decreasing trend in the

solar mode and decreasing-increasing trends in two other radiation modes. For the MS, this

parameter continually declines with increasing the ORC turbine inlet pressure in the solar and

solar and storage modes and exhibits a decreasing-increasing trend in the storage mode. In

Fig. 8b, the CPUE of CCHP of both systems is seen to experience decreasing-increasing

behaviors in all radiation modes. Referring to Fig. 8a, at the ORC turbine inlet pressure of

20
1000 kPa, the electrical CPUE of the MS for solar, solar and storage, and storage modes is

respectively equal to 43.39 $/GJ, 29.62 $/GJ, and 45.83 $/GJ that are the highest values in the

considered range for the evaluated pressure. Additionally, in the MS for the first two

radiation modes, the lowest values of the electrical CPUE occur at the pressure of 2500kPa,

which are respectively equal to 40.75 $/GJ and 28.61 $/GJ, while for the third mode, the

minimum CPUE, 42.79 $/GJ occurs at 2360 kPa. But, for the CS, the lowest values of the

CPUE are 39.68 $/GJ (at 2500 kPa) for the solar mode, 29.35 $/GJ (at 1270 kPa) for the solar

and storage mode, and 43.06 $/GJ (at 1550 kPa) for the storage mode. In Fig. 8b, it is

f
oo
observed that the minimum values related to the CPUE of the CCHP for the MS at the

r
abovementioned radiation modes are respectively 47.01 $/GJ at 2360 kPa, 29.27 $/GJ at 2220
-p
kPa, and 46.65 $/GJ at 2100 kPa; these values are correspondingly 5.6%, 3.5%, and 2.0%
re
lower than those of the CS.
lP

Figs. 9a and 9b demonstrate the behavior of the ICR and exergoeconomic factor versus the
na

ORC turbine inlet temperature, respectively. Due to the improvements in the performance

and the amount of products as a result of increasing the ORC turbine inlet pressure, the size
ur

or production rate of the system’s components is enhanced, resulting in an increase in the


Jo

ICR. The exergoeconomic factor improves as well since the irreversibility declines and,

subsequently, the EDCR declines. At the ORC turbine inlet pressure of 1000 kPa, the ICR of

the MS is respectively 255.6 $/h, 271.1 $/h, and 31.6 $/h for the solar, solar and storage,

and storage modes that are correspondingly 1.5%, 1.6%, and 14.0% higher than those

of the CS. Besides, at 2500 kPa, for the MS at the mentioned radiation modes, this

variable is individually 263.0 $/h, 280.0 $/h, and 39.2 $/h, which are respectively

3.0%, 3.5%, and 26.0% greater compared with those of the conventional one.

Moreover, the MS’s exergoeconomic factor rises from 68.89% to 70.24% for the solar

mode, 61.29% to 62.58% for the solar and storage mode, and 30.09% to 34.80% for the

21
storage mode. These values are 0.06-0.78, 0.25-1.26, and 1.74-3.85 percent-points higher

than those of the CS, respectively.

5.4. Oil mass ratio

Figs. 10a, 10b, and 10c plot the influence of enhancing the OMR on the total EDR, electrical

efficiency and CCHP exergy efficiency. As the OMR increases, the amount of the energy

saved by the HTS rises, which improves the production rate of the products for the storage

mode. Hence, as the production rate of the products decreases, system’s irreversibility

declines in the solar and storage mode and increase in the storage mode, resulting in a

f
oo
reduction in the EDR in the solar and storage mode, and an ascendance in the storage mode

r
as shown in Fig. 10a. Also, the exergy efficiency experiences a decreasing trend in the solar
-p
and storage mode because, under the influence of the mentioned variation, the output exergy
re
rate for this mode grows less than the exergy rate entering the system during charging.
lP

However, in the storage mode, this variable increases due to the dominance of products’
na

exergy rate over the exergy rate entering the system. In accordance with Fig. 10a, as the

OMR enhances from 62% to 72%, the MS’s total exergy destruction rate decreases from
ur

4825.0 kW to 4660.0 kW in the solar and storage mode, but it increases from 855.6 kW to
Jo

1020 kW in the storage mode, that is slightly lower than that of the CS. The modified setup’s

electrical exergy efficiency drops by 1% in the solar and storage mode from 3.6% to 2.6%,

while in the storage mode, this variable improves from 4.8% to 5.0%. Meanwhile, compared

with the CS, the electrical exergy efficiency improves by 0.5 and 0.2 percent-points for the

mentioned operation modes. Also, the MS’s CCHP exergy efficiency reduces by 1.5% in the

solar and storage operation mode, while in storage operation mode, this variable rises by

0.4%. Also, the MS’s CCHP exergy efficiency (Fig. 10c) reduces from 5.5% to 4% in the

solar and storage mode, whereas in the storage mode, this variable enhances from 7.3% to

7.7%. In comparison with the CS, the MS’s CCHP exergy efficiency is correspondingly 0.5

22
and 0.3 percent-points greater for the solar and storage and storage operation modes,

respectively.

Variations in the electrical CPUE and CCHP’s CPUE with the OMR during charging are

plotted in Figs. 11a and 11b, respectively. According to these figures, the CPUE experiences

an improvement in the solar and storage mode, but it declines in the storage mode. Looking at

Fig. 11a, it is obvious that when the OMR varies from 62% to 72%, the electrical CPUE rises

from 27.93 $/GJ to 31.70 $/GJ in the solar and storage mode, but it reduces from 46.10 $/GJ

to 43.01 $/GJ. At the OMR of 62%, the difference in the electrical CPUE between the

f
oo
modified and conventional cycles is -2.2% and 3.4%, while at the OMR of 72%, the

r
difference is -7.0% and 5.4%, for the mentioned radiation modes, respectively. Besides, the
-p
MS’s CPUE of CCHP increases from 28.95 $/GJ to 30.74 $/GJ in the solar and storage mode,
re
but in the storage mode, this variable drops from 47.26 $/GJ to 46.19 $/GJ. Compared with
lP

the CS, at OMR of 62%, the MS’s CPUE is 4.1% and 2.3% greater, and at the OMR of 72%,
na

it is 6.5% and 1.3% higher for the abovementioned radiation modes, individually.

