You are on page 1of 12

Applied Thermal Engineering 181 (2020) 115926

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Optimization and dynamic techno-economic analysis of a novel PVT-based T


smart building energy system
Amirmohammad Behzadi , Ahmad Arabkoohsar, Yongheng Yang

Department of Energy Technology, Aalborg University, Denmark

HIGHLIGHTS

• Transient simulation of a smart building interacted with local grids is modeled.


• The temperature of panels plays a crucial role in the techno-economic indicators.
• PVT panels are found to be the main source of irreversibility in the system.
• AThesignificant reduction in capital costs is achieved by removing the battery.
• exergy efficiency increases by 4.03%, and the total cost decreases by 3.64 €/MWh.

ARTICLE INFO ABSTRACT

Keywords: There is a variety of solar-based energy system designs for buildings. Although these systems are economically
Smart building energy systems profitable, reducing the energy cost of the buildings over time, their penetration has not been that impressive yet
PVT panels due to their high initial cost. In this study, an energy system comprising a few PVT panels (without any batteries)
Dynamic simulation and a heat storage tank is proposed and investigated for smart buildings with two-way interactions with both
Thermodynamic analysis
heat and electricity grids. Removing the battery from the system would result in a sharp reduction of the cost of
Exergy and cost optimization
the system and, thereby, will make incentives for the end-users to adopt the solution. This novel system will not
only supply the buildings’ real-time electricity and domestic hot water needs but also will compensate for a
significant portion of the buildings’ energy expenses by selling the surplus generations to the electricity and heat
networks. The dynamic model of the proposed system is comprehensively analyzed from thermodynamic and
economic points of view using TRNSYS software. Additionally, defining the overall annual exergy efficiency, and
the total product cost as the objective functions, optimization of the design and size of the system employing the
TRNOPT tool has been done. It is shown that the optimized system results in 16.7 €/MWh and 7.7 €/MWh lower
energy costs for electricity and heat of the buildings compared to when the buildings’ demand is only supplied by
heat and electricity grids.

1. Introduction the true definition of smart energy systems [3]. To achieve these, the
electricity grids should be reconfigured into smart grids, and other
Due to the increase in energy consumption and demand for elec­ energy distribution networks (e.g., district cooling and heating, if any)
tricity and heat in households, residential and commercial buildings are should be compatible with the new buildings and energy trading po­
considered as a significant source of greenhouse gas emissions. licies [4]. Low and ultra-low temperature district heating systems are
Buildings are responsible for 40% of domestic primary energy con­ samples of the proposed solutions for the future district heating sys­
sumption and more than 30% of worldwide natural gas consumption tems, for instance [5,6].
[1]. Consequently, many countries have already begun to develop en­ Smart energy buildings are of the essential elements of future smart
ergy conservation plans in their constructions like a policy called 20-20- energy systems [7]. Smart energy buildings have a highly efficient
20, which leads to a 20% increase in the energy efficiency of buildings energy performance and own their individual renewable energy sys­
[2]. Another critical aspect of these reviews in the energy sectors is to tems while also being connected to the local energy distribution grids.
increase the share of renewable energy and move practically towards Smart buildings are supposed to have two-way interactions with the


Corresponding author.
E-mail address: ambe@et.aau.dk (A. Behzadi).

https://doi.org/10.1016/j.applthermaleng.2020.115926
Received 7 April 2020; Received in revised form 31 July 2020; Accepted 15 August 2020
Available online 29 August 2020
1359-4311/ © 2020 Elsevier Ltd. All rights reserved.
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

Nomenclature Subscripts and abbreviations

A Area, m2 0 Ambient
c Product cost, €/MWh CHP Combined heating and power
CP Specific heat capacity, kJ/kg·K Conv Convection
CRF Capital recovery factor D Destruction
Ė Electricity, kWh DHW Domestic hot water
Ex Exergy, kWh F Fuel
G Total incident solar radiation, kW/m2 i Inlet
h Enthalpy, kJ/kg INV DC to AC inverter
ir Interest rate i ith segment
LC loss coefficient between the tank and its environment, kJ/ o Outlet
kg·K·m2 P Product
M Molar mass, kg/kmol PV Photovoltaic
m Mass flow rate, kg/s PVT Photovoltaic/Thermal
mh Water mass flow rate from the PVT, kg/s PY Present year
mL Water mass flow rate to the load, kg/s rad Radiation
NT Number of tubes RY Reference year
P Pressure, kPa SES Smart energy system
Qh Energy input to the tank from the hot fluid stream, kWh TMY Typical meteorological year
QL Energy removed from the tank to supply the load, kWh tot Total
Q loss,top,conv Energy lost through convection at the top surface, kWh
Q loss,top,rad Energy lost through radiation at the top surface, kWh Greek letters
Q loss,back Energy lost at the back of the collector, kWh
Qu Energy added to the flow stream, kWh
τα transmittance-absorptance product for the PVT panel
RB The heat transfer resistance through the back of the PVT,
ηI Energy efficiency
K/kWh
ηII Exergy efficiency
Rk Costs not related to the investment cost and exergy of
γ Fixed operation and maintenance costs coefficient
product, €/h
τ Annual component operation hours
T Temperature, °C
ω Variable operation and maintenance costs coefficient
Ż Cost rate of components, €/h
ŻCI Capital investment cost rate of components, €/h
ŻOM Operating and maintenance cost rate of components, €/h

energy distribution grids as well. Two-way interaction means buying the increase of the PVT area and decrease of heat storage volume in­
energy from the grids whenever the building’s energy systems are not creases the system profit. Entchev et al. [15] compared the dynamic
sufficiently producing and selling the overproduction of the building performance of a hybrid PVT based renewable energy system against
energy system to the grids, for peak-shaving, etc. [8]. Among renew­ the conventional system utilizing a boiler and chiller of a separated
able-based energy resources for buildings, solar-driven systems are the house and office building in Canada. They demonstrated that the de­
most popular choices and available in a variety of designs. Solar-based pendence on the electricity grid is decreased due to the increase in
energy systems create clean, renewable electricity/heat from the sun overall energy saving up to 58%. Comparative dynamic performance
and benefit the environment due to net-zero greenhouse gas emissions, evaluation of a ground source heat pump integrated with PVT and fuel
particularly carbon dioxide (CO2). PV panels with a battery unit for cell system for residential and office buildings located in Italy was
two-way electricity supply for forming net-zero electricity buildings, performed by Canelli et al. [16]. Their results indicated that the re­
solar thermal panels coupled to a thermal storage unit for providing the duction of CO2 emission rate and the operational costs reduction of the
heat demand and domestic hot water of the buildings, and more re­ PVT based system are 36.2% and 28.4% higher than the fuel cell-based,
cently PVT panels with a battery and a thermal storage unit (with and respectively. The influence of operating temperature of PVT and storage
without heat pumps) are of the common types of building solar energy tank volume on overall performance indicators and maximum PVT ef­
systems [9]. ficiency of a PVT system integrated with air temperature and humidity
A PVT set integrated with a battery and a thermal storage system, is regulation unit was carried out by Hu et al. [17]. They showed that at
much helpful for increasing the contribution of buildings in peak- the PVT temperature of 101 °C and storage tank volume of 16.7 L, the
shaving and cost reduction of energy grids, and increasing the share of maximum overall efficiency of 49.77% is achieved.
renewable energy as they cogenerate heat and electricity [10]. In Considering prime energy-saving efficiency to be maximized and
comparison to PV panels, extracting useful thermal energy from the life cycle saving to be minimized as two conflictive objectives, multi-
same aperture area of PVT panels leads to better performance, more objective optimization of a building integrated PVT system for re­
capability of being integrated with other energy technologies, and a sidential hot water application was dynamically performed by Chen
higher overall (electrical plus thermal) efficiency [11]. Since they can et al. [18]. The results of the scatter distribution of the significant
work with little deterioration for more than 20 years, reliability and variables indicated that the optimal value of tank volume and mass flow
life-time are further benefits of PVT panels [12]. In a recent study, a rate change from 99.5 L to 218.6 L and 0.0085 kg/s to 0.011 kg/s,
comparative performance study of a building integrated PVT panels and respectively. The comparison of conventional exergy analysis and ad­
a thermal wheel system from energy and exergy viewpoints has been vanced exergy assessment of a novel energy storage system was studied
studied by Shahsavar and Khanmohammadi [13]. Using TRNSYS soft­ by Liu et al. [19]. They demonstrated that compressor is the most es­
ware, dynamically energy, and thermoeconomic assessment of a solar sential equipment from the quality of energy conversion aspect. Kamel
trigeneration system of a hospital located in Naples was investigated by and Fung [20] compared the dynamic performance evaluation of a PVT
Buonomano et al. [14]. The result revealed that at the optimum region, system against the air source heat pump in a residential building. They

