Next Article in Journal
Ultrasonic Deicing Efficiency Prediction and Validation for a Flat Deicing System
Next Article in Special Issue
Evolution of a Geothermal System with Fluid Re-Injection
Previous Article in Journal
SOH Estimation of Lithium-Ion Battery Pack Based on Integrated State Information from Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Integrated Geothermal Energy Systems for Small-Scale Combined Heat and Power Production: Energy and Economic Investigation

by
Pietropaolo Morrone
and
Angelo Algieri
*
Department of Mechanical, Energy and Management Engineering, University of Calabria, Via P. Bucci, 87036 Arcavacata di Rende (CS), Italy
*
Author to whom correspondence should be addressed.
Submission received: 26 August 2020 / Revised: 16 September 2020 / Accepted: 18 September 2020 / Published: 23 September 2020
(This article belongs to the Special Issue Recent Advances in Geothermal Energy Exploration)

Abstract

:

Featured Application

A geothermal integrated energy system for combined heat and power production in a commercial centre is proposed. The suggested apparatus can be adopted for several small-scale applications (e.g., industries, dwellings, smart communities) to improve their global energy, economic and environmental performance.

Abstract

In recent years, an increasing interest in geothermal energy has been registered in both the scientific community and industry. The present work aims to analyse the energy performance and the economic viability of an innovative high-efficiency geothermal-driven integrated system for a combined heat and power (CHP) application. The system consists of a heat exchanger (HEX) and a transcritical organic Rankine cycle (ORC) that work in parallel to exploit a high-temperature geothermal source (230 °C) and satisfy the energy demand of a commercial centre located in Southern Italy. The ORC and HEX sub-units can operate at partial load to increase the system flexibility and to properly react to continuous changes in energy request. A lumped model was developed to find the proper operating conditions and to evaluate the energy production on an hourly basis over the whole year. In particular, a multi-variable optimisation was implemented to find the most suitable configuration and a 101.4 kWel ORC was selected while the HEX nominal power was 249.5 kWth. The economic viability of the integrated system was evaluated in terms of net present value and payback period and different operating strategies were compared: thermal-driven, electric-driven, and a mixed strategy. The latter turned out to be the best solution according to both energy and economic criteria, with electric and thermal self-consumptions larger than 90%, with no heat dumping and a payback time close to five years.

1. Introduction

Combined heat and power (CHP) generation presents several advantages in terms of lower costs, emissions, fuel consumption, and higher decentralised generation share as compared to conventional separate electric and thermal production [1,2,3]. During the last few years, the research community and the industry have been focusing increased attention on renewable sources, sustainability and global efficiency [4,5,6]. In this framework, geothermal energy represents an interesting resource and, due to its not intermittent availability, an attractive solution for lowering fossil-fuel dependence and their environmental impact [7,8,9]. Furthermore, geothermal sources are slightly dependent on the season, climate and local conditions, presenting the advantages of different viable technologies [10] with direct (e.g., heat production for industrial and domestic users) and indirect applications (i.e., electricity generation at large, medium and small scales).
In geothermal applications, a geo-fluid transports the heat stored underground to the users through wells drilled into the geothermal reservoirs (up to 200–300 °C) [10]. The installed capacity of geothermal power has been increasing rapidly in the last few years [11] (from 6.8 GW in 1995 to an estimate of 21.4 GW for 2020 [12]). On the other hand, the direct use of geothermal sources is one of the first, most flexible and common methods to exploit geothermal potential [13].
A worldwide review of published data shows that the average electric conversion efficiency of geothermal systems is about 12%, while the highest values reach 21% [14]. For this purpose, it is worthy noting that several factors influence the system efficiency, including the conversion technology (i.e., dry steam, single, double and triple flash, binary and hybrid systems), the desired output (e.g., electricity production, combined heat and power generation, polygeneration), the size (large, medium, small, micro-scale), the ambient conditions, the temperature, the flow rate and the physic–chemical properties of the geothermal source [14].
In particular, several technologies can be adopted for geothermal power extraction, which depend on the thermodynamic state of the resource. At high temperatures, dry steam systems are usually adopted [12]. They represent a well-established technology and account for 23% of the global geothermal capacity. Furthermore, single and double flash steam power systems are the most diffused technology (42% for single flash and 19% for double flash) [12].
Finally, binary cycles represent the most recent solution that accounts for the remaining 16% of the total installed capacity in operation [15]. In this context, binary organic Ranking cycles (ORCs) play a crucial role [10,16,17,18]. ORCs represent an attractive and flexible technology for small applications owing to their capability to convert low enthalpy heat sources (i.e., waste heat [19,20,21,22,23] and renewable sources [24,25,26]) to electricity. Specifically, lower maintenance costs, better partial load performance, faster start-up and stop procedures, and higher safety and lifetime (up to 30 years) with respect to conventional installations are guaranteed [7,27]. Furthermore, a geothermal ORC does not need an intermediate thermal oil circuit, as usually present in other ORC applications [28,29] with a significant advantage in terms of performance, cost and architecture complexity [30]. The electric efficiency of the system can reach values up to 20% for the sole power generation [31]. Specifically, several factors affect the performance of ORC units: the thermal level of the external source, the working fluid, as well as the type of expander and operating conditions; therefore, the system architecture can vary significantly with the specific application and a standard setup is not completely established [11,32]. Moreover, the electric power of current geothermal systems is usually larger than 200 kWel with values also higher than 50 MWel [14,15,33,34]. On the other hand, few examples on small-scale ORCs are present, owing to higher specific costs and lower electric production [16,35]. As a consequence, further investigations are required to improve their performance [12,36]. Many scientific works concern geothermal ORC systems for electricity production. Ahmadi et al. recently provided a comprehensive review on this topic [7]. At the same time, the flexibility of geothermal-driven CHP systems based on ORCs has been receiving great attention from both the scientific community and industry [37,38,39,40,41]. In particular, a great amount of energy and techno-economic analysis and optimisation approaches are usually proposed for medium and large-scale geothermal power plants (electric power higher than 200 kWel) and for different system arrangements [37,42,43]. The objective functions combine energy parameters (i.e., first and second law analysis [44,45,46], electric power [41], heat exchanger areas [44,47], etc.) and economic purposes (e.g., specific cost of investment and levelized cost of energy, energy production cost [48,49,50]).
In this work, an innovative transcritical geothermal-driven ORC system for small-scale applications was investigated. To this purpose, the ORC unit is coupled with a traditional geothermal heat exchanger for the direct production of thermal energy. The sub-systems work in parallel to satisfy the electric and thermal demand of a small commercial centre. In a parallel arrangement, the heat is delivered to both the thermal user and the ORC at a high temperature and at a lower flow rate [43,51]. In a series design, the same geothermal flow passes through the ORC and subsequently through the heat exchanger [43,51]. As a consequence, the ORC unit significantly limits the thermal level of the heat generation in series arrangements [52,53]. According to the literature [43], the parallel configuration is the most suitable solution for supply and return temperatures higher or close to 80 and 60 °C [43], respectively, typical of space heating applications [7]. On the other hand, the series configuration is recommended for applications with lower temperatures (supply and return temperature close to 50 °C and 30 °C [25,30,43]) suitable for soil warming, swimming pools, de-icing, etc. [7]. Furthermore, the parallel configuration guarantees better ORC electric performance, owing to the lower condensation temperatures compared to series arrangements and higher flexibility. An innovative high-energy geothermal CHP demonstrator in parallel configuration is developing within the Fongeosec project [54,55]. Specifically, the geothermal source will be used to feed a 5.5 MWel subcritical ORC and a 15 MWth district heating network in Alsace (France) [56]. Parallel arrangements are also available in Bavaria (Germany) with Traunreut and Oberhaching geothermal plants. The global electric and thermal power is equal to 8.4 MWel and 52.0 MWth, respectively, while the net electric process efficiency ranges between 8.9% and 13.1% [57].
A lot of research effort is currently focused on organic Rankine cycles for CHP applications operating in subcritical conditions and driven by relatively low-temperature sources. The novelty of this work consists in the adoption of a transcritical ORC configuration with superheated conditions at the expander inlet. Transcritical systems show, in fact, higher performance compared to subcritical configurations [33,58] and provide a more efficient thermal exchange between the geofluid and organic fluid and lower irreversibility and exergy destruction [59]. Another benefit is the component downsizing due to the higher density of fluid. However, transcritical configurations present higher investment costs than subcritical arrangements owing to their higher operating pressures and larger size of some components (turbines, pumps and evaporators) [60,61]. Nevertheless, transcritical ORCs appear a promising technology to also exploit geothermal sources from an economic perspective [60,61,62]. Song et al. compared different ORC configurations (subcritical and transcritical, saturated and superheated, simple and regenerative cycles) with different organic fluids (isobutane, isopentane, R134a, R245fa, R1233zd, and R1234yf) for the exploitation of a 40 kg/s geothermal source at 180 °C [60]. A bi-variable optimisation was adopted based on the exergy efficiency maximisation and payback period minimisation. The comparison between optimised configurations demonstrated that transcritical systems present a lower payback period compared to subcritical arrangements for all the investigated working fluids, with decrease up to 90%. Furthermore, the authors demonstrated that when the payback period is fixed, the transcritical ORCs provide higher net power, larger long-term economic profit, and lower specific investment cost compared to subcritical systems. The advantages of transcritical configurations were also confirmed for geothermal source temperatures between 140 and 200 °C. Similar results were found by Preißinger et al., who observed a significant increase in the gross power and a noticeable decrease in the payback period comparing transcritical and subcritical power plants with geothermal source temperatures between 100 and 190 °C and different working fluids [61]. Furthermore, Astolfi et al. highlighted that transcritical systems guarantee lower specific costs with respect to subcritical ORCs when the ratio between the critical temperature of the organic fluid and the geothermal source thermal level ranges between 0.8 and 0.9 due to the higher energy production and efficiency that overcomes the larger investment costs. Furthermore, the investigation demonstrated that supercritical ORCs offer better energy and economic performance for high geothermal sources [62]. Nevertheless, to the best of the authors’ knowledge, few works are present in the literature concerning geothermal transcritical configurations.
The proposed unit is able to work at partial loads to fulfil the energy demand of the final users, to increase the system flexibility and to respond efficiently to load variations [30]. It is worthy to note that most of the papers present in the literature do not focus on off-design performance [63]. The internal regeneration has been considered to maximise the electric power and achieve higher efficiency with respect to the mean value of typical geothermal plants (close to 12%) and isobutane has been selected as working fluid owing to its good performance and low specific costs, according to the literature [64,65,66,67]. In particular, isobutane is considered one of the organic fluids with zero ODP (ozone depletion potential), suitable for supercritical ORC applications [33]. Even though the relatively high pressures lead to higher operating costs, the integrated system reaches higher global efficiency for high thermal level geothermal sources. Maloney et al. found that high pressure values are more appropriate for high expander inlet temperatures to maximise the system efficiency [68]. The authors show that only a few investigations are focused on working fluid temperatures larger than 180 °C. Despite this, in the range of temperatures 180–240 °C isobutane revealed good energy performance in transcritical ORCs. In particular, the authors showed that the best isobutane cycles are supercritical for the highest values of temperature of the geofluid investigated (180 °C and higher). Toffolo et al. confirm that isobutane represents one of the most effective fluids for the exploitation of high-temperature geothermal sources through a comprehensive screening of organic fluids [69] and similar results are found in the literature [70,71]. Furthermore, many researchers recommend isobutane as a suitable fluid for supercritical conditions from an energetic and exergetic point of view [72,73] and claim that it is a cost-efficient working fluid [74].
A lumped thermodynamic model has been developed to select the suitable ORC system and the corresponding operating conditions, and to calculate the energy production on hourly basis over the whole year. The proper system configuration has been defined adopting a multi-variable optimisation and different operating strategies (i.e., thermal-driven, electric-driven and mixed-mode) have been compared [2,75,76]. Finally, the economic viability of the proposed technical solution has been investigated through the estimation of the cost of the single sub-units. The payback period and net present value were compared for the different operating strategies.