Figs. 12a and 12b exhibit variations in the ICR and exergoeconomic factor with increasing
ur

the OMR during charging. In the solar and storage mode, the size or production rate of the
Jo

system’s components reduces as the OMR increases because of the reductions in products,

leading to a decrease in the ICR. Conversely, in the storage mode, the ICR decreases as the

OMR increases. With increasing the OMR from 62% to 72%, the ICR of the MS in the solar

and storage mode, which is 2.5% higher than that of the CS, diminishes from 280.0 $/h to

275.1 $/h. Likewise, this variable in the storage mode that is 18.1% bigger compared to the

conventional one, grows from 35.4$/h to 35.7 $/h. Moreover, the exergoeconomic factor

declines in the solar and storage mode, but rises in the storage mode. As can be seen, the

exergoeconomic factor of the MS is 63.13% and 30.62% for the solar and storage mode and

storage mode, respectively. Also, at the OMR of 72%, this variable is 61.74% and 33.43%

23
for the mentioned modes, respectively. It is worth noting that, compared with the MS, the

CS’s exergoeconomic factor of the mentioned radiation modes is 1.19 and 2.52 percent-

points lower at the OMR of 62%, respectively. Furthermore, these variables at the OMR of

72% are correspondingly 0.77 and 2.12 percent-points lower.

5.5. Optimization Results

In order to optimize the MS via the genetic algorithm, the optimum design mode is

determined based on the decision variables in three optimization cases by giving weight

coefficient to objective functions as reported in Table 8. Also, the optimum values of other

f
oo
crucial variables in investigated cases for all radiation modes are thoroughly provided in the

r
same table.
-p
The weight coefficients applied to the CCHP exergy efficiency, the net output electricity, and
re
the CPUE of the CCHP are 0.5, 0, and 0.5 in the first case, respectively. In this case, the
lP

optimum values of the CCHP exergy efficiency, the net output electricity, and the CPUE of
na

the CCHP are correspondingly found to be 12.58%, 941.0 kW, and 46.23 $/GJ for the solar

mode, 6.04%, 759.6 kW, and 28.14 $/GJ for the solar and storage mode, and 8.29%, 760.0
ur

kW, and 43.81 $/GJ for storage mode. In this way, the CCHP exergy efficiency of the
Jo

modified CCHP system improves by 0.48, 0.94, and 0.69 percent-points, its net output

electricity increases by 6.9%, 18.6%, and 23.8%, and also its CPUE declines by 8.1%, 8.3%,

and 8.2% for the low, high, and zero radiation modes, respectively.

The weight coefficients considered for the abovementioned objective functions, in the second

case, are 0.0, 0.5, and 0.5 respectively. Hence, the calculated optimum CCHP exergy

efficiencies for the solar, solar and storage and storage modes are 12.52%, 6.00%, and 8.23%

respectively. These values are 0.42, 0.9, and 0.63 percent-points higher than those calculated

for the baseline design conditions, respectively. Accordingly, the maximum net output

electricity of the modified application is found to be 947.5 kW, 768.5 kW, and 764.3 kW for

24
these modes, respectively, which are 7.6%, 20.0%, and 24.5% grater then the baseline design

conditions. Finally, the lowest CPUE of the CCHP are achieved as 46.71 $/GJ, 28.23 $/GJ,

and 44.01 $/GJ for the aforementioned radiation modes, respectively. This optimization case

reduces these values up to 7.1%, 8.0%, and 7.7%, correspondingly.

As for the third case, the weight coefficient of all objective functions is set to be 1/3. The

optimum values of the CCHP exergy efficiency, net output electricity, and CPUE of the

CCHP individually are found to be 12.55%, 944.3 kW, and 46.33 $/GJ for the first mode,

6.01%, 766.3 kW, and 28.21 $/GJ for the second mode, and 8.28%, 762.9 kW, and 44.02

f
oo
$/GJ for the third mode. The CCHP exergy efficiency, net output electricity, and CPUE of the

r
CCHP are improves as 0.45 percent-points, 19.7%, and 8.1% at the first mode, 0.91 percent-
-p
points, 19.6%, and 7.9% at the second mode, and 0.68 percent-points, 24.3%, and 7.7% at the
re
third mode.
lP

6. Conclusions
na

Exergetic and exergoeconomic aspects of a modified setup of a trigeneration system driven


ur

by solar energy were evaluated and compared with the corresponding conventional system.
Jo

The conventional system is composed of parabolic trough solar collectors in integration with

thermal storage tanks as the solar subsystem, an organic Rankine cycle, a single effect

absorption chiller, and a heating production heat exchanger. Modifications were made on the

system using three heating production heat exchangers, two regenerative organic Rankine

cycles, and a double-effect absorption chiller. The simulation results were compared for the

systems at three operation modes encompassing low radiation (solar mode), high radiation

(solar and storage mode), and zero radiation (storage mode). A parametric study and a multi-

objective optimization were conducted. The main conclusions and outcomes of this study are

as follow:

25
 For solar, solar and storage, and storage operation modes, the overall trigeneration energy

and exergy efficiencies rise by 2.6 and 0.8 percent-points, 4.0 and 0.6 percent-points, and

1.3 and 0.4 percent-points.

 At the same operating base conditions, compared to the conventional system, the modified

setup’s cost per unit exergy decreases by 5.4%, 4.6%, and 2.1% for the above-mentioned

solar radiation modes, respectively.

 At the same operating base conditions, the modified setup’s exergoeconomic factor

increases by 0.4, 0.9, and 2.1 percent-points for the aforementioned solar radiation modes,

f
oo
correspondingly.

r
 An increase in the organic Rankine cycle’s pump inlet temperature increases the exergy
-p
destruction rate and cost per unit exergy of the overall system and reduces the electrical
re
and overall exergy efficiencies, electrical cost per unit exergy, investment cost rate, and
lP

exergoeconomic factor.
na

 As the organic Rankine cycle’s turbine inlet pressure increases, the exergy destruction rate

drops, while the electrical and overall exergy efficiencies, investment cost rate, and
ur

exergoeconomic factor rise. Likewise, the modified setup’s electrical cost per unit exergy
Jo

experiences a decreasing-increasing trend in the storage mode and declines steadily in

other two modes.

 By increasing the oil mass ratio in the solar and storage mode, the exergy destruction rate,

the electrical and overall exergy efficiencies, the investment cost rate and the

exergoeconomic factor decrease, while the electrical cost per unit exergy and the overall

cost per unit exergy increase.