2
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

concluded that using renewable energy resources results in a decrease best operating condition.
in annual CO2 emission and electricity cost by 1734.7 kg and 500 $,
respectively. The variation of electrical power and temperature of heat 2. Cycle description and assumptions
source for a real weather condition in Denmark of a solar-based system
consisting of PVT panels, heat pumps, and heat storage tank was in­ Fig. 1 illustrates the pictorial representation of the proposed smart
vestigated by Dannemand et al. [21]. They showed that PVT panels building system in the TRNSYS studio environment. As seen, the system
supply 60% of the overall heat delivered to the system in January and consists of four PVT panels, heat storage tank, controllers, pumps, and
February. Dynamic investigation of a solar-based system equipped with auxiliary valves. To have a better insight, the schematic diagram of the
PVT panels for providing hot water and power demand of Tunisian proposed smart system is shown in Fig. 2. As depicted, the cold water of
residential houses was carried out by Hazami et al. [22]. They resulted 15 °C enters the tempering valve (state 1). After being monitored by the
that annual thermal and electric efficiencies are 50% and 15%, re­ PVT hot side outlet temperature (T9), it is split proportionally between
spectively. In another study, a thermodynamic analysis of a novel CO2- heat storage tank (state 2) and tee piece 1 (state 3) to provide domestic
based energy storage system was carried out by Liu et al. [23]. They hot water of 45 °C [26]. Since the PVT outlet temperature falls re­
concluded that particular attention should be paid to energy storage markably, controller 1 decides whether the system turns on or off.
due to the sharing of 19.23% of the total irreversibilities of the overall When the system is on, if there is any demand for heat (state 10) or the
system. A smart building system consisting of a PVT panel controlled by capacity of the tank is not yet filled, controller 2 turns the pump 1 on.
an air handling unit under a real climate condition was designed and Otherwise, it turns on the pump 2 to provide hot water for selling to the
investigated by Dahmane et al. [24]. The results indicated that the heat local ultralow-temperature district heating grid (state 14). Monitoring
exploited from PVT panels via ambient air provides 24.2% of the the average temperature of the tank, it is decided when pump 3 turns on
thermal energy demand of the building in winter. They also showed to provide hot water of 40 °C [27] for the local ultralow-temperature
that the electrical power of the PVT panels at an optimum angle in­ district heating grid. Aside from hot water production for building use
creases dramatically in summer. Liu et al. [25] proposed a novel and selling to the district heating network, the PVT panels produce
building integrated PVT system consisting of a heat pump. They re­ electricity which firstly, supplies the building’s real-time electricity
sulted the power consumption and heating coefficient of 3.12 W and demand (state 21) and secondly, is used to further charge the tank in
57.9, respectively. the form of hot water via an electrical coil (state 22). If the tank is fully
The above-discussed scientific works and many other research and charged, the surplus production is sold to the electricity grid. Naturally,
practical projects are examples of solutions proposed for integrating by the time of low electricity production and higher demand, the extra
PVT panels in building energy systems. Despite the promising techno- required electricity is purchased from the grid (state 23).
economic results of almost all of these investigations, PVT panels have The description of the system components, named as type in
not been that broadly used in buildings yet. One of the biggest chal­ TRNSYS, along with the input parameters, are tabulated in Table 1. In
lenges of PVT-integrated building energy systems is the high cost of this regard, input parameters are selected size enough to not only
capital, which the building owners need to pay for that. Although this supply a significant portion of the building’s electricity and domestic
investment goes back to the owners gradually during a fairly short hot water demand but also compensate a major fraction of the build­
period, the initial cost seems a serious challenge for that yet. This work ing’s energy cost by selling its excess heat and electricity to the local
proposes a novel method of employing PVT panels in smart buildings in heat and power grids.Table 2.
which the initial cost of the system is significantly reduced. This is
mainly done via the elimination of the battery unit and the heat pump, 3. Modeling
which are the two major costs of the system. Therefore, the proposed
system includes some PVT panels integrated with a heat storage unit To perform energy, exergy, and economic assessment, the dynamic
only. The system is to make a highly cost-effective and reliable inter­ modeling of the proposed smart building system consisting the PVT
action between the smart building and both the heat and electricity panels, the heat storage tank, pumps, valves, and controllers is analyzed
grids. By this smart configuration, not only a significant portion of the via TRNSYS software as a powerful tool in stimulating renewable en­
building’s real-time electricity demand is supplied, but also its entire ergy systems in transient mode [31]. Besides, the influence of major
domestic hot water need is provided. Besides, by selling surplus heat decision variables on net electricity, energy and exergy efficiencies, and
and electricity production of the system to the local grids, a significant total product cost as objective functions is investigated via parametric
portion of the building’s energy cost is compensated. In a nutshell, to study. Furthermore, finding the best operating condition from perfor­
cover the literature gap in this area, the main novelties of this research mance and economic standpoints, the proposed smart system is opti­
could be summarized as below: mized.