2. Materials and Methods

Figure 1 highlights the scheme of the geothermal-driven system used to satisfy the energy requests of a commercial centre located in Southern Italy. In particular, the investigated area refers to the Phlegraean Fields where high temperature geothermal aquifers are present at a very shallow depth [77,78]. The system consists of an organic Rankine cycle (ORC) for the electric production and a heat exchanger (HEX) able to provide the thermal energy. The two components work in parallel and simultaneously and the total geothermal mass flow rate can be arranged between the ORC and HEX units according to the user energy demand. To this purpose, the geofluid is pumped to the two components from the geothermal reservoir and afterwards is reinjected into the ground.
The adopted parallel configuration guarantees high flexibility to the CHP system, owing to the possibility to split the geothermal mass flow rate in the ORC and HEX units, depending on the hourly user energy request [52]. Furthermore, the parallel arrangement permits one to meet the high temperature heat demand for the space heating and also when the ORC condensation temperature is set to low values to maximise the electric power and efficiency [43,52].
The integrated system was able to exchange the electric energy with the grid and a traditional natural gas packaged firetube boiler was adopted to meet the thermal load when the geothermal production was not sufficient.
In particular, the system can operate adopting a thermal-driven, an electric-driven or a mixed strategy. In the first two cases, the priority is given to the fulfilment of the heat and electricity request, respectively; as a consequence, the geothermal mass flow rate is divided between HEX and ORC sub-systems on the basis of the privileged load. In the mixed operation, the thermal-driven strategy is adopted when the heat demand is present in order to minimise the heat dumping. On the other hand, the geofluid is completely exploited in the ORC apparatus when the thermal request is absent to maximise the electric efficiency and production.
The global performance of the combined system is expressed in terms of electric efficiency (ηel), energy utilisation factor (EUF) and marginal efficiency of electric production (ηmar) [28,79], as follows:
η e l = P e l m ˙ g   ( h g ,   i ˙ n h g ,   o u t )
E U F = P   e l   +   P t h m ˙ g   ( h g , i n h g , o u t )
η m a r = P e l m ˙ g   ( h g ,   i ˙ n h g ,   o u t ) P t h η t h , r e f
where Pel is the ORC electric power, m ˙ g corresponds to the geothermal mass flow rate, hg,in and hg,out represent the enthalpy of the geothermal fluid at the entrance and exit of the integrated apparatus, respectively, and ηth,ref is thermal reference efficiency, based on the European harmonised efficiency reference value for separate heat production [80].

2.1. Organic Rankine Cycle (ORC) Unit

Transcritical organic Rankine cycles (ORCs) were selected for the investigation, owing to their high performance, low payback period and efficient heat exchange with low and medium-temperature heat sources [60,81,82,83,84]. In fact, organic fluids present usually low critical pressures and, as a consequence, transcritical configurations are very attractive, owing to the possibility to guarantee proper performance without operating into extreme and dangerous conditions [16].
The main components of an ORC are a pump, an evaporator, a turbine, and a condenser (Figure 2a) [85,86]. The pump increases the pressure of the organic working fluid to the maximum value (1-2 process) then the evaporator is adopted to obtain the superheated vapour (2-3) that expands in the turbine (3-4) to provide the mechanical work that is converted into electricity by an alternator. Finally, the working fluid is condensed (4-1). An internal heat exchanger (IHE) is often used to improve the system efficiency [87]. To this purpose, the energy content of the organic fluid at the exit of the turbine is exploited (4-4*) to preheat the liquid before the entrance in the evaporator (2-2*). Specifically, centrifugal pumps, radial turbine and shell and tube heat exchangers were selected, according to the literature [19,88,89,90]. Figure 2b depicts the different processes in the T-s diagram.
The performances of the geothermal-driven ORC were characterised by adopting a thermodynamic model developed by the authors in the last years [16,29,91]. The model is general and refers to the ORC unit alone (pump, evaporator, turbine, regenerator and condenser). It is capable of characterising the behaviour of ORC systems in both nominal and partial load operations independently of the energy source. The model includes an intermediate thermal oil circuit when high-temperature sources are adopted. Since this component is not necessary for geothermal application, the algorithm considers only a single heat exchanger that transfers the heat input from the geofluid to the organic fluid directly. The model integrates the Refprop database to define the working fluid properties [92], whereas steady state conditions were imposed [93] and pressure losses in the ORC components were neglected [94], in line with the literature.
The behaviour of the geothermal organic Rankine cycle system at full and part loads has been defined in terms of electric power and efficiency.
The electric power Pel is evaluated as:
P   e l = η e m , t   P t P p / η e m , p   P p , g e o / η e m , g e o     P p , c o o l / η e m , c o o l  
where Pt and Pp represent the ORC turbine and pump power, respectively, while ηem,t and ηem,p are the corresponding electro-mechanical efficiencies. Pp,g and Pp,cool denote the power request of the pumps used to drive the geofluid and the cooling water, whereas ηem,g and ηem,cool are the corresponding electro-mechanical efficiencies.
The electric efficiency of the ORC system ηel,ORC is:
η e l , O R C = P e l m ˙ g ,   O R C   ( h g , O R C   i ˙ n h g , O R C   o u t )
where m ˙ g ,   O R C is the mass flow rate of the geothermal source exploited in the ORC unit while hg,ORC out represents the enthalpy of the geofluid at the exit (out) of the ORC evaporator.
An iterative procedure has been implemented to take into account the variation of temperature and pressure at the turbine inlet when the system operates at partial loads, by using the Stodola’s ellipse approach. The algorithm evaluates the off-design variation of the turbine, pump and heat exchanger performance. More details on ORC partial load operations are available in a previous work of the authors [28]. The ORC model has been validated in previous works and more details are reported in the literature [16,28].

2.2. Heat Exchanger (HEX) Unit

A traditional geothermal heat exchanger (HEX) has been used to satisfy the thermal request of the commercial centre. To this purpose, a shell and tube unit in carbon steel has been considered, owing the high compatibility with geothermal and organic fluids [88]. The thermal power Pth is evaluated as:
P   t h = η t h , H E X   m ˙ g ,   H E X   ( h g , H E X   i ˙ n h g , H E X   o u t )  
where, ηth,HEX represents the HEX thermal efficiency, m ˙ g ,   H E X is the geothermal mass flow rate used in the heat exchanger, hg,HEX out is the enthalpy of the geofluid at the exit of the HEX sub-system. In particular, the geothermal mass flow rate for heating purposes m ˙ g ,   H E X is:
m ˙ g ,   H E X = m ˙ g     m ˙ g ,   O R C
The area of the heat exchanger has been calculated as:
A = P t h U   Δ T l m
U represents the global heat exchange coefficient and ΔTlm is the logarithmic mean temperature difference, defined as follows:
Δ T l m = ( T h , o u t T c , i n ) ( T h , i n T c , o u t ) l n ( T h , o u t T c , i n ) ( T h , i n T c , o u t )
where T is the temperature, the subscripts h and c refer to the hot and cold fluid, respectively, while the subscripts in and out correspond to the inlet and outlet fluid sections.
The global heat transfer coefficient is:
U = 1 d o h i   d i + R f , i   d o d i + d o 2   k     l n ( d o / d i ) + R f , o + 1 h o
where D is the tube diameter, H corresponds to the heat transfer coefficient, K is the thermal conductivity and Rf is the fouling factor. The subscripts i and o refer to the properties inside and outside the tubes. In particular, the Gnielinski equation [95] and the Kern method [96] were adopted to define Hi and Ho, respectively [88].
The same procedure was implemented to define the area of the other heat exchangers of the ORC sub-system (i.e., condenser, evaporator and internal regenerator). In particular, for the sole evaporator, the Krasnoshchekov et al. correlation was adopted to characterise the heat transfer coefficient in supercritical conditions [97,98]. The correlation is based on the Krasnoshchekov and Protopopov rule and the Petukhov and Kirillov equation [98,99]. To this purpose, it is worthy to note that a universal accepted law for the heat transfer coefficient in supercritical regimes is still not available [98,100,101]; however, the adopted rules are among the most accurate correlations available in the literature for supercritical fluids [98,102,103,104].