26
f
r oo
Nomenclature
A
c
Area (m2)
Cost per unit exergy ($/GJ)
-p Greek symbols
re
̇ Cost rate ($/h) αr Absorbance of receiver
Specific heat at constant pressure
Cp γ Intercept factor (o)
(kJ/kgK)
lP

CRF Capital recovery factor δ Declination angle


D Diameter (m) ε Emittance factor; Exergy efficiency (%)
̇ Exergy rate (kW) η Energy efficiency (%)
na

̇ Exergy destruction rate (kW) θ Incident angle (o)


EV Expansion valve θz Zenith angle (o)
f Exergoeconomic factor ρ Reflectance factor, Density (kg/m3)
ur

F1 Collector efficiency factor σ Stefan-Boltzmann constant (kW/m2K4)


FR Heat removal factor ω Hour angle (o)
Gb Solar beam radiation per unit area (W/m2) Subscripts
Jo

Specific enthalpy (kJ/kg); Convection


h 0 Reference condition
heat transfer coefficient (W/m2K)
ir Annual interest rate (%) a Air
kθ Incidence angle modifier ap Aperture
k Thermal conductivity (W/mK) av Average
L Length (m)
̇ Mass flow rate (kg/s) b Beam
M Mass (kg) c Cover; Cooling
NPTSC Number of PTSCs CCHP Combined cooling, heating, and power
Nu Nusselt number chst Charging time
P Pressure (kPa) d Diffuse
Pr Prandtl number dhst Discharging time
Q Heat (kJ) e Exit
̇ Heat transfer rate (kW) el Electrical
Re Reynolds number h Heating
rel/c Electricity to cooling ratio in Inlet
rel/h Electricity to heating ratio L Lost
s Specific entropy (kJ/kgK) net Net
S Absorbed radiation per unit area (W/m2) out Outlet
t Time (s) ph Physical
T Temperature (K) r Receiver; Radiation
U Heat loss coefficient (W/m2K) sol Solar mode
V Valve solst Solar and storage mode

27
W Width (m) st Storage mode
̇ Mechanical power and electricity rate
tlost Total heat lost
(kW)
Z Purchase cost ($) tot Total
̇ Investment cost rate ($/h) u Useful

f
r oo
References
[1]
-p
M.A. Haghghi, S.G. Holagh, S.M. Pesteei, A. Chitsaz, F. Talati, On the performance,
re
economic, and environmental assessment of integrating a solar-based heating system with
conventional heating equipment$\mathsemicolon$ a case study, Therm. Sci. Eng. Prog. 13
(2019) 100392. https://doi.org/10.1016/j.tsep.2019.100392.
lP

[2] A. Khaligh, O.C. Onar, Energy Harvesting: Solar, Wind, and Ocean Energy Conversion
Systems, n.d. https://www.routledge.com/Energy-Harvesting-Solar-Wind-and-Ocean-Energy-
na

Conversion-Systems/Khaligh-Onar/p/book/9781439815083# (accessed April 15, 2021).


[3] S.A. Kalogirou, Solar Energy Engineering - Processes and Systems , 2nd Edition, Academic
ur

Press, 2013. https://www.elsevier.com/books/solar-energy-engineering/kalogirou/978-0-12-


397270-5 (accessed April 15, 2021).
Jo

[4] A.A. AlZahrani, I. Dincer, Energy and exergy analyses of a parabolic trough solar power plant
using carbon dioxide power cycle, Energy Convers. Manag. 158 (2018) 476–488.
https://doi.org/10.1016/j.enconman.2017.12.071.
[5] E.J.C. Cavalcanti, H.P. Motta, Exergoeconomic analysis of a solar-powered/fuel assisted
Rankine cycle for power generation, Energy. 88 (2015) 555–562.
https://doi.org/10.1016/j.energy.2015.05.081.
[6] H. Habibi, A. Chitsaz, K. Javaherdeh, M. Zoghi, M. Ayazpour, Thermo-economic analysis and
optimization of a solar-driven ammonia-water regenerative Rankine cycle and LNG cold
energy, Energy. 149 (2018) 147–160. https://doi.org/10.1016/j.energy.2018.01.157.
[7] J. Oyekale, M. Petrollese, F. Heberle, D. Brüggemann, G. Cau, Exergetic and integrated
exergoeconomic assessments of a hybrid solar-biomass organic Rankine cycle cogeneration
plant, Energy Convers. Manag. 215 (2020) 112905.
https://doi.org/10.1016/j.enconman.2020.112905.
[8] M. Ashouri, M.H. Ahmadi, S.M. Pourkiaei, F.R. Astaraei, R. Ghasempour, T. Ming, J.H.
Hemati, Exergy and exergo-economic analysis and optimization of a solar double pressure
organic Rankine cycle, Therm. Sci. Eng. Prog. 6 (2018) 72–86.
https://doi.org/10.1016/j.tsep.2017.10.002.
[9] F. Calise, D. Capuano, L. Vanoli, Dynamic Simulation and Exergo-Economic Optimization of

28
a Hybrid Solar-Geothermal Cogeneration Plant, Energies. 8 (2015) 2606–2646.
https://doi.org/10.3390/en8042606.
[10] M. Mehrpooya, M. Ashouri, A. Mohammadi, Thermoeconomic analysis and optimization of a
regenerative two-stage organic Rankine cycle coupled with liquefied natural gas and solar
energy, Energy. 126 (2017) 899–914. https://doi.org/10.1016/j.energy.2017.03.064.
[11] V. Zare, A. Moalemian, Parabolic trough solar collectors integrated with a Kalina cycle for
high temperature applications: Energy, exergy and economic analyses, Energy Convers.
Manag. 151 (2017) 681–692. https://doi.org/10.1016/j.enconman.2017.09.028.
[12] A. Mohammadi, M. Mehrpooya, Thermodynamic and economic analyses of hydrogen
production system using high temperature solid oxide electrolyzer integrated with parabolic
trough collector, J. Clean. Prod. 212 (2019) 713–726.
https://doi.org/10.1016/j.jclepro.2018.11.261.
[13] Y.N. Dabwan, P. Gang, J. Li, G. Gao, J. Feng, Development and assessment of integrating
parabolic trough collectors with gas turbine trigeneration system for producing electricity,

f
chilled water, and freshwater, Energy. 162 (2018) 364–379.

oo
https://doi.org/10.1016/j.energy.2018.07.211.
[14] F. Calise, M.D. d’Accadia, A. Macaluso, A. Piacentino, L. Vanoli, Exergetic and

r
exergoeconomic analysis of a novel hybrid solar-geothermal polygeneration system producing
energy and water, Energy -pConvers.
https://doi.org/10.1016/j.enconman.2016.02.029.
Manag. 115 (2016) 200–220.
re
[15] S.M. Alirahmi, S.R. Dabbagh, P. Ahmadi, S. Wongwises, Multi-objective design optimization
of a multi-generation energy system based on geothermal and solar energy, Energy Convers.
lP

Manag. 205 (2020) 112426. https://doi.org/10.1016/j.enconman.2019.112426.