• To propose a novel configuration of employing PVT panels for


3.1. Energy and exergy analysis
electricity generation and hot water production in a real residential
(just to make the result as realistic as possible, a case study in
Energy and exergy analysis, as criteria for evaluating the quantity
Esbjerg city, Denmark, is considered for the simulations).

and quality of energy conversion of the proposed smart system, are
To model a smart building energy system interacting with local
carried out, respectively. In this regard, mass, energy, exergy balances
electricity and district heating networks.

in the transient model for each component are written as below, re­
To decrease the investment costs by eliminating the major costs of
spectively [32]:
the system (battery and heat pump) to motivate consumers to adopt
and invest in such smart energy systems. min = mout (1)
• To analyze the proposed smart energy system dynamically from
energy, exergy, and economic point of view. Q W = mout hout min hin (2)
• To perform a parametric study to evaluate the effect of major de­
cision variables on annual performance indicators and economic ExQ ExW = mout exout min exin + ExD (3)
indicant.
• To optimize the proposed smart system, maximizing the overall where D, W, Q subscriptions correspond to the exergy destruction, ex­
exergy efficiency and minimizing the total product cost. ergy of work and heat, respectively. Neglecting the potential and ki­
• To determine the optimal value of major parameters to obtain the netic exergies, the overall specific exergy is equivalent to the sum of

3
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

Fig. 1. Pictorial view of the TRNSYS model for the proposed smart building system in the simulation studio environment.

Fig. 2. Schematic diagram of the proposed smart building.

4
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

Table 1
Description of the system components in TRNSYS (types) and input parameters [28–30].
PVT panel Type 155 applies a link with Matlab on which the thermodynamic formula of PVT panels is written. A PVT panel not only generates electricity from photovoltaic
cells but also recovers waste heat of panel via water passing through tubes bonded to an absorber plate located beneath the PV cells.
Number of PVT panels 4
l (m) 2
w (m) 2
Number of tubes 20
Bond thickness (m) 0.01
Bond thermal conductivity (kJ/h·m·K) 1386
Tube diameter (m) 0.01
Reflectance 0.15
Emissivity 0.9
1st order IAM 0.1
U-value of roof material (kJ/h.m2.K) 0.667
PV cell reference temperature (°C) 30
PV cell reference electrical efficiency (%) 20
PV cell reference thermal efficiency (%) 50
2
PV cell reference radiation (kJ/h.m ) 3600
2
Efficiency modifier-radiation (h⋅m /kJ) 0.000025
Electrical efficiency modifier-temperature (%/°C) −0.5
Thermal efficiency modifier-temperature (%/°C) −0.2
Heat storage tank Type 4 models a multi-node heat storage tank with an internal electrical coil generating hot water from input electricity. In this model, an accurate calculation is
achieved if the parameters are determined in a logical domain.
V (m3) 0.3
Tank loss coefficient (kJ/h.m2.K) 2.5
Number of nodes 6
Height of node 1,…,6 (m) 0.3
Tset point for Tank (°C) 50
The maximum heating rate of element for the tank (kJ/h) 9000
Pump Type 3b models a pump with a user-specified maximum mass flow rate and maximum electricity depending on an input control signal having a value between 1
and 0.
Maximum electricity (kJ/h) 240
η Pump (%) 90
Heat exchanger Type 91 models a constant effectiveness heat exchanger, which is independent of the system configuration.
Heat exchanger effectiveness 0.85
Valves Type 11b and 11f model the tempering valve and flow diverter controlled by an external input signal, respectively. In addition to an input signal, type 11f works
according to a set point temperature determining the split ratio of outlets to reach a specific temperature. Type 11h represents the tee piece mixing two inlet
water streams into single outlet water.
Weather data Type 109 is utilized to explain the annual weather information of Esbjerg city in Denmark in the form of a typical meteorological year (TMY2) format.

Table 2
The rate of exergy destruction, the exergy of product, and exergy of fuel for the ith component.
Component Exergy destruction Exergy of product Exergy of fuel

Tempering valve ĖxD,T.V=Ėx1-Ėx2-Ėx3 ĖxP,T.V= Ėx2+Ėx3 ĖxF,T.V=Ėx1


Heat storage tank ĖxD,tank =Ėx2+Ėx8+Ėx22-Ėx4+Ėx9 ĖxP,tank=Ėx4-Ėx2 ĖxF,tank = (Ėx8-Ėx9)+ Ėx22
Pump 1 ĖxD,Pump1= Ėx4+Ėx18-Ėx5 ĖxP,Pump1=Ėx5-Ėx4 ĖxF,Pump1=Ėx18
Tee piece 2 ĖxD,T.P2=Ėx5+Ėx16-Ėx6 ĖxP,T.P2= Ėx6 ĖxF,T.P2=Ėx5+Ėx16
PVT ĖxD,PVT=Ėx6+Ėx17-Ėx7- Ėx19 ĖxP,PVT=Ėx19+(Ėx7-Ėx6) ĖxF,PVT=Ėx17
Diverter ĖxD,Div=Ėx7-Ėx8-Ėx11 ĖxP,Div=Ėx11+Ėx8 ĖxF,Div=Ėx7
Inverter ĖxD,Inv=ĖPVT,DC-ĖPVT,AC ĖxP,Inv= ĖPVT,AC ĖxF,Inv= ĖPVT,DC
Tee piece 1 ĖxD,T.P1=Ėx3+Ėx9-Ėx10 ĖxP,T.P1=Ėx10 ĖxF,T.P1=Ėx3+Ėx9
Pump 3 ĖxD,Pump3=Ė20+Ė12-Ė13 ĖxP,Pump3=Ė13-Ė12 ĖxF,Pump3=Ė20
Heat exchanger ĖxD,HEx =Ėx11+Ėx13-Ėx14-Ėx15 ĖxP,HEx=Ėx14-Ėx13 ĖxF,HEx=Ėx11-Ėx15
Pump 2 ĖxD,Pump2=Ėx19+Ėx15-Ėx16 ĖxP,Pump3=Ėx16-Ėx15 ĖxF,Pump3=Ėx19

physical and chemical exergies as following: Qloss, top, rad = hrad A (T¯PVT Tsky ) (8)
ph
ex i = ex i ex ich (4) T¯abs T¯inside
Qloss, back = A
RB (9)
ex iph = (hi h0) T0 (si s0 ) (5)
Qu = m6 CP (T7 T6) (10)

3.1.1. The PVT panels Furthermore, in Eq. (6), Qabsorbed is the energy rate absorbed by the
Defining Qloss,top,conv, Qloss,top,rad, Qloss,back, and Qu as the energy rate PVT panel (does not contain PVT produced electricity) which is written
lost to the ambient through convection and radiation at the top surface as follows [33,34]:
and back of the panel and the energy rate added to the water, respec­
Qabsorbed = G (l × w )( )(1 PV ) (11)
tively, the energy balance is written as following [30,33]:
Here, l, w, and G denote the length and width of the panel and total
Qabsorbed = Qloss, top, conv + Qloss, top, rad + Qloss, back + Qu (6) incident solar radiation, and η and τα are PV efficiency and the trans­
mittance-absorptance product for the solar panel, respectively.
Qloss, top, conv = houter A (T¯PVT Tamb) (7) Eventually, the PVT produced electricity can be calculated as below

5
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

[35]: Table 3
Purchased cost for each component [40–43].
EPVT = G (l × w )( ) PV (12)
Component Zk ($)