2.3. Operating Conditions

The performances of the integrated energy systems have been investigated. To this purpose, 1 kg/s of geothermal water at 230 °C was considered as energy input [16]. The geofluid was used to drive both the ORC and HEX units depending on the energy request of the investigated users while the minimum reinjection temperature was fixed to 70 °C to prevent scaling and fouling problems within system components and pipes, according to the literature [82,105,106]. The thermal efficiency of the heating process from the geothermal source was fixed to 0.92 [80] and the minimum pinch-point temperature was considered equal to 10 °C [16]. This value was increased if the constraint on the minimum reinjection temperature was not fulfilled.
The analysis was carried out adopting transcritical ORC units with and without the internal heat regenerator (simple and IHE configuration, respectively) owing to their high performance with geothermal sources [16]. Isobutane was selected as organic fluid due to its thermal stability, low specific cost and attractive efficiency, according to the literature [67,107,108]. Furthermore, isobutane is suitable for supercritical ORC applications with high-temperature geothermal sources [33,68,69], as already pointed out in the introduction section. The ratio between the critical temperature of the working fluid and the geothermal source temperature was equal to 0.81, within the suggested interval (0.8–0.9) to minimise the specific investment cost and the levelized cost of energy [62]. ORC condensation temperature was fixed to 30 °C while the maximum pressure was set to 65.32 bar, corresponding to 1.80 times the critical value (pcrit = 36.29 bar). At part load operations, the maximum pressure can be reduced up to 1.03 pcrit depending on the energy request.
The temperature at the entrance of the turbine varies with the pressure and the maximum value was fixed to 220 °C, based on the geothermal source temperature.
For the investigation, the maximum turbine and pump efficiencies were imposed equal to 0.80 and 0.70 [109], respectively, while the electro-mechanical efficiencies were 0.95 [110]. The effectiveness of the internal heat exchanger was set to 95% and the vapour temperature at the IHE exit (point 4*) was 10 °C higher than the condensation temperature [16,111]. The efficiency and head of the pump used for the cooling circuit were equal to 80% and 10 m, respectively, while the corresponding pinch point temperature was 5 °C [81,106]. Table 1 summarises the operating conditions and the main assumptions adopted for the analysis.

2.4. Electric and Thermal Demand

The integrated geothermal system was adopted to satisfy the energy request of a commercial centre in Southern Italy [112]. The yearly electric and thermal demands were equal to 533.1 MWhel and 396.3 MWhth, respectively. The electric load depended mainly on the lighting system and heating, ventilation, and air conditioning (HVAC) units, while the thermal demand was based on space heating and hot water request. The analysis was performed on an hourly basis and the electric and thermal loads during the year are depicted in Figure 3. The highest electric request (144.0 kWhel) was registered in August when the thermal demand was absent or presented the minimum values. As expected, the maximum heat load (427.9 kWhth) was noticed in winter, owing to the space heating demand.

2.5. Economic Analysis

The economic viability of the integrated geothermal system has been evaluated. To this purpose, the costs of the different components were calculated, adopting the bare module equipment methodology, which takes into account direct and indirect costs [113]. In particular, the reference purchased cost C j 0 (in US dollar) of the generic component j was evaluated according to the following equation:
log C j 0 = K 1 , j + K 2 , j log S j + K 3 , j ( log S j ) 2
where S is the size or capacity parameter of the component j and it corresponds to the area (in m2) of heat exchangers or to the power (in kW) of pumps and turbines [113]. The superscript 0 refers to the base conditions with carbon steel material and ambient operating pressure, whereas the coefficients k1,j, k2,j and k3,j vary with the selected sub-systems (Table 2).
The reference purchased costs of pumps and heat exchangers are corrected to take into account the adopted construction materials and operating pressures:
C j = C j 0     ( B 1 , j + B 2 , j   F M , j     F P , j )
where the coefficients b1,j and b2,j depend on the equipment type, while FM,j and FP,j represent the material and the pressure factors of the generic component j. Specifically, the pressure factors are evaluated according to:
log F P , j = C 1 , j + C 2 , j log p j + C 3 , j ( log p j ) 2
where pj is the operating pressure. The coefficients b1,j, b2,j, c1,j, c2,j, c3,j and the material factors FM,j are shown in Table 2 [19,88,89,90].
The turbine purchased cost was evaluated as a function of the corresponding reference cost [60,114,115]:
C j = C j 0     F B M , j
where the corresponding bare module factor FBM is in Table 2 [19,88,89,90].
Finally, the total cost (in euro) of the integrated geothermal system was calculated according to the following equation:
C = C E P C I 2019 C E P C I 2001     e E U R / U S D     j = 1 N C j
Here, the chemical engineering plant cost indexes (CEPCI) are used to account for inflation and update to 2019 the parameters of Table 2 (referred to as 2001). Specifically, CEPCI indexes correspond to 397.0 and 607.5 for 2001 and 2019, respectively [113,116,117]. Furthermore, the European Central Bank reference exchange rate (eEUR/USD) from euro (EUR) to US dollar (USD) is adopted to express the total cost of the geothermal combined system in euros [118].
Other assumptions for the economic analysis are summarised in Table 3. In particular, the investment period has been set to 20 years and the interest rate to 2%. The electricity and natural gas prices are based on Italian tariffs for non-domestic users [119] and the maintenance costs of the integrated geothermal-driven system have been assumed to be equal to 2% of the initial investment [28]. The reference exchange rate from euro to US dollar is 1.12 €/$ (1 July 2020 value) [118].

3. Results and Discussion

The energy performance of a geothermal integrated system for combined heat and power (CHP) production was modelled and analysed. The geothermal source drives a transcritical organic Rankine cycle (ORC) unit and a heat exchanger (HEX) to satisfy the electric and thermal demand of a small commercial centre [112]. The available mass flow rate and thermal level of the geothermal source were fixed to 1 kg/s and 230 °C, respectively. Full and partial load operations were considered and the annual energy balances were evaluated on an hourly basis.

3.1. Performance of the Integrated Geothermal System

First, a parametric investigation was carried out, and the effect of the maximum temperature (Tmax) on the ORC performance was analysed. The maximum pressure was fixed to 65.3 bar (1.80 pcrit) while the condensation temperature was set to 30 °C. As expected [29], a continuous increase in the ORC electric efficiency with the maximum temperature was noticed for both simple and regenerative configurations (Figure 4). The simple arrangement guarantees higher performance for temperature values lower than 165 °C and a plateau (ηel ≈ 13.4%) was reached for Tmax higher than 200 °C. Conversely, the IHE configuration guarantees a significant raise in the electric efficiency when the temperature moves from 140 °C to 220 °C (6.7% to 18.7%, respectively) while the maximum electric power (102.3 kWel) is registered when the temperature at the turbine entrance is equal to 195 °C.
A multi-variable optimisation was adopted to select the most suitable configuration of the ORC geothermal-driven apparatus, and the electric power and efficiency have been considered as objective parameters. Figure 5 highlights that a Pareto frontier exists and the “minimum distance” criterion was adopted to define the proper ORC arrangement according to the literature [120,121]. The method suggests selecting the system configuration that minimises the dimensionless distance to the ideal point (the red circle in Figure 5) that is characterised by the maximum ORC electric efficiency (ηel,ORCmax) and the maximum electric power (Pel,max), according to the following equation:
d = min ( ( η e l , O R C   max η e l , O R C     i η e l , O R C   max η e l , O R C   min ) 2 + ( P e l ,   max P e l ,     i P e l ,   max P e l ,   min ) 2 )
The subscript k refers to the generic kth ORC configuration that is characterised by the electric efficiency ηel,ORC k and the electric power Pel,k, whereas ηel,ORC min and Pel,min corresponds to the minimum ORC electric efficiency and power on the Pareto frontier, respectively.
In particular, the transcritical ORC with the internal heat exchanger (IHE) and temperature at the entrance of the turbine equal to 205 °C was selected as the most suitable configuration. The corresponding electric power and efficiency were equal to 101.4 kW and 17.4%, respectively. At full load, the geothermal mass flow rate (1 kg/s) was fully exploited in the ORC and the HEX thermal production is absent. At partial load the mass flow rate of the organic fluid was reduced and, as a consequence, a rate of the geofluid was used to drive the HEX sub-system. It is worth noting that a minimum geothermal flow is required for the proper operation of the HEX heat exchanger. For this reason, the integrated CHP system only produces electricity when the electric load is larger than 91.0%, corresponding to the purple area in Figure 6.
Figure 6a highlights the geothermal flow used in the ORC, the electric and thermal power of the integrated CHP system as a function of the electric load. Furthermore, the energy utilisation factor, the marginal efficiency, and the electric efficiency are plotted in Figure 6b. In particular, the minimum electric load is equal to 50%, owing to the constraints of the minimum pressure at the ORC turbine inlet (1.03 pcrit). The corresponding electric and thermal power are equal to 50.7 kWel and 249.5 kWth, whereas electric efficiency is 8.7%. A slight variation in the marginal efficiency is observed (from 16.0% to 17.7%) whereas EUF is more sensitive to the load and ranges between 25.2% and 50.9% (Figure 6b). The geothermal temperature at the exit of the evaporator ranges between 95.9 °C (at full load) and 70.0 °C (at minimum). For this purpose, it is worth noting that the adopted parallel arrangement permits to satisfy the high-temperature heat request for all the operating conditions.