[16] M. Delpisheh, M.A. Haghghi, H. Athari, M. Mehrpooya, Desalinated water and hydrogen
na

generation from seawater via a desalination unit and a low temperature electrolysis using a
novel solar-based setup, Int. J. Hydrogen Energy. 46 (2021) 7211–7229.
https://doi.org/10.1016/j.ijhydene.2020.11.215.
ur

[17] M. Delpisheh, M.A. Haghghi, M. Mehrpooya, A. Chitsaz, H. Athari, Design and financial
parametric assessment and optimization of a novel solar-driven freshwater and hydrogen
Jo

cogeneration system with thermal energy storage, Sustain. Energy Technol. Assessments. 45
(2021) 101096. https://doi.org/10.1016/j.seta.2021.101096.
[18] F.A. Al-Sulaiman, F. Hamdullahpur, I. Dincer, Performance assessment of a novel system
using parabolic trough solar collectors for combined cooling, heating, and power production,
Renew. Energy. 48 (2012) 161–172. https://doi.org/10.1016/j.renene.2012.04.034.
[19] F.A. Al-Sulaiman, I. Dincer, F. Hamdullahpur, Exergy modeling of a new solar driven
trigeneration system, Sol. Energy. 85 (2011) 2228–2243.
https://doi.org/10.1016/j.solener.2011.06.009.
[20] F.A. Al-Sulaiman, I. Dincer, F. Hamdullahpur, Thermoeconomic optimization of three
trigeneration systems using organic Rankine cycles: Part II - Applications, Energy Convers.
Manag. 69 (2013) 209–216. https://doi.org/10.1016/j.enconman.2012.12.032.
[21] M.A. Haghghi, S.M. Pesteei, A. Chitsaz, J. Hosseinpour, Thermodynamic investigation of a
new combined cooling, heating, and power (CCHP) system driven by parabolic trough solar
collectors (PTSCs): A case study, Appl. Therm. Eng. 163 (2019) 114329.
https://doi.org/10.1016/j.applthermaleng.2019.114329.
[22] M.A. Haghghi, Z. Mohammadi, S.M. Pesteei, A. Chitsaz, K. Parham, Exergoeconomic
evaluation of a system driven by parabolic trough solar collectors for combined cooling,
heating, and power generation; a case study, Energy. 192 (2020) 116594.

29
https://doi.org/10.1016/j.energy.2019.116594.
[23] A. Baghernejad, M. Yaghoubi, K. Jafarpur, Exergoeconomic optimization and environmental
analysis of a novel solar-trigeneration system for heating, cooling and power production
purpose, Sol. Energy. 134 (2016) 165–179. https://doi.org/10.1016/j.solener.2016.04.046.
[24] F. Khalid, I. Dincer, M.A. Rosen, Thermoeconomic analysis of a solar-biomass integrated
multigeneration system for a community, Appl. Therm. Eng. 120 (2017) 645–653.
https://doi.org/10.1016/j.applthermaleng.2017.03.040.
[25] J. Wang, Z. Lu, M. Li, N. Lior, W. Li, Energy, exergy, exergoeconomic and environmental
(4E) analysis of a distributed generation solar-assisted CCHP (combined cooling, heating and
power) gas turbine system, Energy. 175 (2019) 1246–1258.
https://doi.org/10.1016/j.energy.2019.03.147.
[26] E. Bellos, C. Tzivanidis, K. Torosian, Energetic, exergetic and financial evaluation of a solar
driven trigeneration system, Therm. Sci. Eng. Prog. 7 (2018) 99–106.
https://doi.org/10.1016/j.tsep.2018.06.001.

f
oo
[27] Y.N. Dabwan, G. Pei, A novel integrated solar gas turbine trigeneration system for production
of power, heat and cooling: Thermodynamic-economic-environmental analysis, Renew.
Energy. 152 (2020) 925–941. https://doi.org/10.1016/j.renene.2020.01.088.

r
[28] -p
R. Gomri, R. Hakimi, Second law analysis of double effect vapour absorption cooler system,
Energy Convers. Manag.
https://doi.org/10.1016/j.enconman.2007.09.033.
49 (2008) 3343–3348.
re
[29] A. Bejan, G. Tsatsaronis, M.J. Moran, Thermal design and optimization, John Wiley & Sons,
lP

Ltd, 1995.
[30] Y. Cao, A.M. Mohamed, M. Dahari, M. Delpisheh, M.A. Haghghi, Performance enhancement
and multi-objective optimization of a solar-driven setup with storage process using an
na

innovative modification, J. Energy Storage. 32 (2020).


https://doi.org/10.1016/j.est.2020.101956.
ur

[31] V. Dudley, G. Kolb, A. Mahoney, T. Mancini, C. Matthews, M. Sloan, D. Kearney, Test


results: SEGS LS-2 solar collector, Office of Scientific and Technical Information (OSTI),
1994. https://doi.org/10.2172/70756.
Jo

[32] Y. Cao, M.A. Haghghi, M. Shamsaiee, H. Athari, M. Ghaemi, M.A. Rosen, Evaluation and
optimization of a novel geothermal-driven hydrogen production system using an electrolyser
fed by a two-stage organic Rankine cycle with different working fluids, J. Energy Storage. 32
(2020) 101766. https://doi.org/10.1016/j.est.2020.101766.

30
Jo
ur
na

31
lP
re
-p
r oo
f
9 PTSC 1

8 7

V-1
V-4
17 SolHX 10
6

EV-2 StP-2
SolP
HST
16 11
2

StP-1 EV-1
15 12
5
CS
33 13
14
4 3

f
V-3 ORC eva V-2

oo
18 19
ORCP

ORCT

r
22 -p 20
21
HPHX
re
Des 23 24
lP

31 36
Cond
na

25 35
REV

32 30
26
29 SP 28
ur

27
SHX Abs Eva
Jo

33 SEV 34
40 39 38 37
PTSC Parabolic trough solar collectors V Valve Abs Absorber
SolHX Solar heat exchanger HPHX Heatin production heat exchanger Eva Evaporator
HST Hot storage tank ORC eva ORC evaporator Cond Condenser
CST Cold storage tank ORCT ORC turbine SP Solution pump
StP Storage pump ORCP ORC pump REV Refrigeration expansion valve
SolP Solar pump Des Desorber SEV Solution expansion vavle
EV Expansion valve SHX Solution heat exchanger

Fig. 1a The plot of the conventional solar-based trigeneration system driven by parabolic
trough solar collectors.