Heat storage tank ZTank = c1V


3.1.2. The heat storage tank c1 = 7581.066 €/m3
Transiently, the energy balance for the ith node of the heat storage PVT panel ZPVT = c2 APVT
tank is calculated as following [36]: c2 = 898.55 €/m2
Pump ZPump = c3 f (W Pump )0.71 c3 = 1603.91 €/(kWh)0.71
dT
mi CP i = i mh CP (Th Ti ) + i mL CP (TL Ti ) + UAi (Tenv Ti ) f=1+
0.2
dt 1 P
Heat exchanger
+ i CP (Ti 1 Ti ) + i CP (Ti Ti + 1) + Qi (13) ZHEX = c4 ( 0.093)
AAHX 0.78
c4 = 390 $
Valve (Diverter, tee piece, ZValve = 135.4 €/unit
Here, Th, TL, mh , and mL denote the temperature of the water going tempering valve)
into the tank from the PVT and the temperature of the water to the load, Controller ZController = 266.4 €/unit
water mass flow rate from the PVT, and to the load/grid, respectively.
Also, αi, βi, and γi correspond to the different control signals based on
smart system design. Furthermore, Qi and U denote the electricity re­ CRF =
ir (1 + ir ) n
ceived via the electrical coil to heat the water and the loss coefficient (1 + ir )n 1 (20)
between the tank and environment, respectively. Finally, QL and Qh are
the sensible energy rate extracted from the heat storage tank proving Eventually, using Marshal and Swift cost equation and defining CI as
the load demands and the energy rate input to the heat storage tank the cost index, the value of Żk is converted from reference to the present
from PVT output hot water as following, respectively [33]: year as follows [39]:
PY CIPY
QL = mL CP (T1 TL ) (14) Zk = Zk ×
CIRY (21)
Qh = mh CP (Th T6) (15)

What stands out from Eqs. (11) and (12) is that the load/grid-sup­ 3.3. Performance evaluation
plied water always leaves the first node, and the PVT output water
enters the sixth node. For evaluation the performance of the proposed smart building
To carry out the exergy assessment, the exergy efficiency for the ith system from energy and exergy point of views, a net produced elec­
component is defined as below: tricity, the electricity to/from the grid, energy, and exergy efficiencies
are written as following, respectively:
ExP, i
=
II , i
ExF , i (16) Enet = EPVT EHeater EPumps (22)

where ExP , i and ExF , i are the rate of exergy of fuel and product, re­ Egrid = Eload Enet (23)
spectively. In this regard, the exergy destruction rate, along with the
rate of exergy of fuel and product for each component of the proposed Enet + Qload + Q grid
smart building system, are tabulated in Table 2. = × 100
(24)
I
QSun

3.2. Economic analysis Enet + Ex10 + Ex14


= × 100
II
Ex17 (25)
To evaluate the proposed smart system from an economic stand­
point, the total annual product cost is calculated and compared with the Here, EHeater is the needed water to run the electrical coil for hot
electricity and heat production costs due to the use of the conventional water production. In Eq. (23), when Eload is more than Enet , the load
method in Denmark. In this regard, the total cost of the system, which is need is provided from the grid. Conversely, the net produced extra
electricity should be sold to the electrical grid. Also, Qload and Qgrid
equal to the sum of the annual Levelized capital investment cost and the
denote the heats supplying the load need and sold to the ultra-low
yearly Levelized operation and maintenance cost is calculated as fol­
temperature local grid, respectively. Accordingly, considering Denmark
lows:
as the case study, the temperatures of the load and ultra-low local space
Zk = Zk + Zk
cl OM
(17) heating grid are assumed to be 45 °C and 40 °C, respectively. Also, E17 in
Eq. (25) is the rate of exergy emanated from the sun, which is measured
cl CRF as follows:
Zk = ( ) Zk
k (18)
4
1 T0 4 T0
Ex17 = 1 + QSun
OM
Zk =( k
) Zk + k EP , k + Rk 3 TS 3 TS (26)
k (19)

Here, τk denotes the annual working hours, and γk and ωk are the QSun = 4 × G × l × w (27)
coefficients corresponding to the fixed and variable operation and where TS and T0 are, respectively, the sun and ambient temperatures
maintenance costs, respectively. Also, EP, k and Rk are the exergy of based on Esbjerg hourly weather data. Moreover, the total product cost
product and other costs not related to the investment cost and exergy of in €/kWh as the economic index is calculated as the following:
product, respectively. In Eq. (19), the second and third terms are neg­
ligible because of being small compared with the first term [37,38]. nk PY
Z
k=1 k
Also, the purchased cost for each component (Zk) is tabulated in cP, total = np
i=1
Ex pi (28)
Table 3. Moreover, defining ir and n as the interest rate and the number
of operating years, the capital recovery factor (CRF) is written as fol­ Finally, the PV electrical efficiency as the PVT performance in­
lows: dicator from electricity conversion standalone is calculated as follows:

6
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

EPVT
= × 100
PV
4×G×l×w (29)

4. Results and discussion

In this work, for making the simulations and assessments close to


real states, a typical house with real heat and electricity demand pro­
files as well as real availability of resources such as solar irradiation,
etc., has been considered as the case study.
Implementing TRNSYS as a far-reaching tool in the simulation of
smart energy systems in transient mode, a parametric study is carried
out to assess the performance of the proposed system from thermo­
dynamic and economic standpoints. Besides, the optimum values of
major decision variables, as well as the optimal values of annual exergy
efficiency and total product cost, are determined to perform single-
objective optimization using the TRNOPT tool of TRNSYS.

4.1. Case study Fig. 3. Average daily electricity load profile [44].

The mentioned case study of this work is a detached house in the


city of Esbjerg, located in the southwest of Denmark. Accordingly, the
typical daily hot water draw-off for a Danish single-family house is
tabulated in Table 4 [5]. Moreover, the average daily electricity load
profile of a Danish house is illustrated in Fig. 3 [44].

4.2. Parametric study

The standard method of sizing solar-based building energy systems


is seeking for an equal annual production and consumption of elec­
tricity. Here also the same principle is taken. Therefore, the primary
constraint for the system sizing and the number of PVT panels is the
almost equal values of electricity bought/sold from/to the electricity
networks over an entire year. The variation of annual electricity
bought/sold to/from the grid with the number of PVT panels is illu­
strated in Fig. 4. As observed, as the number of panels increases, while
the value of electricity sold to the network increases, the value of Fig. 4. Variation of annual electricity bought/sold to/from the grid with the
electricity bought from the network decreases. So, to supply the number of PVT panels.
building’s real-time demand, the dependence on the local network will
decrease. Here, since the conjunction of the two lines is near to 4 pa­
nels, choosing 4 panels leads to the almost equal values of annual variables is carried out. For this, major decision variables are the PVT
electricity bought/sold from/to the electricity. outlet temperature, number of identical water tubes bonded to the
Considering net electricity, a net produced hot water volume, and absorber plate of the PVT, the length of the PVT at a constant width, the
exergy efficiency as the performance indicators along with the total volume of the heat storage tank, and the average tank loss coefficient
product cost as an economic indicant, the parametric study of major per unit area.
Since the operating temperature plays a crucial role in the PVT
Table 4 panels performance, the influence of the outlet temperature on the
Draw-off profile of a Danish single-family [5]. thermodynamic/economic aspects of the proposed smart energy system
Day
is shown in Fig. 5. As the PVT outlet temperature increases, the panel
electrical and thermal efficiencies decrease; so, the net produced elec­
Time Volume (m3) Duration (s) tricity, and total volume of generated hot water will decrease, respec­
tively, as depicted in Fig. 5(a). Referring to Eq. (25), since the increase
10:35 0.125 600
10:45 0.015 150
of outlet temperature of PVT leads to a lower net produced electricity
11:05 0.125 600 and produced hot water volume, at a constant rate of exergy emanated
11:15 0.015 150 from the sun, the trend of variation of overall exergy efficiency is the
11:35 0.042 300 same as the trend of net electricity variation as illustrated in Fig. 5(b).
11:55 0.042 300
Also, the figure reveals that when the PVT outlet temperature increases
Night 40 °C, the total product cost increases from 19.9 €/MWh to 18.2
€/MWh, which is rational due to the decrease in the exergy of product,
Time Volume (m3) Duration (s)
at a constant total capital investment and operating and maintenance
22:35 0.125 600 costs as inferred from Eq. (28).
22:45 0.015 150 PVT area is a critical decision parameter which linearly affects the
23:05 0.125 600 purchased cost of the panel as inferred from Table 3. At a constant
23:15 0.015 150
width, the effect of the length of the PVT panels is presented in Fig. 6.
23:35 0.042 300
23:55 0.042 300 According to Eq. (12), by increasing the length of panels, at a constant
width, the PVT output generated electricity is increased, so the net