3.2. Influence of the Operating Strategy

The yearly electric and thermal productions of the geothermal-driven CHP system were evaluated on an hourly basis considering the electric and thermal request of a commercial centre [112] sited in the Phlegraean Fields area where geothermal sources at a temperature higher than 200 °C are present [77,78]. The energy system can operate according to a thermal-driven, electric-driven, or mixed strategy. Depending on the energy request of the final user, a variable operation is registered all over the year.
Figure 7 shows the comparison between the three operating strategies through the monthly energy balances of the CHP system. The electric demand presents a regular trend all over the year (around 44 MWh/month), while the thermal request variation is significant, ranging from 3.7 to 75.4 MWh/month (in August and January, respectively).
The thermal-driven and mixed strategies show a similar behaviour. Both operating modes guarantee no thermal dumping and a significant self-consumption during the winter period, ranging from 89.0% to 100.0%. Furthermore, the thermal-driven strategy satisfies 100% of the user heat request during the summer, whereas mixed-mode ranges from 64.8% to 88.9% in the same months. On the contrary, the mixed strategy significantly increases the electric self-consumption to values always higher than 87.6%, while thermal strategy never satisfies the electric request (the maximum self-consumption is 69.1%). In this case, a significant integration is necessary during summertime (up to 90% of the electric load).
The grid integration is present, although very small, when the mixed operation is adopted, despite the electric production always being larger than the energy demand. In fact, the electric and thermal loads are not simultaneous during the 24 h.
The electric-driven operation produces very high electric self-consumption and moderate excess to be injected into the grid. Conversely, noticeable thermal dumping is registered. In fact, the electric-driven strategy forces a homogenous operation of the apparatus, with a mean electric load equal to 68.1% for all the year, owing to the continuous electric request. The system operates with intermediate electric performance and significant thermal production, as well as when the thermal demand is absent.
On the contrary, the thermal-driven strategy guarantees a higher mean load (80.6%) with fewer operating hours (3692 h/year) because the thermal request is limited to the central hours of the day. This circumstance corresponds to lower equivalent operation hours (2977) as compared with the electric mode (6028). It is worthy to note that the mixed strategy reduces significantly the thermal dumping producing only electric energy at full load when the thermal request is absent. At the same time, the system operates in the high-performance range with 92.4% mean load and 8095 equivalent hours of operation. A slight excess of electric energy is noticed, due to the difference in the production and request timing.
Figure 8 shows the annual balance for the three operating modes. The mixed strategy satisfies both thermal and electric demand with small integration (4.7% and 9.3% of the thermal and electric load, respectively) with 59.2% electric surplus and no thermal dumping. If the heat-driven mode is used, the integrated CHP apparatus is able to satisfy up to 95.7% of the thermal demand, while the electric self-consumption drops to 47.5%.
Furthermore, the economic viability of the integrated energy system was analysed and the different operating strategies were compared. To this purpose, the investment period was fixed to 20 years and the main assumptions for the economic investigation are summarised in Table 3. The cost of the integrated CHP system (ORC and HEX sub-systems) was evaluated equal to EUR 636,786 adopting the bare model cost [113]. Specifically, the initial investment for the selected supercritical ORC unit was evaluated equal to EUR 551,838, with a specific cost lower than 5500 €/kWel, while the cost of the HEX sub-system corresponds to about EUR 85,000. The absolute and percentage costs of the different components are visible in Figure 9. The turbine revealed the highest rate (36.5% of the integrated system), whereas the share costs of the four heat exchangers were similar and ranged between 13.3% of the HEX unit and 18.0% of the condenser. The pumps costs correspond to about 4%.
The analysis reveals that the operating strategy has a noticeable impact on the economic performance of the integrated system when the same architecture is selected. To this purpose, the net present values (NPV) for the three operating modes are compared in Figure 10.
The results highlight that the mixed strategy also guarantees better economic performance: the payback period is lower than five years (4.9 years) and the NPV at the end of the investment period is equal to EUR 1,617,456. A slight increase in the payback time (PBT) is registered when the electric-driven mode is adopted (6.1 years). On the other hand, the thermal-driven strategy is characterised by a significant PBT rise to 11.0 years, and the NPV reduces to less than EUR 430,000, owing to the noticeable electric integration from the grid (52.7%).

3.3. Hourly Energy Balance for the Selected Integrated System

The previous analysis demonstrates that the mixed operating strategy guarantees the best techno-economic performance for the investigated integrated energy system. To evaluate in more detail the behaviour of the geothermal-driven CHP apparatus, Figure 11 depicts the distribution of hourly global electric efficiency and EUF index adopting mixed operations. The electric effectiveness ranges between 7.7% and 16.3%. As already observed, the maximum electric efficiency refers to full load operation with the sole electric production (Figure 6). It is worth noting that a concentration of high electric and low EUF points is registered during the hot season due to the low thermal request during this period. In this case, most of the geothermal source is used for electric production at high electric load, with good electric efficiency and low thermal performance. Furthermore, in order to guarantee the complete exploitation of the geothermal source, the integrated unit produces the sole electric energy at full load when the thermal request is absent.
Figure 12 shows the hourly energy balances in typical winter and spring days (3 January and 3 April). The CHP unit is capable of fulfilling the thermal load almost completely in January and a negligible integration is needed at 7:00 h (0.5% of the daily energy demand), whereas no thermal surplus is registered. A slightly higher integration is necessary in April (3.7% of the daily demand) due to the thermal request at 15:00 h, lower than the minimum thermal power of the combined apparatus. As far as the electric production is concerned, a very high self-consumption is guaranteed for both reference days (89.3% and 96.2% in January and April, respectively) and integration is required during the daytime, when the thermal request is present. Conversely, when there is no thermal demand, the whole geothermal source is used to produce electric energy. In this case, a higher electric surplus to be injected to the grid is noticed (39.1% in January and 47.6% in April).
The thermal demand decreases noticeably during the hot season (Figure 13) and it is present for a few hours per day. The thermal self-consumption is always higher than 94% and the system works at full load for 21 h in July, satisfying almost completely the daily electric request (values higher than 99% are found). The electric energy injected to the grid is 62.6% and 49.2% of the daily production in July and October, respectively.
Table 4 summarises the yearly techno-economic performance of the integrated geothermal CHP system with mixed operation strategy. The global electric efficiency is equal to 13.4% and confirms that the parallel transcritical configuration guarantees high performance in line with medium and large-scale geothermal CHP systems [57]. The energy utilisation factor reaches 19.2% while the marginal efficiency of electric production is 14.4%.
As already observed, the thermal surplus is absent, whereas the electricity injected into the grid is about 315 MWhel. The natural gas consumption reduces to 4512.6 Sm3 and corresponds to 9.8% of the quantity that should be used to satisfy the total thermal request of the commercial centre, adopting a traditional gas-fired boiler.
The significant fossil fuel saving, and the noticeable rate of electric self-consumption, have a positive influence on the environmental impact of the integrated energy system. For this purpose, the standard and LCA (life cycle assessment) methodologies have been used [122,123] to evaluate the carbon dioxide (CO2) and greenhouse gas (GHG) emissions, expressed as tons of CO2 (tCO2) and tons of CO2 equivalent (tCO2,eq), respectively. According to the standard approach, the emission factor for the electricity generation in Italy is equal to 0.343 tCO2/MWh based on the national energy mix with both renewable and non-renewable sources, whereas the emission factor for the natural gas combustion is 0.202 tCO2/MWh. The corresponding factors are equal to 0.424 tCO2,eq/MWh and 0.240 tCO2,eq/MWh when the LCA method is considered [122,123].
The investigation highlights that geothermal exploitation in the suggested CHP system assures an annual decrease in the CO2 and GHG emissions larger than 250 tCO2 and 300 tCO2,eq, respectively, compared to the traditional methods to fulfil the user’s energy request (electricity from the grid and thermal energy from a gas-fired boiler) and demonstrates that the integrated apparatus with the mixed operating strategy guarantees proper energy, economic and environmental performance. The proposed CHP geothermal system, developed for a commercial centre, can be easily extended and adapted to other applications (e.g., industries, dwellings, smart communities) owing to the high flexibility and the possibility to adopt different operating strategies depending on the user’s energy requests. In this way, the distributed exploitation of renewable sources and local small-scale energy production can drive a faster and softer transition from fossil fuels and traditional technologies towards cleaner and more sustainable energy solutions.

4. Conclusions

A techno-economic analysis was carried out to investigate the energy performance and the viability of an innovative CHP system, based on a transcritical geothermal-driven ORC unit and a heat exchanger (HEX) for the exploitation of a high-temperature geothermal source (230 °C). The sub-systems operate in parallel to satisfy the energy request of a commercial centre. Depending on the energy demand, the geofluid drives both the ORC and HEX sub-units: at full electric load, the geothermal water drives the sole ORC and no thermal production is present, whereas at partial load a fraction of the geofluid feeds the HEX system.
A thermodynamic model was developed to predict the performance of the geothermal-driven integrated system at full and part loads. First of all, the influence of the maximum temperature on the ORC behaviour was evaluated through a parametric investigation for both simple and regenerative ORC arrangements and the proper system configuration was defined by adopting a multi-variable optimisation. To this purpose, electric power and efficiency were considered as objective parameters and the “minimum distance” method was used to find the most suitable apparatus. In particular, the optimised system consists of a regenerative 101.4 kW kWel ORC unit with 17.4% electric efficiency and a geothermal heat exchanger that guarantees 249.5 kWth.
As a test case, the application of the integrated CHP system to a commercial centre located in Phlegraean Fields (Southern Italy) was considered and the payback time was used to evaluate the economic viability. To this aim, the costs of the different components were estimated through the bare module cost methodology and three operating strategies were compared: electric- and thermal-driven and mixed strategies. For each operating mode, the yearly electric and thermal energy balances were evaluated on an hourly basis. The results highlight that mixed-mode represents the best solution according to both energy and economic points of view. Particularly, the integrated system provides an electric and thermal self-consumption higher than 90%, while thermal and electric strategies revealed lower performance (electric self-consumption lower than 50% for the thermal-driven mode and thermal self-consumption less than 70% for the electric strategy). Furthermore, the mixed strategy assures no heat dumping and the highest equivalent hours of operation—i.e., 8095—whereas thermal- and electric-driven strategies show 2977 and 6028 h, respectively. Finally, the payback time lower than five years confirms that the proposed system represents a viable technical solution for small-scale and combined energy production, able to give a contribution in the transition towards cleaner and more sustainable technologies. Specifically, the geothermal exploitation in the suggested integrated apparatus provides a yearly CO2 emission reduction of more than 250 tons and a greenhouse gas saving of more than 300 tons of CO2 equivalent per year.

Author Contributions

Conceptualization, A.A. and P.M.; methodology, A.A. and P.M.; software, A.A. and P.M.; validation, A.A. and P.M.; formal analysis, A.A. and P.M.; investigation, A.A. and P.M.; resources, A.A. and P.M.; data curation, A.A. and P.M.; writing—original draft preparation, A.A. and P.M.; visualization, A.A. and P.M.; supervision, A.A. and P.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

AArea of heat exchanger (m2)
bBare module coefficient (-)
CCost (€)
cBare module coefficient (-)
DTube diameter (m)
dMinimum dimensionless distance (-)
eEUR/USDEuro/US dollar exchange rate (€/$)
FMBare module factor (-)
FMMaterial factor (-)
FpPressure factor (-)
HHeat transfer coefficient (W/m2 K)
hEnthalpy (J/kg)
KThermal conductivity (W/m K)
kBare module coefficient (-)
m ˙ Mass flow rate (kg/s)
PPower (W)
pPressure (Pa)
RfFouling factor (m2 K/W)
SSize parameter (m2 or kW)
sEntropy (J/kg K)
TTemperature (°C)
UGlobal heat exchange coefficient (W/m2 K)
Greek characters
ΔDifference
ηEfficiency
Subscripts and superscripts
cCold
coolCooling system
elElectric
emElectro-mechanical
eqEquivalent
gGeothermal
hHot
iInside
inInlet
jGeneric component
kGeneric configuration
lmLogarithmic mean
marMarginal
maxMaximum
minMinimum
oOutside
outOutlet
pPump
refReference
tTurbine
thThermal
0Base conditions
Acronyms
CEPCIChemical engineering plant cost index
CHPCombined heat and power
EUFEnergy utilisation factor
GHGGreenhouse gas
HEXHeat exchanger
HVACHeating, ventilation and air conditioning
IHEInternal heat exchanger
LCALife cycle assessment
ODPOzone depletion potential
ORCOrganic Rankine cycle