32
10 PTSC 1

9 8

V-1
V-

18 SolHX 11
7
SolP

EV-2 StP-2 HST


17 12

StP-1 EV-1
2
16 13
6
CST
15 14
5 4 3

f
V-3 ORC eva-2 ORC eva-1 V-2

oo
32 33 20
19

r
ORCT-2 -p ORCT-1
ORCP-4

ORCP-

re
38 35 34 27 21
lP

OFH-2 OFH-1 22
28 29
37 26
ORCP-

36
na

HPHX-3 HPHX-2
ORCP-3 23
39 40
ur

25
HPHX-1 HPD
24
Jo

59 58 30 31
54
Cond
53
REV-1 51 48 47
41 52
50 SEV-1 49
REV-2

LPD SHX-1

42 55
46
57 SEV-2 56
Abs Eva SHX-2
43
44 45
63 62 61 60 SP

PTSC Parabolic trough solar collectors V Valve LPD Low pressure desorber
SolHX Solar heat exchanger HPHX Heatin production heat exchanger SHX Solution heat exchanger
HST Hot storage tank ORC eva ORC evaporator Abs Absorber
CST Cold storage tank ORCT ORC turbine Eva Evaporator
StP Storage pump ORCP ORC pump Cond Condenser
SolP Solar pump OFH Open feed heater SP Solution pump

33
EV Expansion valve HPD High pressure desorber REV Refrigeration expansion valve
SEV Solution expansion vavle

Fig. 1b The plot of the modified solar-based trigeneration system driven by parabolic
trough solar collectors.

80
This paper
70 Al-Sulaiman et al. [18]
Sandia test results [31]
Heat loss per unit area [W/m2]

60

f
50

oo
40

r
30 -p
re
20
lP

10

0
na

90 120 150 180 210 240 270 300


Average temperature above ambient [°C]
Fig. 2 Comparing this paper results with Refs. [18,31] for collectors based on the effect of
ur

variation in average temperature above ambient on the heat loss rate per unit area.
Jo

34
f
r oo
-p
re
lP
na
ur
Jo

a b c
Fig. 3 Sankey diagrams indicating the exergoeconomic parameters of the modified setup: (a)
solar, (b) solar and storage, and (c) storage modes.

35
Exergy destruction rate (sol and st) [kW] 1800 4850 8.5

Exergy destruction rate (solst) [kW]

Exergy efficiency (electricity) [%]


7.5
4825
1600

6.5
4800
1400 eel-sol-a eel-sol-b
5.5 eel-solst-a eel-solst-b
Exd-sol-b eel-st-a eel-st-b
4775
Exd-solst-b
Exd-st-b 4.5
1200
Exd-sol-a
Exd-solst-a 4750
Exd-st-a 3.5

1000
4725
2.5

f
800 4700 1.5

oo
360 365 370 375 380 385 390 360 365 370 375 380 385 390
ORC pump inlet temperature [K] ORC pump inlet temperature [K]
a b

r
12.5

11.5
-p
re
Exergy efficiency (CCHP) [%]

10.5
lP

9.5 eCCHP-sol-a
eCCHP-solst-a
eCCHP-st-a
8.5
na

7.5

6.5 eCCHP-sol-b
ur

eCCHP-solst-b
5.5 eCCHP-st-b
Jo

4.5

3.5
360 365 370 375 380 385 390
ORC pump inlet temperature [K]
c
Fig. 4 Variation in (a) total EDR, (b) electrical exergy efficiency, and (c) CCHP exergy
efficiency versus ORC pump inlet temperature.

36
50

Cost per unit exergy (electricity) [$/GJ]


cel-sol-a
cel-solst-a
cel-st-a
45

40

35 cel-sol-b
cel-solst-b
cel-st-b

30

f
oo
25
360 365 370 375 380 385 390

r
ORC pump inlet temperature [K]

55
-p a
re
Cost per unit exergy (CCHP) [$/GJ]

50
lP

45
na

cCCHP-sol-b
cCCHP-solst-b
40 cCCHP-st-b
ur

cCCHP-sol-a
cCCHP-solst-a
Jo

35 cCCHP-st-a

30

25
360 365 370 375 380 385 390
ORC pump inlet temperature [K]
b
Fig. 5 Variation in CPUE for (a) generated electricity and (b) CCHP versus ORC pump
inlet temperature.

37
280 38

Investment cost rate (sol and solst) [$/h]


275

Investment cost rate (st) [$/h]


36

270
34
Zsol-a
265 Zsolst-a
Zst-a
Zsol-b
32
Zsolst-b
260 Zst-b

30
255

f
oo
250 28
360 365 370 375 380 385 390

r
ORC pump inlet temperature [K]

75
-pa
re
70
lP

65
Exergoeconomic factor [%]

60
na

55
fsol-a
50 fsolst-a
ur

fst-a
fsol-b
45
fsolst-b
Jo

fst-b
40

35

30

25
360 365 370 375 380 385 390
ORC pump inlet temperature [K]
b
Fig. 6 Variation in (a) ICR and (b) exergoeconomic factor versus ORC pump inlet
temperature.

38
Exergy destruction rate (sol and st) [kW] 1800 4850 9

1700 eel-sol-a

Exergy destruction rate (solst) [kW]


eel-solst-a

Exergy efficiency (electricity) [%]


8
eel-st-a
1600 4830
Exd-sol-a
Exd-solst-a 7
1500
Exd-st-a
1400 4810
6
eel-sol-b
1300 eel-solst-b
Exd-sol-b
eel-st-b
Exd-solst-b 5
1200 Exd-st-b 4790

1100 4

1000 4770
3
900

f
800 4750 2

oo
1000 1250 1500 1750 2000 2250 2500 1000 1250 1500 1750 2000 2250 2500
ORC turbine inlet pressure [kPa] ORC turbine inlet pressure [kPa]
a b

r
13

12
-p
re
Exergy efficiency (CCHP) [%]

11
eCCHP-sol-a
lP

10 eCCHP-solst-a
eCCHP-st-a
9 eCCHP-sol-b
eCCHP-solst-b
na

8 eCCHP-st-b

7
ur

6
Jo

4
1000 1250 1500 1750 2000 2250 2500
ORC turbine inlet pressure [kPa]
c
Fig. 7 Variation in (a) total EDR, (b) electrical exergy efficiency, and (c) CCHP exergy
efficiency versus ORC turbine inlet pressure.

39
50

Cost per unit exergy (electricity) [$/GJ]


45

40

cel-sol-b
cel-solst-b
35 cel-st-b
cel-sol-a
cel-solst-a
cel-st-a
30

f
oo
25
1000 1250 1500 1750 2000 2250 2500

r
ORC turbine inlet pressure [kPa]

55
-pa
re
Cost per unit exergy (CCHP) [$/GJ]

50
lP

45
na

cCCHP-sol-b
cCCHP-solst-b
40 cCCHP-st-b
ur

cCCHP-sol-a
cCCHP-solst-a
Jo

35 cCCHP-st-a

30

25
1000 1250 1500 1750 2000 2250 2500
ORC turbine inlet pressure [kPa]
b
Fig. 8 Variation in CPUE for (a) generated electricity and (b) CCHP versus ORC turbine
inlet pressure.