7
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

Fig. 5. The influence of the PVT outlet temperature on thermodynamic and economic indicators.

Fig. 6. The influence of the length of PVT panels at a constant width on thermodynamic and economic indicators.

electricity increases too. In addition, according to Fig. 6(a), at constant respectively, the numerator and denominator of Eq. (25) and Eq. (28);
width, as the panel length increases from 0.5 m to 3.5 m, the yearly ergo, the yearly overall exergy efficiency increases, and the total pro­
produced hot water increases by about 950 m3 which is due to the rise duct cost decreases. Noteworthy, Fig. 7 exhibits that when the number
of the energy rate absorbed by the PVT panel to heat the water as in­ of tubes varies from 20 to 32, the value of exergy efficiency decreases,
ferred from Eq. (11). What stands out from Fig. 6(b) is that a larger and the product unit cost remains rather constant. Consequently, the
panel length means higher overall exergy of product, so higher overall PVT panels with 20 tubes are selected as a suitable option for the
exergy efficiency, as well as lower total product cost, are expected. proposed smart system.
Another characteristic of the PVT panel is the number of water tubes Due to the simultaneous interaction between the heat source, i.e.,
bonded to the absorber plate. Fig. 7(a) reveals that the net generated PVT panels and the heat demand, i.e., building, the heat storage tank
electricity increases by about 7.3 kWh, as the number of tubes increases equipped with an electrical coil is the other vital equipment in the
from 5 to 32. What the figure further depicts is that the energy rate proposed system. In this component, the volume of the tank is an es­
added to the water increases with an increase in the number of the tube. sential physical parameter which linearly affects the capital investment
Hence the annual produced hot water increases by about 19 m3. Re­ and operating and maintenance costs of the tank, as inferred from
ferring to Fig. 7(b), rising the number of PVT tubes increases, Table 3. Also, if the volume of the tank is not selected accurately, the

Fig. 7. The influence of the number of the tube on thermodynamic and economic indicators.

8
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

Fig. 8. The influence of the volume of the tank on thermodynamic and economic indicators.

Fig. 9. The influence of the average tank loss coefficient on thermodynamic and economic indicators.

portion of the renewable-based resource, i.e., the solar system, in pro­ 4.3. Results of exergy and economic analysis
viding the building’s real-time hot water demand will decrease. A
bigger heat storage tank results in an increase in operating hours of The findings related to energy, exergy, and economic assessment are
pump 2 and 3 to be filled; therefore, the value of electrical consump­ tabulated in Table 5. In this regard, the value of the annual produced
tions will increase, and net electricity along with overall exergy effi­ and consumption electricities, generated hot water volume, energy and
ciency will both decrease slightly, as illustrated in Fig. 8. In opposition exergy efficiencies, and total product cost at the operating condition
to the trend of the performance indicators, i.e., the net electricity and over time are calculated implementing TRNSYS software.
exergy efficiency, the value of the economic index, i.e., the annual total Referring to Table 5, the PVTs generated electricity in the whole
product cost, increases from 14.9 €/MWh to 18.2 €/MWh. Moreover, year is 3647.4 kWh where 400.2 kWh and 47.1 kWh of it is, respec­
referring to the right side of Fig. 8(a), when the tank volume increases tively, consumed in the electrical coil of the tank and pumps to produce
from 0.01 m3 to 0.45 m3, the yearly generated hot water volume de­ hot water and run the system. Else, the extra generated electricity of
creases about 4 m3, which is negligible. 3200.1 kWh is directly utilized to supply the housing demand. The table
The tank loss coefficient is another influential parameter playing an also reveals that since the annual electricity sold to the network is equal
important role in the performance of not only the heat storage tank but to the annual electricity bought from it, the ultimate goal of ap­
also the overall integrated system. An efficiently insulated tank can proaching a true net-zero energy building on which the annual energy
retain heats for days, reduce energy consumption, speed up the heating bill of the building is just close to zero is achieved. Moreover, Table 5
process, and maintain the desired operating temperature. The effect of indicates that from 526.2 m3 of annual generated hot water, 123.4 m3 is
the average tank loss coefficient on the smart energy system perfor­ used to provide the domestic hot water need. Otherwise, when the solar
mance/economic features is depicted in Fig. 9. As illustrated in Fig. 9, energy is available, and simultaneously there isn’t any load demand, the
by increasing the heat storage tank loss coefficient from 0 to 5 kJ/ extra 402.8 m3 produced hot water is sold to the ultra-low temperature
(h.m2·K), the auxiliary heater electricity consumption increases; hence, local district heating grid to reduce the overall annual energy expenses.
the annual net electricity along with overall exergy efficiency will de­ The table further shows that under the Esbjerg weather data situation,
crease. Moreover, the yearly produced hot water increases from annual energy and exergy efficiencies of the proposed smart system are
529.2 m3 to 526.06 m3 as the loss coefficient increases from 0 to 3.5 kJ/ 61% and 45.04%, respectively. Eventually, the results of the economic
(h.m2·K). Subsequently, when the loss coefficient increases 3.5 kJ/ analysis indicate that the annual total product cost is equal to 16.9 €/h.
(h.m2·K) to 5 kJ/(h.m2·K), the volume decreases up to 526.5 m3. Ac­ The comparison between 16.9 €/MWh of the total product cost with 30
cording to Fig. 9(b), the yearly total product cost increases with an €/MWh and 21 €/MWh fees for electricity and hot water production,
increase in the heat storage tank loss coefficient. This is reasonable respectively, via CHP waste to the energy power plant as the conven­
because the increase in the tank loss coefficient means lower exergy of tional way in Denmark reveals the significance of the proposed smart
product is generated; hence, the total product cost of the system in­ system from the economic standpoint. Thus, the main superiority of this
creases, as shown in Fig. 9(b). system compared to its competitors is its simplicity, lower number of
equipment, and, thus, its lower costs, including maintenance costs.