References

  1. Chen, X.; Si, Y.; Liu, C.; Chen, L.; Xue, X.; Guo, Y.; Mei, S. The Value and Optimal Sizes of Energy Storage Units in Solar-Assist Cogeneration Energy Hubs. Appl. Sci. 2020, 10, 4994. [Google Scholar] [CrossRef]
  2. Amber, K.; Day, T.; Ratyal, N.; Kiani, A.; Ahmad, R. Techno, Economic and Environmental Assessment of a Combined Heat and Power (CHP) System—A Case Study for a University Campus. Energies 2018, 11, 1133. [Google Scholar] [CrossRef] [Green Version]
  3. Li, D.; Xu, X.; Yu, D.; Dong, M.; Liu, H. Rule Based Coordinated Control of Domestic Combined Micro-CHP and Energy Storage System for Optimal Daily Cost. Appl. Sci. 2017, 8, 8. [Google Scholar] [CrossRef] [Green Version]
  4. Carlos de Oliveira Matias, J.; Godina, R.; Pouresmaeil, E. Sustainable Energy Systems: Optimization and Efficiency. Appl. Sci. 2020, 10, 4405. [Google Scholar] [CrossRef]
  5. Amelio, M.; Morrone, P. Residential Cogeneration and Trigeneration. In Current Trends and Future Developments on (Bio-) Membrane; Elsevier: Amsterdam, The Netherlands, 2020; pp. 141–175. ISBN 978-0-12-817807-2. [Google Scholar]
  6. Algieri, A.; Andiloro, S.; Tamburino, V.; Zema, D.A. The potential of agricultural residues for energy production in Calabria (Southern Italy). Renew. Sustain. Energy Rev. 2019, 104, 1–14. [Google Scholar] [CrossRef]
  7. Ahmadi, A.; El Haj Assad, M.; Jamali, D.H.; Kumar, R.; Li, Z.X.; Salameh, T.; Al-Shabi, M.; Ehyaei, M.A. Applications of geothermal organic Rankine Cycle for electricity production. J. Clean. Prod. 2020, 274, 122950. [Google Scholar] [CrossRef]
  8. Jocić, N.; Müller, J.; Požar, T.; Bertermann, D. Renewable Energy Sources in a Post-Socialist Transitional Environment: The Influence of Social Geographic Factors on Potential Utilization of Very Shallow Geothermal Energy within Heating Systems in Small Serbian Town of Ub. Appl. Sci. 2020, 10, 2739. [Google Scholar] [CrossRef] [Green Version]
  9. Yu, G.; Yu, Z. Research on a Coupled Total-Flow and Single-Flash (TF-SF) System for Power and Freshwater Generation from Geothermal Source. Appl. Sci. 2020, 10, 2689. [Google Scholar] [CrossRef] [Green Version]
  10. Moya, D.; Aldás, C.; Kaparaju, P. Geothermal energy: Power plant technology and direct heat applications. Renew. Sustain. Energy Rev. 2018, 94, 889–901. [Google Scholar] [CrossRef]
  11. Bianchi, M.; Branchini, L.; Pascale, A.D.; Melino, F.; Ottaviano, S.; Peretto, A.; Torricelli, N.; Zampieri, G. Performance and operation of micro-ORC energy system using geothermal heat source. Energy Procedia 2018, 148, 384–391. [Google Scholar] [CrossRef]
  12. Anderson, A.; Rezaie, B. Geothermal technology: Trends and potential role in a sustainable future. Appl. Energy 2019, 248, 18–34. [Google Scholar] [CrossRef]
  13. Lund, J.W.; Boyd, T.L. Direct utilization of geothermal energy 2015 worldwide review. Geothermics 2016, 60, 66–93. [Google Scholar] [CrossRef]
  14. Zarrouk, S.J.; Moon, H. Efficiency of geothermal power plants: A worldwide review. Geothermics 2014, 51, 142–153. [Google Scholar] [CrossRef]
  15. Bertani, R. Geothermal power generation in the world 2010–2014 update report. Geothermics 2016, 60, 31–43. [Google Scholar] [CrossRef]
  16. Algieri, A. Energy Exploitation of High-Temperature Geothermal Sources in Volcanic Areas—A Possible ORC Application in Phlegraean Fields (Southern Italy). Energies 2018, 11, 618. [Google Scholar] [CrossRef] [Green Version]
  17. Nami, H.; Anvari-Moghaddam, A. Geothermal driven micro-CCHP for domestic application—Exergy, economic and sustainability analysis. Energy 2020, 207, 118195. [Google Scholar] [CrossRef]
  18. Ehyaei, M.A.; Ahmadi, A.; El Haj Assad, M.; Rosen, M.A. Investigation of an integrated system combining an Organic Rankine Cycle and absorption chiller driven by geothermal energy: Energy, exergy, and economic analyses and optimization. J. Clean. Prod. 2020, 258, 120780. [Google Scholar] [CrossRef]
  19. Li, P.; Han, Z.; Jia, X.; Mei, Z.; Han, X.; Wang, Z. An Improved Analysis Method for Organic Rankine Cycles Based on Radial-Inflow Turbine Efficiency Prediction. Appl. Sci. 2018, 9, 49. [Google Scholar] [CrossRef] [Green Version]
  20. Gutierrez, J.C.; Valencia Ochoa, G.; Duarte-Forero, J. Regenerative Organic Rankine Cycle as Bottoming Cycle of an Industrial Gas Engine: Traditional and Advanced Exergetic Analysis. Appl. Sci. 2020, 10, 4411. [Google Scholar] [CrossRef]
  21. Alshammari, F.; Karvountzis-Kontakiotis, A.; Pesiridis, A.; Minton, T. Radial Expander Design for an Engine Organic Rankine Cycle Waste Heat Recovery System. Energy Procedia 2017, 129, 285–292. [Google Scholar] [CrossRef]
  22. Bellos, E.; Tzivanidis, C. Investigation of a hybrid ORC driven by waste heat and solar energy. Energy Convers. Manag. 2018, 156, 427–439. [Google Scholar] [CrossRef]
  23. Peris, B.; Navarro-Esbrí, J.; Molés, F.; Martí, J.P.; Mota-Babiloni, A. Experimental characterization of an Organic Rankine Cycle (ORC) for micro-scale CHP applications. Appl. Therm. Eng. 2015, 79, 1–8. [Google Scholar] [CrossRef]
  24. Flores, R.A.; Aviña Jiménez, H.M.; González, E.P.; González Uribe, L.A. Aerothermodynamic design of 10 kW radial inflow turbine for an organic flashing cycle using low-enthalpy resources. J. Clean. Prod. 2020, 251, 119713. [Google Scholar] [CrossRef]
  25. Kaczmarczyk, M.; Tomaszewska, B.; Operacz, A. Sustainable Utilization of Low Enthalpy Geothermal Resources to Electricity Generation through a Cascade System. Energies 2020, 13, 2495. [Google Scholar] [CrossRef]
  26. Pili, R.; Eyerer, S.; Dawo, F.; Wieland, C.; Spliethoff, H. Development of a non-linear state estimator for advanced control of an ORC test rig for geothermal application. Renew. Energy 2020, 161, 676–690. [Google Scholar] [CrossRef]
  27. Dong, L.; Liu, H.; Riffat, S. Development of small-scale and micro-scale biomass-fuelled CHP systems—A literature review. Appl. Therm. Eng. 2009, 29, 2119–2126. [Google Scholar] [CrossRef] [Green Version]
  28. Morrone, P.; Algieri, A.; Castiglione, T. Hybridisation of biomass and concentrated solar power systems in transcritical organic Rankine cycles: A micro combined heat and power application. Energy Convers. Manag. 2019, 180, 757–768. [Google Scholar] [CrossRef]
  29. Algieri, A.; Morrone, P. Techno-economic Analysis of Biomass-fired ORC Systems for Single-family Combined Heat and Power (CHP) Applications. Energy Procedia 2014, 45, 1285–1294. [Google Scholar] [CrossRef] [Green Version]
  30. Eyerer, S.; Dawo, F.; Wieland, C.; Spliethoff, H. Advanced ORC architecture for geothermal combined heat and power generation. Energy 2020, 205, 117967. [Google Scholar] [CrossRef]
  31. Drescher, U.; Brüggemann, D. Fluid selection for the Organic Rankine Cycle (ORC) in biomass power and heat plants. Appl. Therm. Eng. 2007, 27, 223–228. [Google Scholar] [CrossRef]
  32. Calise, F.; Cappiello, F.L.; Dentice d’Accadia, M.; Vicidomini, M. Thermo-Economic Analysis of Hybrid Solar-Geothermal Polygeneration Plants in Different Configurations. Energies 2020, 13, 2391. [Google Scholar] [CrossRef]
  33. Lecompte, S.; Ntavou, E.; Tchanche, B.; Kosmadakis, G.; Pillai, A.; Manolakos, D.; De Paepe, M. Review of Experimental Research on Supercritical and Transcritical Thermodynamic Cycles Designed for Heat Recovery Application. Appl. Sci. 2019, 9, 2571. [Google Scholar] [CrossRef] [Green Version]
  34. Macchi, E.; Astolfi, M. Organic Rankine Cycle (ORC) Power Systems: Technologies and Applications; Woodhead Publishing: Cambridge, UK, 2016; ISBN 978-0-08-100510-1. [Google Scholar]
  35. Rivera Diaz, A.; Kaya, E.; Zarrouk, S.J. Reinjection in geothermal fields—A worldwide review update. Renew. Sustain. Energy Rev. 2016, 53, 105–162. [Google Scholar] [CrossRef]
  36. El Haj Assad, M.; Bani-Hani, E.; Khalil, M. Performance of geothermal power plants (single, dual, and binary) to compensate for LHC-CERN power consumption: Comparative study. Geotherm. Energy 2017, 5, 17. [Google Scholar] [CrossRef]
  37. Lee, I.; Tester, J.W.; You, F. Systems analysis, design, and optimization of geothermal energy systems for power production and polygeneration: State-of-the-art and future challenges. Renew. Sustain. Energy Rev. 2019, 109, 551–577. [Google Scholar] [CrossRef]
  38. Heberle, F.; Brüggemann, D. Thermoeconomic Analysis of Hybrid Power Plant Concepts for Geothermal Combined Heat and Power Generation. Energies 2014, 7, 4482–4497. [Google Scholar] [CrossRef] [Green Version]
  39. Lecompte, S.; Huisseune, H.; Van den Broek, M.; De Schampheleire, S.; De Paepe, M. Part load based thermo-economic optimization of the Organic Rankine Cycle (ORC) applied to a combined heat and power (CHP) system. Appl. Energy 2013, 111, 871–881. [Google Scholar] [CrossRef]
  40. Eller, T.; Heberle, F.; Brüggemann, D. Transient Simulation of Geothermal Combined Heat and Power Generation for a Resilient Energetic and Economic Evaluation. Energies 2019, 12, 894. [Google Scholar] [CrossRef] [Green Version]