40
280 40

Investment cost rate (sol and solst) [$/h]


Zsol-a
Zsolst-a
Zst-a 38
275

Investment cost rate (st) [$/h]


36
270

34
Zsol-b
265
Zsolst-b
Zst-b 32

260
30

255
28

f
oo
250 26
1000 1250 1500 1750 2000 2250 2500

r
ORC turbine inlet pressure [kPa]

75
-p
a
re
70
lP

65
Exergoeconomic factor [%]

60
na

55
fsol-a
50 fsolst-a
ur

fst-a
fsol-b
45
fsolst-b
Jo

fst-b
40

35

30

25
1000 1250 1500 1750 2000 2250 2500
ORC turbine inlet pressure [kPa]
b
Fig. 9 Variation in (a) ICR and (b) exergoeconomic factor versus ORC turbine inlet
pressure.

41
1050 4900 5.5

Exd-solst-a

Exergy destruction rate (solst) [kW]


1025 5

Exergy efficiency (electricity) [%]


Exd-st-a
Exergy destruction rate (st) [kW]

Exd-solst-b 4850
1000 Exd-st-b 4.5
eel-solst-a
eel-st-a
975 4
4800 eel-solst-b
eel-st-b
950 3.5

4750
925 3

900 2.5
4700
875 2

f
850 4650 1.5

oo
62 63 64 65 66 67 68 69 70 71 72 62 63 64 65 66 67 68 69 70 71 72
Mass ratio of oil at charging time [%] Mass ratio of oil at charging time [%]
a b

r
8

7.5
-p
re
Exergy efficiency (CCHP) [%]

6.5 eCCHP-solst-a
lP

eCCHP-st-a
6 eCCHP-solst-b
eCCHP-st-b
na

5.5

5
ur

4.5

4
Jo

3.5

3
62 63 64 65 66 67 68 69 70 71 72
Mass ratio of oil at charging time [%]
c
Fig. 10 Variation in (a) total EDR, (b) electrical exergy efficiency, and (c) CCHP exergy
efficiency versus OMR.

42
50

Cost per unit exergy (electricity) [$/GJ]


45

40 cel-solst-a
cel-st-a
cel-solst-b
cel-st-b
35

30

f
oo
25
62 63 64 65 66 67 68 69 70 71 72

r
Mass ratio of oil at charging time [%]

50
-p
a
re
Cost per unit exergy (CCHP) [$/GJ]

lP

45
na

cCCHP-solst-a
40
cCCHP-st-a
cCCHP-solst-b
cCCHP-st-b
ur

35
Jo

30

25
62 63 64 65 66 67 68 69 70 71 72
Mass ratio of oil at charging time [%]
b
Fig. 11 Variation in CPUE for (a) generated electricity and (b) CCHP versus OMR

43
36 280

Investment cost rate (solst) [$/h]


35 278

Investment cost rate (st) [$/h]


34 276

33 Zsolst-a 274
Zst-a
Zsolst-b
32 Zst-b 272

31 270

30 268

f
oo
29 266
62 64 66 68 70 72

r
Mass ratio of oil at charging time [%]

70
-pa
re
65
lP

60
Exergoeconomic factor [%]

55
na

50 fsolst-a
fst-a
45 fsolst-b
ur

fst-b
40
Jo

35

30

25

20
62 63 64 65 66 67 68 69 70 71 72
Mass ratio of oil at charging time [%]
b
Fig. 12 Variation in (a) ICR and (b) exergoeconomic factor versus OMR.

44
Table 1 Input data to simulate the conventional and modified trigeneration systems.
Parameter Value
Ambient conditions [3]
Pressure 101.3 kPa
Temperature 298.15 K
High beam radiation rate 850 W/m2
Low beam radiation rate 500 W/m2

ORC and RORC [18]


ORC turbine efficiency 80%
ORC pump efficiency 80%
Electrical generator efficiency 97%
Pinch point temperature of ORC evaporator 40 oC

f
oo
ORC [18]
ORC turbine inlet pressure 2000 kPa

r
ORC pump inlet temperature 365 K
Mass flow rate of n-octane
-p 7.5 kg/s
re
RORC [This study]
lP

First and third ORC pumps inlet temperature 365 K


High pressure ORC turbine inlet pressure 2000 kPa
Pressure ratio of ORC turbines 2.5
na

Pressure ratio of first outlet in high pressure ORC turbine to


5
its inlet
ur

Pressure ratio of first outlet in low pressure ORC turbine to


8
its inlet
Jo

Mass flow rate of n-octane in first RORC 7.5 kg/s


Mass flow rate of n-octane in second RORC 6 kg/s

Absorption chillers [28]


Higher temperature in DEACH 393.15 K
Solution pump efficiency 95%
Effectiveness of solution heat exchangers 70%

Solar subsystem
Sun temperature [18,19,21] 6000 K
Therminol-66 inlet flow rate to HST 10 kg/s
Mass ratio of therminol-66 at charging time 65%
Effectiveness of solar heat exchanger 85%
PTSC outlet temperature 600 K
PTSC width 5.76 m
PTSC length 12.27 m
Incidence angle modifier 1a
Receiver inlet diameter 45 mm

45
Receiver outlet diameter 52 mm
Cover inlet diameter 70 mm
Cover outlet diameter 75 mm
Mirror reflectance factor 0.931
Intercept factor 0.93
Glass cover’s transmittance 0.94
Receiver’s absorbance 0.94
Stefan-Boltzmann constant 5.76 10-8 kW/m2 K4
Receiver emittance 0.92
Cover emittance 0.87

Exergoeconomic [17]
Maintenance factor 1.06

f
Component life time 20 year

oo
Annual interest rate 12%
Chemical engineering plant cost index in 2019 607.5

r
Optimization [32]
Individuals in the population
-p 32
re
Number of generations 64
Maximum mutation rate 0.25
lP

Minimum mutation rate 0.0005


Initial mutation rate 0.25
Crossover probability 0.85
na

a
For validation
ur
Jo

46
Table 2 Heat transfer coefficient relations for the parabolic trough solar collector [3,16,17].
Type of Local Relation
coefficient
Convection Between cover and Nu  k
ambient hc,ca = a a
Dc,o
0.4+0.54  Re0.52 0.1 < Re < 1000
Nua = 
4.364 1000 < Re < 50000