9
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

Table 5 Table 7
The value of annual performance indicators and Logical domain for major decision variables.
economic indices of the proposed smart system.
Parameter Range
ĖPVT (kWh) 3647.4
Ėhetaer (kWh) 400.2 TPVT (°C) 40 < TPVT < 80
Ėpump (kWh) 47.1 NT 5 < NT < 32
Ėnet (kWh) 3200.1 LPVT (m) 0.5 < LPVT < 3.5
Ėload (kWh) 3392.7 VTank (m3) 0.01 < VTank < 0.45
ĖBought (kWh) 2083 LC (kJ/h·m2·K) 0 < LC < 5
ĖSold (kWh) 1937.5
QBought (kWh) 385.7
QSold (kWh) 7033.4
Vtotal (m3) 526.2
Vload (m3) 123.4 4.4. Results of optimization
Vgrid (m3) 402.8
ηPV (%) 14
ηI (%) 61 When simulation models are used to design a system, it is usually
ηII (%) 45.04 challenging to determine the parameter values that result in the best
CP,tot (€/MWh) 16.9 system performance. This is sometimes due to time constraints since it
is time-consuming for a user to change the input values and run the
simulation. Sometimes time is not a problem, but because of the com­
Table 6 plexity of the system, the user is just not capable of understanding the
Results of exergy analysis. nonlinear interactions of the various parameters. The goal of a single-
Component ηII (%) objective optimization problem is to find the best solution for a specific
ED (kWh) EF (kWh) EP (kWh)
objective function. The TRNSYS optimization library (also known as
Tempering valve 0 0.32 0.32 100 TRNOPT) is a tool that couples the TRNSYS simulation with the GenOpt
Heat storage tank 56.23 85.51 29.28 34.25 program for finding the optimum value of objective function along with
Pump 1 0.28 31.66 31.38 99.12
the value of major decision variables at optimal condition. The pro­
Tee piece 2 0 24.88 24.88 100
PVT panels 5937.88 17842.69 11904.81 66.72 posed smart system is optimized from thermodynamic and economic
Diverter 0.07 115.92 115.85 99.93 standpoints implementing a single-optimization method using the
Inverter 11.91 3659.32 3647.41 99.67 TRNOPT tool of TRNSYS. For this, the annual total exergy efficiency as
Tee piece 1 0.63 5.96 5.33 89.42 the performance indicator as well as the total product cost as the eco­
Pump 3 0.42 10.03 9.61 95.81
Heat exchanger 42.45 77.09 34.64 44.93
nomic index are maximized and minimized, respectively. What stands
Pump 2 0.72 5.41 4.69 86.69 out from the parametric study is that the performance of the smart
system is intensely influenced by the PVT outlet temperature, number
of PVT tubes, the length of the PVT, the volume of the heat storage
Also, to evaluate the performance of the proposed system from an en­ tank, and the average tank loss coefficient per unit area. Consequently,
vironmental point of view, the comparison of this system with the an­ to find the optimal operating condition, the rational domain for the
nual CO2 emission index of CHP waste to the energy power plant as the major decision variables are tabulated in Table 7.
conventional way in Denmark for electricity and heat production is Considering the yearly total product cost, to be minimized, and
added. Referring to the latest environmental report for Danish elec­ annual second law efficiency, to be maximized, as single-objective
tricity and CHP plant [45], the proposed solar system leads to a de­ functions, the value of major decision variables at the optimum op­
crease in CO2 emission of 220 kg/MWh, which is considerable. eration condition are listed in Table 8. As Table 8 indicates, to reach the
To assess the performance of the proposed smart system from the minimum total product cost, the optimal value of the PVT outlet tem­
quality of energy conversion facet, the yearly exergy destruction, ex­ perature, number of PVT tubes, the PVT length, the volume of the tank,
ergy of fuel, exergy of product, and exergy efficiency of each compo­ and the average tank loss coefficient are, respectively, 40 °C, 32, 3.5 m,
nent are listed in Table 6. By and large, it is established beyond doubt 0.043 m3, 0.12 kJ/hm2K. On the other hand, for maximum total exergy
that temperature difference, mixing, and chemical reaction are the most efficiency, the optimal values of the variables mentioned above are
prominent factors in deviation of the energy systems from an efficient equal to 40 °C, 28, 3.5 m, 0.45 m3, 0 kJ/hm2K, respectively. In com­
performance. In this regard, the measurement and comparison of ex­ parison to the annual total product cost and overall exergy efficiency of
ergy destruction for every component give a better insight to clarify the 16.9 €/MWh and 45.04% at operating conditions, the optimal values of
source of irreversibility. Table 6 reveals that PVT panels with the an­ 13.26 €/MWh and 49.07% reveal the significance of optimization of the
nual exergy destruction of 5937.88 kWh, which equals 98% of de­ proposed smart system.
struction, is the major source of inefficiency. Also, in comparison to the
exergy destruction of other components, the values of 56.23 kWh and
42.45 kWh associated to the heat storage tank and heat exchanger,
respectively, indicate their importance from the exergy standpoint. Table 8
Accordingly, since the temperature difference between the hot and cold Results of optimization.
sides of the heat storage tank is more than the differences in the heat
Annual total product cost as the objective function
exchanger, the tank has higher exergy destruction in a year.
Furthermore, tee piece 1 has the highest exergy destruction, 0.63 TPVT (°C) NT LPVT (m) VTank (m3) LC (kJ/h·m2·K) cP,tot (€/MWh)
kWh, among all of the valves due to the higher temperature difference
between entered and mixed waters. Moreover, the yearly exergy de­ 40 32 3.5 0.043 0.12 13.26
structions of pumps 2 and 3 are more than the destruction of pump1 as Yearly overall exergy efficiency as the objective function
a result of higher annual operation hours. Eventually, referring to
TPVT (°C) NT LPVT (m) VTank (m3) LC (kJ/h·m2·K) ηII (%)
Table 6, the majority of components have an exergy efficiency above
89%, indicating that the efficient equipment from the quality of energy 40 28 3.5 0.45 0 49.07
conversion aspect is implemented in the proposed system.