  41. Fiaschi, D.; Lifshitz, A.; Manfrida, G.; Tempesti, D. An innovative ORC power plant layout for heat and power generation from medium- to low-temperature geothermal resources. Energy Convers. Manag. 2014, 88, 883–893. [Google Scholar] [CrossRef]
  42. Li, T.; Zhu, J.; Zhang, W. Comparative analysis of series and parallel geothermal systems combined power, heat and oil recovery in oilfield. Appl. Therm. Eng. 2013, 50, 1132–1141. [Google Scholar] [CrossRef]
  43. Van Erdeweghe, S.; Van Bael, J.; Laenen, B.; D’haeseleer, W. Comparison of series/parallel configuration for a low-T geothermal CHP plant, coupled to thermal networks. Renew. Energy 2017, 111, 494–505. [Google Scholar] [CrossRef]
  44. Mohammadzadeh Bina, S.; Jalilinasrabady, S.; Fujii, H. Energy, economic and environmental (3E) aspects of internal heat exchanger for ORC geothermal power plants. Energy 2017, 140, 1096–1106. [Google Scholar] [CrossRef]
  45. Marty, F.; Sochard, S.; Serra, S.; Reneaume, J.-M. Multi-objective approach for a combined heat and power geothermal plant optimization. Chem. Prod. Process Model. 2020, 1. [Google Scholar] [CrossRef]
  46. Pastor-Martinez, E.; Rubio-Maya, C.; Ambriz-Díaz, V.M.; Belman-Flores, J.M.; Pacheco-Ibarra, J.J. Energetic and exergetic performance comparison of different polygeneration arrangements utilizing geothermal energy in cascade. Energy Convers. Manag. 2018, 168, 252–269. [Google Scholar] [CrossRef]
  47. El-Emam, R.S.; Dincer, I. Exergy and exergoeconomic analyses and optimization of geothermal organic Rankine cycle. Appl. Therm. Eng. 2013, 59, 435–444. [Google Scholar] [CrossRef]
  48. Toselli, D.; Heberle, F.; Brüggemann, D. Techno-Economic Analysis of Hybrid Binary Cycles with Geothermal Energy and Biogas Waste Heat Recovery. Energies 2019, 12, 1969. [Google Scholar] [CrossRef] [Green Version]
  49. Van Erdeweghe, S.; Van Bael, J.; Laenen, B.; D’haeseleer, W. Design and off-design optimization procedure for low-temperature geothermal organic Rankine cycles. Appl. Energy 2019, 242, 716–731. [Google Scholar] [CrossRef]
  50. Astolfi, M.; La Diega, L.N.; Romano, M.C.; Merlo, U.; Filippini, S.; Macchi, E. Techno-economic optimization of a geothermal ORC with novel “Emeritus” heat rejection units in hot climates. Renew. Energy 2020, 147, 2810–2821. [Google Scholar] [CrossRef]
  51. Van Erdeweghe, S.; Van Bael, J.; Laenen, B.; D’haeseleer, W. Optimal configuration for a low-temperature geothermal CHP plant based on thermoeconomic optimization. Energy 2019, 179, 323–335. [Google Scholar] [CrossRef]
  52. Wieland, C.; Meinel, D.; Eyerer, S.; Spliethoff, H. Innovative CHP concept for ORC and its benefit compared to conventional concepts. Appl. Energy 2016, 183, 478–490. [Google Scholar] [CrossRef]
  53. Habka, M.; Ajib, S. Evaluation of mixtures performances in Organic Rankine Cycle when utilizing the geothermal water with and without cogeneration. Appl. Energy 2015, 154, 567–576. [Google Scholar] [CrossRef]
  54. Marty, F.; Serra, S.; Sochard, S.; Reneaume, J.-M. Simultaneous optimization of the district heating network topology and the Organic Rankine Cycle sizing of a geothermal plant. Energy 2018, 159, 1060–1074. [Google Scholar] [CrossRef]
  55. Marty, F.; Serra, S.; Sochard, S.; Reneaume, J.-M. Economic optimization of a combined heat and power plant: Heat vs electricity. Energy Procedia 2017, 116, 138–151. [Google Scholar] [CrossRef]
  56. Ademe. Fongeosec—Conception d’un Démonstrateur de Centrale Géothermique Haute Enthalpie. Available online: https://www.ademe.fr/fongeosec (accessed on 3 September 2020).
  57. Eyerer, S.; Schifflechner, C.; Hofbauer, S.; Bauer, W.; Wieland, C.; Spliethoff, H. Combined heat and power from hydrothermal geothermal resources in Germany: An assessment of the potential. Renew. Sustain. Energy Rev. 2020, 120, 109661. [Google Scholar] [CrossRef]
  58. Algieri, A.; Morrone, P. Comparative energetic analysis of high-temperature subcritical and transcritical Organic Rankine Cycle (ORC). A biomass application in the Sibari district. Appl. Therm. Eng. 2012, 36, 236–244. [Google Scholar] [CrossRef]
  59. Cau, G.; Cocco, D. Comparison of Medium-size Concentrating Solar Power Plants based on Parabolic Trough and Linear Fresnel Collectors. Energy Procedia 2014, 45, 101–110. [Google Scholar] [CrossRef]
  60. Song, J.; Loo, P.; Teo, J.; Markides, C.N. Thermo-Economic Optimization of Organic Rankine Cycle (ORC) Systems for Geothermal Power Generation: A Comparative Study of System Configurations. Front. Energy Res. 2020, 8, 6. [Google Scholar] [CrossRef] [Green Version]
  61. Preißinger, M.; Heberle, F.; Brüggemann, D. Advanced Organic Rankine Cycle for geothermal application. Int. J. Low Carb. Tech. 2013, 8, i62–i68. [Google Scholar] [CrossRef] [Green Version]
  62. Astolfi, M.; Romano, M.C.; Bombarda, P.; Macchi, E. Binary ORC (organic Rankine cycles) power plants for the exploitation of medium–low temperature geothermal sources—Part A: Thermodynamic optimization. Energy 2014, 66, 423–434. [Google Scholar] [CrossRef]
  63. Van Erdeweghe, S.; Van Bael, J.; D’haeseleer, W. Optimal configuration, design and control of a binary geothermal combined heat-and-power plant. Energy Convers. Manag. 2019, 198, 111854. [Google Scholar] [CrossRef]
  64. Manente, G.; Toffolo, A.; Lazzaretto, A.; Paci, M. An Organic Rankine Cycle off-design model for the search of the optimal control strategy. Energy 2013, 58, 97–106. [Google Scholar] [CrossRef]
  65. Nasruddin, N.; Dwi Saputra, I.; Mentari, T.; Bardow, A.; Marcelina, O.; Berlin, S. Exergy, exergoeconomic, and exergoenvironmental optimization of the geothermal binary cycle power plant at Ampallas, West Sulawesi, Indonesia. Therm. Sci. Eng. Prog. 2020, 19, 100625. [Google Scholar] [CrossRef]
  66. Dumont, O.; Dickes, R.; De Rosa, M.; Douglas, R.; Lemort, V. Technical and economic optimization of subcritical, wet expansion and transcritical Organic Rankine Cycle (ORC) systems coupled with a biogas power plant. Energy Convers. Manag. 2018, 157, 294–306. [Google Scholar] [CrossRef] [Green Version]
  67. Garg, P.; Orosz, M.S.; Kumar, P. Thermo-economic evaluation of ORCs for various working fluids. Appl. Therm. Eng. 2016, 109, 841–853. [Google Scholar] [CrossRef] [Green Version]
  68. Moloney, F.; Almatrafi, E.; Goswami, D.Y. Working fluid parametric analysis for regenerative supercritical organic Rankine cycles for medium geothermal reservoir temperatures. Energy Procedia 2017, 129, 599–606. [Google Scholar] [CrossRef]
  69. Toffolo, A.; Lazzaretto, A.; Manente, G.; Paci, M. A multi-criteria approach for the optimal selection of working fluid and design parameters in Organic Rankine Cycle systems. Appl. Energy 2014, 121, 219–232. [Google Scholar] [CrossRef]
  70. Akbari, A.D.; Mahmoudi, S.M.S. Thermoeconomic analysis & optimization of the combined supercritical CO2 (carbon dioxide) recompression Brayton/organic Rankine cycle. Energy 2014, 78, 501–512. [Google Scholar] [CrossRef]
  71. Heberle, F.; Brüggemann, D. Exergy based fluid selection for a geothermal Organic Rankine Cycle for combined heat and power generation. Appl. Therm. Eng. 2010, 30, 1326–1332. [Google Scholar] [CrossRef]
  72. Wang, X.; Levy, E.K.; Pan, C.; Romero, C.E.; Banerjee, A.; Rubio-Maya, C.; Pan, L. Working fluid selection for organic Rankine cycle power generation using hot produced supercritical CO2 from a geothermal reservoir. Appl. Therm. Eng. 2019, 149, 1287–1304. [Google Scholar] [CrossRef]
  73. Gawlik, K.; Hassani, V. Advanced Binary Cycles: Optimum Working Fluids. In Proceedings of the IECEC-97 Thirty-Second Intersociety Energy Conversion Engineering Conference (Cat. No.97CH6203), Honolulu, HI, USA, 27 July–1 August 1997; Volume 3, pp. 1809–1814. [Google Scholar]
  74. Heberle, F.; Brüggemann, D. Thermo-Economic Evaluation of Organic Rankine Cycles for Geothermal Power Generation Using Zeotropic Mixtures. Energies 2015, 8, 2097–2124. [Google Scholar] [CrossRef] [Green Version]
  75. Askari, I.; Calise, F.; Vicidomini, M. Design and Comparative Techno-Economic Analysis of Two Solar Polygeneration Systems Applied for Electricity, Cooling and Fresh Water Production. Energies 2019, 12, 4401. [Google Scholar] [CrossRef] [Green Version]
  76. Ghaem Sigarchian, S.; Malmquist, A.; Martin, V. Design Optimization of a Complex Polygeneration System for a Hospital. Energies 2018, 11, 1071. [Google Scholar] [CrossRef] [Green Version]
  77. Carlino, S.; Troiano, A.; Di Giuseppe, M.G.; Tramelli, A.; Troise, C.; Somma, R.; De Natale, G. Exploitation of geothermal energy in active volcanic areas: A numerical modelling applied to high temperature Mofete geothermal field, at Campi Flegrei caldera (Southern Italy). Renew. Energy 2016, 87, 54–66. [Google Scholar] [CrossRef]