Radiation Between cover and hr,ca = εcv  σ  Tc +T0   Tc2 +T02 
ambient
Ac
hr,cr  Tr,av +  hc,ca + hr,ca   T0
Ar 
Tc =
A
hr,cr + c  hc,ca + hr,ca 

f
Ar

oo
Radiation Between cover and   Tc  Tr,av   Tc2  Tr,av
2

receiver h r,cr 
1 Ar  1 
   1

r
-p r Ac   cv 
Convection Inside receiver Nur  k r
hc,r,in =
re
Dr,in
0.023  Re0.8  Pr0.4 Re > 2300
lP

Nur = 
4.364 Re < 2300
na
ur
Jo

47
Table 3 Simulation relations for the thermal storage tanks [18-22].
HST CST
Charging period
̇ ̇ ̇ ̇ ̇ ̇
̇ [ ] ̇ [ ]

Q HST Q CST
THST = TCST =
CpHST  MHST CpCST  MCST
∑ ̇
∑ ̇
Storing period
Δt Δt c
+
THST = THST + h
 -UA HST ×  THST - T0   TCST
+
= TCST +  -UA CST ×  TCST - T0  
CpHST  MHST CpCST  MCST 

f
oo
Q HST,tlost  Cp,HST  M HST  THST  THST

 QCST,tlost = CpCST  MCST  TCST - TCST
+


r
Discharging period
Q out,HST =  Q HST - Q HST,tlost
-p Q out,CST =  Q CST - Q CST,tlost
re
lP
na
ur
Jo

48
Table 4 Formulation of energy and exergy relations and efficiencies for the conventional and
modified systems [25].
Parameter Relation
Net output electricity rate
̇ ∑ ̇ ∑ ̇

Cooling load ̇ ̇ ̇
{
̇ ̇

Heating load ̇ ̇
̇ { ̇ ̇ ̇ ̇
̇ ̇

f
oo
Electricity to cooling ̇ ̇
ratio
Electricity to heating ̇ ̇

r
ratio -p
Inlet exergy to the PTSCs
̇ [ (( ) ( ) ) (( ) ( ))]
re
̇ ̇
lP

Cooling exergy rate


̇ {
̇ ̇
na

Heating exergy rate ̇ ̇


̇ {
̇ ̇ ̇ ̇ ̇ ̇
ur

Electrical energy [ ̇ ̇ ]
Jo

efficiency
CCHP energy efficiency [( ̇ ̇ ̇ ) ̇ ]
Electrical exergy
[ ̇ ̇ ]
efficiency
CCHP exergy efficiency [( ̇ ̇ ̇ ) ̇ ]

49
Table 5 Exergoeconomic modeling information for the components of the modified setup [17,20,22,29].
Component Purchased cost Cost rate balance Auxiliary relation
PTSC ZPTSC = 230× Aap,tot ̇ ̇ ̇ ------
HST ZHST = 494.9+808  VHST ̇ ̇ ̇ ̇ cL,HST = c11
CST ZCST = 494.9+808  VCST ̇ ̇ ̇ ̇ cL,CST = c15
Solar HX ̇ ̇ ̇ ̇ ̇
ZSolHX = 130 ASolHX 0.093 c8 = c9
0.78

Storage pump-1 ̇ ̇ ̇ ̇ ̇ ̇ ̇
[ ] [ ]

Storage pump-2 ̇ ̇ ̇ ̇ ̇ ̇ ̇
[ ] [ ]

Solar pump ̇ ̇ ̇ ̇ ̇ ̇ ̇
[ ] [ ]

f
oo
EV-1 ZEV-1 = 0 ̇ ̇ ̇ ------
EV-2 ZEV-2 = 0 ̇ ̇ ̇ ------
̇ ̇ ̇ ̇ ̇
ZORC eva-1 = 309.14   AORC eva-1 
ORC evaporator-1

r
0.85 c3 =c4
ORC pump-1
ORC pump-2
[ ̇
[ ̇
]
]
-p ̇
̇
̇
̇ ̇
̇
̇
̇

̇
̇ ̇

̇
̇

̇
re
ORC turbine-1 [ ̇ ] ̇ ̇ ̇ ̇ ̇ ̇ c20=c21=c22
̇ ̇ ̇ ̇ ̇
ZHPHX-1 = 130   AHPHX-1 0.093
Heating process 0.78
c22 =c23
lP

HX-1
̇ ̇ ̇ ̇ ̇
ZHPHX-2 = 130   AHPHX-2 0.093
Heating process 0.78 c24 =c25
HX-2
na

Open feed heater-1 [ ̇ ] ̇ ̇ ̇ ̇ ------


[ ̇ ]
̇ ̇ ̇ ̇ ̇
ZORC eva-2 = 309.14   AORC eva-2 
ORC evaporator-2 0.85 c4 =c5
ur

ORC pump-3 [ ̇ ] ̇ ̇ ̇ ̇ ̇ ̇ ̇

ORC pump-4 ̇ ̇ ̇ ̇
Jo

[ ̇ ] ̇ ̇ ̇

ORC turbine-2 [ ̇ ] ̇ ̇ ̇ ̇ ̇ ̇ c33 =c34 =c35


̇ ̇ ̇ ̇ ̇
ZHPHX-3 = 130   AHPHX-3 0.093
Heating process 0.78
c35 =c36
HX-3
Open feed heater-2 [ ̇ ] ̇ ̇ ̇ ̇ ------
[ ̇ ]
High-pressure ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇
ZHPD = 17500   AHPD 100
0.6

desorber ̇ ̇ ̇ ̇
Low-pressure ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇
ZLPD = 17500   ALPD 100
0.6

desorber ̇ ̇ ̇ ̇
̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇
ZCond = 8000   ACond 100
Condenser 0.6

̇ ̇ ̇
̇ ̇ ̇ ̇ ̇
ZEva = 16000   AEva 100
Evaporator 0.6 c42 =c43
̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇ ̇
ZAbs = 16000   A Abs 100
Absorber 0.6

̇ ̇ ̇
Refrigerant EV-1 ZREV-1 = 0 ̇ ̇ ̇ ------
Refrigerant EV-2 ZREV-2 = 0 ̇ ̇ ̇ ------
̇ ̇ ̇ ̇ ̇
ZSHX-1 = 12000   ASHX-1 100
Solution HX-1 0.6 c48 =c49
Solution HX-2 ̇ ̇ ̇ ̇ ̇
ZSHX-2 = 12000   ASHX-2 100
0.6 c55 =c56