10
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

5. Conclusion applthermaleng.2019.04.075.
[2] Z. Liu, W. Li, Y. Chen, Y. Luo, L. Zhang, Review of energy conservation technologies
for fresh air supply in zero energy buildings, Appl. Therm. Eng. 148 (2019)
In this research paper, a smart building system integrated with PVT 544–556, https://doi.org/10.1016/j.applthermaleng.2018.11.085.
panels is proposed to not only provide the building’s real-time elec­ [3] H. Lund, P.A. Østergaard, D. Connolly, B.V. Mathiesen, Smart energy and smart
tricity demand but also to supply the annual domestic hot water need. energy systems, Energy 137 (2017) 556–565, https://doi.org/10.1016/J.ENERGY.
2017.05.123.
Unlike the conventional systems using batteries to store the electricity [4] A. Arabkoohsar, M. Sadi, Technical comparison of different solar-powered ab­
or also heat pump to boost heat production of the system, the proposed sorption chiller designs for co-supply of heat and cold networks, Energy Convers.
system only employs a heat storage tank to make a smart interaction Manag. 206 (2020) 112343, , https://doi.org/10.1016/j.enconman.2019.112343.
[5] A. Arabkoohsar, Non-uniform temperature district heating system with decen­
between the building and district heating and electricity networks. The tralized heat pumps and standalone storage tanks, Energy 170 (2019) 931–941,
aim is to decrease the cost of the system compared to other similar https://doi.org/10.1016/j.energy.2018.12.209.
solutions and thereby increase the motivation of costumers to adopt the [6] A. Arabkoohsar, A.S. Alsagri, A new generation of district heating system with
neighborhood-scale heat pumps and advanced pipes, a solution for future renew­
system in their buildings. In this regard, the sum of capital investment
able-based energy systems, Energy 193 (2020), https://doi.org/10.1016/j.energy.
and operating and maintenance costs, including fixed and variables 2019.116781.
costs, is evaluated, and a significant reduction in investment costs is [7] H. Lund, S. Werner, R. Wiltshire, S. Svendsen, J.E. Thorsen, F. Hvelplund, et al., 4th
achieved due to not using a battery. Then, the ratio of capital invest­ Generation District Heating (4GDH). Integrating smart thermal grids into future
sustainable energy systems, Energy 68 (2014) 1–11, https://doi.org/10.1016/j.
ment, and operating and maintenance cost to the annual produced energy.2014.02.089.
electricity and heat, named as total product unit cost, is evaluated and [8] X. Guan, Z. Xu QSJ, Energy efficient buildings facilitated by microgrid, IEEE Trans.
compared with the electricity and hot water production costs in Smart Grid 1 (2010) 243–252.
[9] F. Huide, Z. Xuxin, M. Lei, Z. Tao, W. Qixing, S. Hongyuan, A comparative study on
Denmark. Additionally, comprehensive energy and exergy assessment three types of solar utilization technologies for buildings: Photovoltaic, solar
are performed transiently under the hourly weather data information of thermal and hybrid photovoltaic/thermal systems, Energy Convers. Manag. 140
Esbjerg city (as a case study) using TRNSYS software. Besides, the in­ (2017) 1–13, https://doi.org/10.1016/j.enconman.2017.02.059.
[10] A. Behzadi, E. Gholamian, P. Ahmadi, A. Habibollahzade, M. Ashjaee, Energy, ex­
fluence of the PVT outlet temperature, the number of PVT tubes, the ergy and exergoeconomic (3E) analyses and multi-objective optimization of a solar
length of the PVT, the volume of the heat storage tank, and the average and geothermal based integrated energy system, Appl. Therm. Eng. 143 (2018)
tank loss coefficient per unit area on the proposed smart system is 1011–1022, https://doi.org/10.1016/j.applthermaleng.2018.08.034.
[11] Y. Khanjari, A.B. Kasaeian, F. Pourfayaz, Evaluating the environmental parameters
performed via parametric study. For this, the annual variation of net affecting the performance of photovoltaic thermal system using nanofluid, Appl.
electricity, a net produced hot water volume, exergy efficiency, and Therm. Eng. 115 (2017) 178–187, https://doi.org/10.1016/j.applthermaleng.2016.
total product cost with major decision variables is investigated. 12.104.
[12] C. Zhou, R. Liang, J. Zhang, A. Riaz, Experimental study on the cogeneration per­
Moreover, by calculating the exergy destruction, the performance of
formance of roll-bond-PVT heat pump system with single stage compression during
each component from the quality of energy conversion is evaluated and summer, Appl. Therm. Eng. 149 (2019) 249–261, https://doi.org/10.1016/j.
compared. Eventually, the proposed smart building system is optimized applthermaleng.2018.11.120.
using the TRNOPT tool of TRNSYS to minimize the yearly total product [13] A. Shahsavar, S. Khanmohammadi, Feasibility of a hybrid BIPV/T and thermal
wheel system for exhaust air heat recovery: Energy and exergy assessment and
cost and maximize the annual 2nd law efficiency. It is noted that the multi-objective optimization, Appl. Therm. Eng. 146 (2019) 104–122, https://doi.
proposed study does not have any geographical or architectural lim­ org/10.1016/j.applthermaleng.2018.09.101.
itations. Main conclusions of this study are: [14] A. Buonomano, F. Calise, G. Ferruzzi, L. Vanoli, A novel renewable polygeneration
system for hospital buildings: Design, simulation and thermo-economic optimiza­

• From annual PVT electricity of 3647.4 kWh, 400.2 kWh and 47.1
tion, Appl. Therm. Eng. 67 (2014) 43–60, https://doi.org/10.1016/j.
applthermaleng.2014.03.008.
kWh is consumed in the heat storage tank and pumps, respectively, [15] E. Entchev, L. Yang, M. Ghorab, E.J. Lee, Performance analysis of a hybrid re­
newable microgeneration system in load sharing applications, Appl. Therm. Eng. 71
to supply domestic hot water and the remaining electricity of 3200.1 (2014) 697–704, https://doi.org/10.1016/j.applthermaleng.2013.10.057.
kWh is used to provide the house real-time electricity demand. [16] M. Canelli, E. Entchev, M. Sasso, L. Yang, M. Ghorab, Dynamic simulations of hy­
• The proposed smart system not only supplies the entire annual do­ brid energy systems in load sharing application, Appl. Therm. Eng. 78 (2015)
315–325, https://doi.org/10.1016/j.applthermaleng.2014.12.061.
mestic hot water demand of the house, i.e., 123.4 m3, but also
[17] H. Hu, D. Yuan, T. Wang, Y. Jiang, Dynamic performance of high concentration
compensates the yearly energy costs by selling the surplus hot water photovoltaic/thermal system with air temperature and humidity regulation system
of 402.8 m3 to the local low-temperature district heating grid. (HCPVTH), Appl. Therm. Eng. 146 (2019) 577–587, https://doi.org/10.1016/j.

• PVT panels, heat storage tank, and heat exchanger with the annual applthermaleng.2018.10.028.
[18] J.F. Chen, L. Zhang, Y.J. Dai, Performance analysis and multi-objective optimiza­
exergy destruction of 5937.88 kWh, 56.23 kWh, 42.45 kWh are, tion of a hybrid photovoltaic/thermal collector for domestic hot water application,
respectively, the main source of irreversibility. Energy 143 (2018) 500–516, https://doi.org/10.1016/j.energy.2017.10.143.

• The annual total produced cost of 16.9 €/MWh in comparison to the [19] Z. Liu, Z. Liu, X. Yang, H. Zhai, X. Yang, Advanced exergy and exergoeconomic
analysis of a novel liquid carbon dioxide energy storage system, Energy Convers.
30 €/MWh and 21 €/MWh fees for electricity and hot water gen­ Manag. 205 (2020) 112391, , https://doi.org/10.1016/j.enconman.2019.112391.
eration, respectively, using conventional way in Denmark demon­ [20] R.S. Kamel, A.S. Fung, Modeling, simulation and feasibility analysis of residential
strates the importance of the present study from an economic point BIPV/T+ASHP system in cold climate - Canada, Energy Build. 82 (2014) 758–770,
https://doi.org/10.1016/j.enbuild.2014.07.081.
of view.

[21] M. Dannemand, B. Perers, S. Furbo, Performance of a demonstration solar PVT
What stands out from optimization is that in optimal conditions, assisted heat pump system with cold buffer storage and domestic hot water storage
while the maximum value for annual exergy efficiency is 49.07%, tanks, Energy Build. 188–189 (2019) 46–57, https://doi.org/10.1016/j.enbuild.
2018.12.042.
the minimum total product cost is 13.26 €/MWh. [22] M. Hazami, A. Riahi, F. Mehdaoui, O. Nouicer, A. Farhat, Energetic and exergetic
performances analysis of a PV/T (photovoltaic thermal) solar system tested and
Declaration of Competing Interest simulated under to Tunisian (North Africa) climatic conditions, Energy 107 (2016)
78–94, https://doi.org/10.1016/j.energy.2016.03.134.
[23] Z. Liu, B. Liu, J. Guo, X. Xin, X. Yang, Conventional and advanced exergy analysis of
The authors declare that they have no known competing financial a novel transcritical compressed carbon dioxide energy storage system, Energy
Convers. Manag. 198 (2019) 111807, , https://doi.org/10.1016/j.enconman.2019.
interests or personal relationships that could have appeared to influ­
111807.
ence the work reported in this paper. [24] M. Ahmed-Dahmane, A. Malek, T. Zitoun, Design and analysis of a BIPV/T system
with two applications controlled by an air handling unit, Energy Convers. Manag.
References 175 (2018) 49–66, https://doi.org/10.1016/j.enconman.2018.08.090.
[25] Z. Liu, W. Li, L. Zhang, Z. Wu, Y. Luo, Energy & Buildings Experimental study and
performance analysis of solar-driven exhaust air thermoelectric heat pump recovery
[1] G. Leroux, N. Le Pierrès, L. Stephan, E. Wurtz, J. Anger, L. Mora, Pilot-scale ex­ system, Energy Build. 186 (2019) 46–55, https://doi.org/10.1016/j.enbuild.2019.
perimental study of an innovative low-energy and low-cost cooling system for 01.017.
buildings, Appl. Therm. Eng. 157 (2019) 113665, , https://doi.org/10.1016/j. [26] M. Lagache, P. Ungerer, A. Boutin, A.H. Fuchs, Prediction of thermodynamic

11
A. Behzadi, et al. Applied Thermal Engineering 181 (2020) 115926

derivative properties of fluids by Monte Carlo simulation, Phys. Chem. Chem. Phys. [36] M. Shoaib, A. Khan, A. Waheed, T. Talha, M. Wajahat, F. Sarfraz, Con fi guration
3 (2001) 4333–4339, https://doi.org/10.1039/b104150a. based modeling and performance analysis of single e ff ect solar absorption cooling
[27] A. Moallemi, A. Arabkoohsar, F.J.P. Pujatti, R.M. Valle, K.A.R. Ismail, Non-uniform system in TRNSYS, Energy Convers. Manag. 157 (2018) 351–363, https://doi.org/
temperature district heating system with decentralized heat storage units, a reliable 10.1016/j.enconman.2017.12.024.
solution for heat supply, Energy 167 (2018) 80–91, https://doi.org/10.1016/j. [37] V. Zare, S.M.S. Mahmoudi, M. Yari, M. Amidpour, Thermoeconomic analysis and
energy.2018.10.188. optimization of an ammonia–water power/cooling cogeneration cycle, Energy 47
[28] M. Compton, B. Rezaie, Investigating steam turbine feasibility to improve the sus­ (2012) 271–283.
tainability of a biomass boiler using TRNSYS, Sustain Cities Soc. 43 (2018) 86–94, [38] S. Soltani, S.M.S. Mahmoudi, M. Yari, T. Morosuk, M.A. Rosen, V. Zare, A com­
https://doi.org/10.1016/j.scs.2018.08.032. parative exergoeconomic analysis of two biomass and co-firing combined power
[29] S.K. Wisconsin-Madison, E.E. Station, 1988 undefined. TRNSYS-A transient system plants, Energy Convers. Manag. 76 (2013) 83–91, https://doi.org/10.1016/j.
simulation program, CiNiiAcJp, n.d. enconman.2013.07.030.
[30] A. Behzadi, A. Habibollahzade, P. Ahmadi, E. Gholamian, E. Houshfar, Multi-ob­ [39] E. Indicators, Marshall&Swift Equipment Cost Index, Chem. Eng. 72 (2011).
jective design optimization of a solar based system for electricity, cooling, and [40] A. Behzadi, E. Gholamian, E. Houshfar, M. Ashjaee, A. Habibollahzade,
hydrogen production, Energy 169 (2019) 696–709, https://doi.org/10.1016/j. Thermoeconomic analysis of a hybrid PVT solar system integrated with double
energy.2018.12.047. effect absorption chiller for cooling/hydrogen production, Energy Equip. Syst. 6
[31] Y. Abbassi, E. Baniasadi, H. Ahmadikia, Comparative performance analysis of dif­ (2018).
ferent solar desiccant dehumidification systems, Energy Build. 150 (2017) 37–51, [41] M.G. Kim, G.S. Joe, T.H. Leigh, D.U. Shin, M.S. Yeo, K.W. Kim, Economic Feasibility
https://doi.org/10.1016/j.enbuild.2017.05.075. Considering Capacity of Thermal Storage Tank for Energy Balancing System, Energy
[32] A. Behzadi, A. Arabkoohsar, E. Gholamian, Multi-criteria optimization of a biomass- Procedia 78 (2015) 2082–2087, https://doi.org/10.1016/j.egypro.2015.11.237.
fired proton exchange membrane fuel cell integrated with organic rankine cycle/ [42] I. Dinçer, M. Rosen, A. Marc, P. Ahmadi, Optimization of energy systems, n.d.
thermoelectric generator using different gasification agents, Energy 201 (2020) [43] M. Yari, A.S. Mehr, S.M.S. Mahmoudi, M. Santarelli, A comparative study of two
117640, , https://doi.org/10.1016/j.energy.2020.117640. SOFC based cogeneration systems fed by municipal solid waste by means of either
[33] A.S. Klein, TRNSYS-A transient system simulation program. Univ Wisconsin- the gasifier or digester, Energy 114 (2016) 586–602, https://doi.org/10.1016/j.
Madison, Eng. Exp. Stn. Rep. (1988) 38–112. energy.2016.08.035.
[34] M. Sadi, A. Arabkoohsar, Exergoeconomic analysis of a combined solar-waste [44] K. Foteinaki, R. Li, C. Rode, R.K. Andersen, Energy & Buildings Modelling house­
driven power plant, Renew. Energy (2019) 883–893, https://doi.org/10.1016/j. hold electricity load profiles based on Danish time-use survey data, vol. 202, 2019.
renene.2019.04.070. https://doi.org/10.1016/j.enbuild.2019.109355.
[35] D. Jonas, M. Lämmle, D. Theis, S. Schneider, G. Frey, Performance modeling of PVT [45] Energinet, Environmental report 2018, Environ Rep Danish Electr CHP 2017 Statur
collectors : Implementation , validation and parameter identi fi cation approach Year, 2018, pp. 1–5. doi:16/19207-5.
using TRNSYS 193 (2019) 51–64. https://doi.org/10.1016/j.solener.2019.09.047.

12

You might also like