  78. Carlino, S.; Somma, R.; Troise, C.; De Natale, G. The geothermal exploration of Campanian volcanoes: Historical review and future development. Renew. Sustain. Energy Rev. 2012, 16, 1004–1030. [Google Scholar] [CrossRef]
  79. Harvey, L.D.D. Energy and the New Reality 1: Energy Efficiency and the Demand for Energy Services; Earthscan: London, UK; Washington, DC, USA, 2010; ISBN 978-1-84407-912-4. [Google Scholar]
  80. European Commission. Commission Delegated Regulation (EU) 2015/2402 of 12 October 2015 Reviewing Harmonised Efficiency Reference Values for Separate Production of Electricity and Heat in Application of Directive 2012/27/EU of the European Parliament and of the Council and Repealing Commission Implementing Decision 2011/877/EU. 2015. Available online: https://eur-lex.europa.eu/eli/reg_del/2015/2402/oj (accessed on 16 July 2020).
  81. Li, J.; Liu, Q.; Ge, Z.; Duan, Y.; Yang, Z. Thermodynamic performance analyses and optimization of subcritical and transcritical organic Rankine cycles using R1234ze(E) for 100–200 °C heat sources. Energy Convers. Manag. 2017, 149, 140–154. [Google Scholar] [CrossRef]
  82. Gao, T.; Liu, C. Off-Design Performances of Subcritical and Supercritical Organic Rankine Cycles in Geothermal Power Systems under an Optimal Control Strategy. Energies 2017, 10, 1185. [Google Scholar] [CrossRef] [Green Version]
  83. Shengjun, Z.; Huaixin, W.; Tao, G. Performance comparison and parametric optimization of subcritical Organic Rankine Cycle (ORC) and transcritical power cycle system for low-temperature geothermal power generation. Appl. Energy 2011, 88, 2740–2754. [Google Scholar] [CrossRef]
  84. Yang, M.-H.; Yeh, R.-H. Economic performances optimization of the transcritical Rankine cycle systems in geothermal application. Energy Convers. Manag. 2015, 95, 20–31. [Google Scholar] [CrossRef]
  85. Wei, D.; Liu, C.; Geng, Z. Conversion of Low-Grade Heat from Multiple Streams in Methanol to Olefin (MTO) Process Based on Organic Rankine Cycle (ORC). Appl. Sci. 2020, 10, 3617. [Google Scholar] [CrossRef]
  86. Unamba, C.K.; Sapin, P.; Li, X.; Song, J.; Wang, K.; Shu, G.; Tian, H.; Markides, C.N. Operational Optimisation of a Non-Recuperative 1-kWe Organic Rankine Cycle Engine Prototype. Appl. Sci. 2019, 9, 3024. [Google Scholar] [CrossRef] [Green Version]
  87. Valencia, G.; Duarte, J.; Isaza-Roldan, C. Thermoeconomic Analysis of Different Exhaust Waste-Heat Recovery Systems for Natural Gas Engine Based on ORC. Appl. Sci. 2019, 9, 4017. [Google Scholar] [CrossRef] [Green Version]
  88. Karimi, S.; Mansouri, S. A comparative profitability study of geothermal electricity production in developed and developing countries: Exergoeconomic analysis and optimization of different ORC configurations. Renew. Energy 2018, 115, 600–619. [Google Scholar] [CrossRef]
  89. Kazemi, S.; Nor, M.I.M.; Teoh, W.H. Thermodynamic and economic investigation of an ionic liquid as a new proposed geothermal fluid in different organic Rankine cycles for energy production. Energy 2020, 193, 116722. [Google Scholar] [CrossRef]
  90. Samadi, F.; Kazemi, N. Exergoeconomic analysis of zeotropic mixture on the new proposed organic Rankine cycle for energy production from geothermal resources. Renew. Energy 2020, 152, 1250–1265. [Google Scholar] [CrossRef]
  91. Algieri, A.; Morrone, P. Energy analysis of Organic Rankine Cycles for biomass applications. Therm. Sci. 2015, 19, 193–205. [Google Scholar] [CrossRef]
  92. Lemmon, E.W.; Bell, I.H.; Huber, M.L.; McLinden, M.O. Refprop Documentation, Release 10.0. 2018. Available online: https://trc.nist.gov/refprop/REFPROP.PDF (accessed on 16 July 2020).
  93. Algieri, A.; Morrone, P.; Bova, S. Techno-Economic Analysis of Biofuel, Solar and Wind Multi-Source Small-Scale CHP Systems. Energies 2020, 13, 3002. [Google Scholar] [CrossRef]
  94. Zhang, X.; Cao, M.; Yang, X.; Guo, H.; Wang, J. Economic Analysis of Organic Rankine Cycle Using R123 and R245fa as Working Fluids and a Demonstration Project Report. Appl. Sci. 2019, 9, 288. [Google Scholar] [CrossRef] [Green Version]
  95. Gnielinski, V. New equations for heat and mass-transfer in turbulent pipe and channel flow. Int. Chem. Eng. 1976, 16, 359–368. [Google Scholar]
  96. Kern, D.Q. Process Heat Transfer; McGraw-Hill Classic Textbook Reissue Series; McGraw-Hill: Pennsylvania, NY, USA, 1990; ISBN 978-0-07-034190-6. [Google Scholar]
  97. Krasnoshchekov, E.A.; Protopopov, V.S.; Van, F.; Kuraeva, I.V. Experimental Investigation of Heat Transfer for Carbon Dioxide in the Supercritical Region. In Proceedings of the 2nd All-Soviet Union Conference on Heat and Mass Transfer, Minsk, Belarus, May 1964; Volume 1, pp. 26–35. [Google Scholar]
  98. Pioro, I.L.; Khartabil, H.F.; Duffey, R.B. Heat transfer to supercritical fluids flowing in channels—Empirical correlations (survey). Nucl. Eng. Des. 2004, 230, 69–91. [Google Scholar] [CrossRef]
  99. Petukhov, B.S.; Krasnoshchekov, E.A.; Protopopov, V.S. An Investigation of Heat Transfer to Fluids Flowing in Pipes under Supercritical Conditions. In Proceedings of the International Developments in Heat Transfer, Boulder, CO, USA, 28 August–1 September 1961; pp. 569–578. [Google Scholar]
  100. Chang, W.; Chu, X.; Binte Shaik Fareed, A.F.; Pandey, S.; Luo, J.; Weigand, B.; Laurien, E. Heat transfer prediction of supercritical water with artificial neural networks. Appl. Therm. Eng. 2018, 131, 815–824. [Google Scholar] [CrossRef]
  101. Vijayan, P.K.; Nayak, A.K.; Kumar, N. Governing Differential Equations for Natural Circulation Systems. In Single-Phase, Two-Phase and Supercritical Natural Circulation Systems; Elsevier: Amsterdam, The Netherlands, 2019; pp. 69–118. ISBN 978-0-08-102486-7. [Google Scholar]
  102. Baik, Y.-J.; Kim, M.; Chang, K.-C.; Lee, Y.-S.; Yoon, H.-K. Power enhancement potential of a mixture transcritical cycle for a low-temperature geothermal power generation. Energy 2012, 47, 70–76. [Google Scholar] [CrossRef]
  103. Baik, Y.-J.; Kim, M.; Chang, K.C.; Kim, S.J. Power-based performance comparison between carbon dioxide and R125 transcritical cycles for a low-grade heat source. Appl. Energy 2011, 88, 892–898. [Google Scholar] [CrossRef]
  104. He, J.; Dang, C.; Hihara, E. Supercritical heat transfer characteristics of R1233zd(E) in vertically upward flow. Int. J. Heat Mass Transf. 2018, 127, 497–505. [Google Scholar] [CrossRef]
  105. Alibaba, M.; Pourdarbani, R.; Hasan Khoshgoftar Manesh, M.; Herrera-Miranda, I.; Gallardo-Bernal, I.; Hernández-Hernández, J.L. Conventional and Advanced Exergy-Based Analysis of Hybrid Geothermal–Solar Power Plant Based on ORC Cycle. Appl. Sci. 2020, 10, 5206. [Google Scholar] [CrossRef]
  106. Liu, Q.; Duan, Y.; Yang, Z. Performance analyses of geothermal organic Rankine cycles with selected hydrocarbon working fluids. Energy 2013, 63, 123–132. [Google Scholar] [CrossRef]
  107. Dai, X.; Shi, L.; An, Q.; Qian, W. Screening of hydrocarbons as supercritical ORCs working fluids by thermal stability. Energy Convers. Manag. 2016, 126, 632–637. [Google Scholar] [CrossRef]
  108. Chacartegui, R.; Sánchez, D.; Muñoz, J.M.; Sánchez, T. Alternative ORC bottoming cycles FOR combined cycle power plants. Appl. Energy 2009, 86, 2162–2170. [Google Scholar] [CrossRef]
  109. Calise, F.; d’Accadia, M.D.; Vicidomini, M.; Scarpellino, M. Design and simulation of a prototype of a small-scale solar CHP system based on evacuated flat-plate solar collectors and Organic Rankine Cycle. Energy Convers. Manag. 2015, 90, 347–363. [Google Scholar] [CrossRef]
  110. Algieri, A.; Morrone, P. Energetic analysis of biomass-fired ORC systems for micro-scale combined heat and power (CHP) generation. A possible application to the Italian residential sector. Appl. Therm. Eng. 2014, 71, 751–759. [Google Scholar] [CrossRef]
  111. Xi, H.; Li, M.-J.; He, Y.-L.; Tao, W.-Q. A graphical criterion for working fluid selection and thermodynamic system comparison in waste heat recovery. Appl. Therm. Eng. 2015, 89, 772–782. [Google Scholar] [CrossRef]
  112. Morrone, P.; Algieri, A.; Castiglione, T.; Perrone, D.; Bova, S. Investigation of the Energy Performance of Multi-Source Integrated CHP Systems for Small-Scale Applications. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2019; Volume 2191, p. 020115. [Google Scholar] [CrossRef]
  113. Turton, R.; Shaeiwitz, J.A.; Bhattacharyya, D.; Whiting, W.B. Analysis, Synthesis, and Design of Chemical Processes, 5th ed.; Pearson Education: New York, NY, USA, 2018; ISBN 978-0-13-417740-3. [Google Scholar]
  114. Zhang, C.; Liu, C.; Xu, X.; Li, Q.; Wang, S. Energetic, exergetic, economic and environmental (4E) analysis and multi-factor evaluation method of low GWP fluids in trans-critical organic Rankine cycles. Energy 2019, 168, 332–345. [Google Scholar] [CrossRef]