50
Solution pump ̇ ̇ ̇ ̇ ̇ ̇ ̇
[ ] [ ]
Solution EV-1 ZSEV-1 = 0 ̇ ̇ ̇ ------
Solution EV-2 ZSEV-2 = 0 ̇ ̇ ̇ ------

f
r oo
-p
re
lP
na
ur
Jo

51
Table 6 Validation of thermodynamic modeling (comparing the results of
this paper (conventional system) with Al-Sulaiman et al. [18]).
Parameter This work Ref. [18] Absolute percentage error [%]
̇ [kg/s] 7.09 7.1 0.14
̇ [kg/s] 13.92 13.9 0.14
̇ [kg/s] 4.18 4.2 0.47
̇ [kg/s] 9.74 9.7 0.41
T8 [K] 435.4 444.1 1.95
T9,sol [K] 405.9 405 0.22
T9,solst [K] 426.8 433 1.4
T10 [K] 576.3 576 0.05
T14 [K] 396 405 2.22
h19,sol [kJ/kg] 981.2 985.6 0.45
h19,solst [kJ/kg] 644.4 646.3 0.29
h19,st [kJ/kg] 810.3 806.2 0.51

f
oo
h20,sol [kJ/kg] 862.1 866 0.45
h20,solst [kJ/kg] 573.1 574.6 0.26
h20,st [kJ/kg] 710.3 706.8 0.49

r
-p
re
lP
na
ur
Jo

52
Table 7 Comparison of thermodynamic and exergoeconomic results for the conventional and
modified systems.
Term Solar mode Solar and storage mode Storage mode
Relative Relative Relative
percentage percentage percentage
MS CS MS CS MS CS
difference difference difference
[%] [%] [%]
̇ [kW] 880.3 797.1 10.4 640.2 540.9 18.4 613.9 596.1 3.0
rel/c [-] 2.56 3.99 -36.0 1.86 2.71 -31.4 1.78 2.98 -40.3
rel/h [-] 0.202 0.180 12.5 0.207 0.185 12.0 0.199 0.188 5.6
ηel [%] 15.3 13.9 10.4 6.2 5.3 18.4 9.2 8.9 3.0
ηCCHP [%] 97.2 94.6 2.7 39.8 35.8 11.2 60.8 59.5 2.1
εel [%] 7.7 7.0 10.4 3.3 2.8 18.4 4.9 4.7 3.0

f
εCCHP [%] 12.1 11.3 6.3 5.1 4.5 14.3 7.6 7.2 6.1

oo
̇ [kW] 1693.0 1689.7 -0.2 4807.2 4775 -0.7 899.5 902.9 0.4
cel [$/GJ] 39.9 41.4 3.8 29.8 28.8 -3.4 43.3 44.8 3.5

r
[$/GJ] 50.3 47.6 -5.4 30.7 29.3 -4.6 47.7 46.7 -2.1
̇
̇
[$/h]
[$/h]
114.4
247.1
131.1
251.2
14.6
1.7
58.1
102.7
-p 66.3
113.5
14.1
10.5
92.9
174
99
173.7
6.6
-0.2
re
̇ [$/h] 254.2 259.1 1.9 269.3 275.5 2.3 30.1 35.2 16.9
̇ [$/h] 112.5 112.4 -0.1 170.4 168.3 -1.2 69.0 73.2 6.1
lP

ftot 69.3 69.7 0.6 61.2 62.1 1.5 30.4 32.5 6.9
na
ur
Jo

53
Table 8 Thermodynamic and exergoeconomic optimization results for the modified setup.
Parameter Solar mode Solar and storage mode Storage mode
ω1 [-] 0.5 0.0 1/3 0.5 0.0 1/3 0.5 0.0 1/3
ω2 [-] 0.0 0.5 1/3 0.0 0.5 1/3 0.0 0.5 1/3
ω3 [-] 0.5 0.5 1/3 0.5 0.5 1/3 0.5 0.5 1/3
P20 [kPa] 2496 2438 2324 2310 2238 2468 2333 2451 2479
P20/ P33 [-] 2.36 2.66 2.33 2.87 2.68 2.00 2.55 2.59 2.47
P20/ P21 [-] 5.19 5.89 4.86 5.82 5.93 6.36 4.10 4.13 4.10
P33/ P34 [-] 8.18 8.79 9.55 7.42 7.12 7.38 7.15 7.42 9.42
T25 and T36 [K] 360.1 360.2 360.1 360.0 360.0 360.8 360.0 360.4 360.6
Tpp-ORC eva [K] 48.6 48.7 48.4 48.5 49.5 49.0 37.2 40.4 32.7
T1 [K] 613.0 592.5 605.8 584.7 583.3 595.0 648.6 635.0 643.7
ORM [%] ----- ----- ----- 62.0 62.1 62.1 71.9 72 71.9
̇ [kW] 941.0 947.5 944.3 759.6 768.5 766.3 760.0 764.3 762.9
ηel [%] 16.45 16.51 16.45 7.44 7.49 7.47 10.35 10.33 10.36
ηCCHP [%] 97.15 97.15 97.15 44.19 44.49 44.72 62.93 62.58 63.12
εel [%] 8.31 8.27 8.29 3.97 3.94 3.96 5.49 5.47 5.48

f
εCCHP [%]

oo
12.58 12.52 12.55 6.04 6.00 6.01 8.29 8.23 8.28
̇ [kW] 1677 1680 1679 4815 4822 4817 1011 1029 1017
cel [$/GJ] 40.80 41.44 41.32 26.82 27.64 26.97 41.27 41.45 41.65

r
cCCHP [$/GJ] 46.23 46.71 46.33 28.14 28.23 28.26 43.81 44.01 44.02
̇ [$/h]
̇
̇ [$/h]
[$/h]
138.7
253.3
141.3
254.3
140.5
255.9
-p
76.4
127.7
76.5
126.7
74.4
126.9
113.4
194.6
113.8
195.0
114.5
195.9
re
261.7 262.9 264.5 281.4 283.4 280.3 36.8 37.4 37.9
̇ [$/h] 110.0 111.0 110.7 168.7 169.4 168.3 79.9 79.8 79.3
ftot [%] 70.40 70.30 70.50 62.52 62.59 62.48 31.55 31.90 32.33
lP
na
ur
Jo

54
Highlights:

 Modification and optimization procedures are conducted for a solar-powered system.

 The goals are to enhance the exergetic and exergoeconomic outcomes.

 Three different solar radiation operational modes are considered during a day.

 Comprehensive parametric study is performed for the radiation operational modes.

 The exergy efficiency and exergoeconomic factor are improved by the modification.

of
ro
-p
re
lP
na
ur
Jo
Declaration of interests

☒ 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.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like