  115. Meng, F.; Wang, E.; Zhang, B.; Zhang, F.; Zhao, C. Thermo-economic analysis of transcritical CO2 power cycle and comparison with Kalina cycle and ORC for a low-temperature heat source. Energy Convers. Manag. 2019, 195, 1295–1308. [Google Scholar] [CrossRef]
  116. Chemical Engineering. Plant Cost Index. Available online: https://www.chemengonline.com/site/plant-cost-index/ (accessed on 16 July 2020).
  117. Okoro, O.V.; Faloye, F.D. Comparative Assessment of Thermo-Syngas Fermentative and Liquefaction Technologies as Waste Plastics Repurposing Strategies. AgriEngineering 2020, 2, 378–392. [Google Scholar] [CrossRef]
  118. European Central Bank. ECB Euro Reference Exchange Rate: US Dollar (USD). Available online: https://www.ecb.europa.eu/stats/policy_and_exchange_rates/euro_reference_exchange_rates (accessed on 1 July 2020).
  119. Eurostat. Database. Available online: https://ec.europa.eu/eurostat/data/database (accessed on 16 July 2020).
  120. Dinçer, İ.; Rosen, M.; Ahmadi, P. Optimization of Energy Systems; John Wiley & Sons: Chichester, West Sussex, UK, 2018; ISBN 978-1-118-89450-7. [Google Scholar]
  121. Algieri, A.; Zema, D.A.; Nicotra, A.; Zimbone, S.M. Potential energy exploitation in collective irrigation systems using pumps as turbines: A case study in Calabria (Southern Italy). J. Clean. Prod. 2020, 257, 120538. [Google Scholar] [CrossRef]
  122. Koffi, B.; Cerutti, A.K.; Duerr, M.; Iancu, A.; Kona, A.; Janssens-Maenhout. CoM Default Emission Factors for the Member States of the European Union—Version 2017. Available online: https://jeodpp.jrc.ec.europa.eu/ftp/jrc-opendata/COM-EF/dataset/comw/JRC-CoM-EF-CoMW-EF-2017.pdf (accessed on 16 July 2020).
  123. Cerutti, A.K.; Duerr, M.; Iancu, A.; Janssens-Maenhout, G.; Koffi, B.; Kona, A.; Covenant of Mayors & Mayors Adapt Offices; European Commission; Joint Research Centre. Covenant of Mayors for Climate and Energy: Default Emission Factors for Local Emission Inventories; Version 2017; Publications Office of the European Union: Luxembourg, 2017; ISBN 978-92-79-71479-5. [Google Scholar]
Figure 1. Simplified scheme of the integrated geothermal-driven CHP system.
Figure 1. Simplified scheme of the integrated geothermal-driven CHP system.
Applsci 10 06639 g001
Figure 2. Typical system layout (a) and T,s diagram (b) of an organic Rankine cycle (ORC) with internal regenerator (IHE) and transcritical conditions.
Figure 2. Typical system layout (a) and T,s diagram (b) of an organic Rankine cycle (ORC) with internal regenerator (IHE) and transcritical conditions.
Applsci 10 06639 g002
Figure 3. Electric and thermal demand of the investigated commercial centre.
Figure 3. Electric and thermal demand of the investigated commercial centre.
Applsci 10 06639 g003
Figure 4. Influence of the maximum temperature on ORC electric power and efficiency. Simple and IHE configurations.
Figure 4. Influence of the maximum temperature on ORC electric power and efficiency. Simple and IHE configurations.
Applsci 10 06639 g004
Figure 5. Multi-objective optimisation for the selection of the proper system configuration.
Figure 5. Multi-objective optimisation for the selection of the proper system configuration.
Applsci 10 06639 g005
Figure 6. Influence of the electric load on thermal power, electric power and geothermal flow at the entrance of ORC sub-system (a). Effect of the electric load on EUF, electric and marginal efficiency (b).
Figure 6. Influence of the electric load on thermal power, electric power and geothermal flow at the entrance of ORC sub-system (a). Effect of the electric load on EUF, electric and marginal efficiency (b).
Applsci 10 06639 g006
Figure 7. Monthly electric and thermal balances for mixed (a,b) thermal- (c,d) and electric-driven (e,f) strategy.
Figure 7. Monthly electric and thermal balances for mixed (a,b) thermal- (c,d) and electric-driven (e,f) strategy.
Applsci 10 06639 g007
Figure 8. Comparison of energy balance on yearly basis for mixed, electric-driven and thermal-driven strategy.
Figure 8. Comparison of energy balance on yearly basis for mixed, electric-driven and thermal-driven strategy.
Applsci 10 06639 g008
Figure 9. Components cost of the integrated combined energy system.
Figure 9. Components cost of the integrated combined energy system.
Applsci 10 06639 g009
Figure 10. Influence of the operating strategy on the economic viability of the integrated energy systems.
Figure 10. Influence of the operating strategy on the economic viability of the integrated energy systems.
Applsci 10 06639 g010
Figure 11. Hourly electric efficiency (a) and energy utilisation factor (b) for the integrated geothermal energy system. Mixed strategy.
Figure 11. Hourly electric efficiency (a) and energy utilisation factor (b) for the integrated geothermal energy system. Mixed strategy.
Applsci 10 06639 g011
Figure 12. Hourly thermal and electric balance of the integrated energy system unit in January (a,c) and April (b,d) typical days. Mixed strategy.
Figure 12. Hourly thermal and electric balance of the integrated energy system unit in January (a,c) and April (b,d) typical days. Mixed strategy.
Applsci 10 06639 g012
Figure 13. Hourly thermal and electric balance of the integrated energy system unit in July (a,c) and October (b,d) typical days. Mixed strategy.
Figure 13. Hourly thermal and electric balance of the integrated energy system unit in July (a,c) and October (b,d) typical days. Mixed strategy.
Applsci 10 06639 g013
Table 1. Operating conditions and main assumptions.
Table 1. Operating conditions and main assumptions.
Parameters
Mass flow rate of geothermal fluid(kg/s)1
Temperature of geothermal fluid(°C)230
Minimum reinjection temperature(°C)70
Pinch-point temperature in geothermal circuit(°C)10
Geothermal heat exchanger efficiency(-)0.92
Geothermal pump efficiency(-)0.70
ORC configuration Transcritical
ORC working fluid Isobutane
ORC maximum temperature range(°C)140.0–220.0
ORC maximum pressure range(bar)37.38–65.32
ORC condensation temperature(°C)30.0
ORC condensation pressure(bar)4.05
Maximum ORC pump efficiency(-)0.70
Maximum ORC turbine efficiency(-)0.80
Electro-mechanical efficiency(-)0.95
Cooling pump efficiency(-)0.80
Head of cooling pump(m)10
Pinch-point temperature in cooling system(°C)5
Heat exchanger inner tube diameter(mm)10.92
Heat exchanger outer tube diameter(mm)12.70
Gas-fired auxiliary boiler efficiency(-)0.90
Natural gas lower heating value(MJ/Sm3)34.54
Reference thermal efficiency(-)0.92
Table 2. Equipment cost assumptions for bare module methodology.
Table 2. Equipment cost assumptions for bare module methodology.
ParameterHeat ExchangerPumpTurbine
k14.32473.38922.2476
k2−0.30030.05361.4965
k30.16340.1538−0.1618
c10.0388−0.39350
c2−0.11270.39570
c30.0818−0.00230
b11.631.89-
b21.661.35-
FM1.001.60-
FBM--3.5
Table 3. Further assumptions for the economic analysis.
Table 3. Further assumptions for the economic analysis.
ParameterUnitValue
Investment Period(Years)20
Interest rate(%)2
Specific revenue for the saved thermal energy(c€/kWhth)6.7
Specific revenue for the saved electricity(c€/kWhel)18.8
Specific value of the electricity injected into the grid(c€/kWh)10.0
Specific cost of the electricity withdrawn from the grid(c€/kWh)18.8
Specific cost of natural gas(c€/kWh)6.0
CEPCI2001(-)397.0
CEPCI2019(-)607.5
EUR/USD exchange rate(EUR/USD)1.12
Maintenance cost/Investment cost(%)2.0
Table 4. Optimised integrated geothermal CHP systems.
Table 4. Optimised integrated geothermal CHP systems.
ParameterUnitValue
Maximum Electric Power(kWel)101.4
Maximum Thermal Power(kWth)249.5
Electric Production(MWhel)820.8
Thermal Production(MWhth)357.4
Electric Self-consumption(%)94.9
Electric Surplus(%)59.0
Electric Integration(%)5.1
Thermal Self-consumption(%)90.2
Thermal Surplus(%)0.0
Thermal Integration(%)9.8
Global Electric Efficiency(%)13.4
Energy Utilisation Factor(%)19.3
Equivalent Operating hours(h)8095
Natural gas consumption(Sm3)4512.6
Natural gas saving(Sm3)41,386.0
Initial investment(k€)636.8
Net present value(k€)1617.5
Payback period(years)4.90
Reduced CO2 emissions(tCO2)253.8
Reduced GHG emissions(tCO2,eq)309.9

Share and Cite

MDPI and ACS Style

Morrone, P.; Algieri, A. Integrated Geothermal Energy Systems for Small-Scale Combined Heat and Power Production: Energy and Economic Investigation. Appl. Sci. 2020, 10, 6639. https://0-doi-org.brum.beds.ac.uk/10.3390/app10196639

AMA Style

Morrone P, Algieri A. Integrated Geothermal Energy Systems for Small-Scale Combined Heat and Power Production: Energy and Economic Investigation. Applied Sciences. 2020; 10(19):6639. https://0-doi-org.brum.beds.ac.uk/10.3390/app10196639

Chicago/Turabian Style

Morrone, Pietropaolo, and Angelo Algieri. 2020. "Integrated Geothermal Energy Systems for Small-Scale Combined Heat and Power Production: Energy and Economic Investigation" Applied Sciences 10, no. 19: 6639. https://0-doi-org.brum.beds.ac.uk/10.3390/app10196639

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop