Accepted Manuscript Simulation of polygeneration systems
Francesco Calise, Giulio de Notaristefani di Vastogirardi, Massimo Dentice d’ Accadia, Maria Vicidomini PII:
S0360-5442(18)31572-X
DOI:
10.1016/j.energy.2018.08.052
Reference:
EGY 13528
To appear in:
Energy
Received Date:
19 December 2017
Accepted Date:
07 August 2018
Please cite this article as: Francesco Calise, Giulio de Notaristefani di Vastogirardi, Massimo Dentice d’Accadia, Maria Vicidomini, Simulation of polygeneration systems, Energy (2018), doi: 10.1016/j.energy.2018.08.052
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT 1 of 67
1
Simulation of polygeneration systems
2 3
Francesco Calise*, Giulio de Notaristefani di Vastogirardi, Massimo Dentice d’Accadia, Maria Vicidomini
4 5 6 7 8 9 10
Department of Industrial Engineering, University of Naples Federico II, Napoli 80125, Italy;
11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51
[email protected] [email protected] [email protected] * Correspondence:
[email protected]; Tel.: +39-081-768-2301
Abstract: This work aims at presenting the current works concerning the polygeneration systems simulation, by specially focusing on the potential integration of different technologies into a single system. Polygeneration allows one to produce energy vectors (power, heating and cooling) as wells as other useful products (hydrogen, syngas, biodiesel, fertilizers, drinking water etc.) by converting one or multiple energy sources. Polygeneration system can be fuelled by renewable sources (geothermal, solar, biomass, wind, hydro), as well as fossil fuels (natural gas, coal, hydrogen, etc.). In this paper innovative energy technologies, such as fuel cells and conventional ones are taken into account, by also focusing on the control strategies implemented for the proper management of polygeneration systems in general. Works regarding energy, economic and exergy analyses and system optimizations are also illustrated. Keywords: renewable energy; polygeneration; distributed generation; dynamic simulations 1.
Introduction
In the last decades, the international energy consumption has considerably increased, principally due to the economic and demographic growth of developing countries. So, more and more urgent issues arose, regarding the environmental impact and availability of fossil or non-renewable fuels (such as oil, gas, nuclear, carbon), which today satisfy most of the worldwide energy demand. At the same time, the consumption of potable water is also significantly increasing. Therefore, some noteworthy actions were taken in order to head towards a sustainable development, aimed at enhancing the energy performance of systems and processes and increasing the use of renewable energy sources [1]. In this framework, polygeneration systems represent an attractive and innovative concept. Such systems can provide simultaneously various energy vectors (power, heating, cooling) and other products (fresh water, syngas, urea, etc.). They become particularly interesting if coupled to renewable energy sources. In spite of their potential, the concept of polygeneration is still scarcely used. In fact, only 10% of the world power generation is obtained by means of polygeneration systems. Exceptions are the countries as The Netherlands, Finland and Denmark, where this ratio increases up to 30%–50%. Presently, many European Union countries are encouraging the development of sustainable and efficient systems based on polygeneration, since it is recognised as a strategic technology able to reduce greenhouse gas emissions. The same strategy was also adopted by the USA, aiming at cutting down the costs of energy production, mainly in the industrial sector. In order to pursue the integration of polygeneration technologies, industries and academic institutions are presently carrying out a noteworthy research effort, for developing and designing plants based on the use of both renewable and fossil energy sources. As for systems fuelled by fossil fuels, numerous investigations aiming at optimizing both trigeneration and cogeneration power plants were performed. As for polygeneration systems fuelled by renewable energy sources, most R&D activities are focused on hybrid plants in which traditional technologies (Stirling, gas turbine, internal combustion engines, combined cycle, etc.) are coupled to renewable energy sources (particularly biomass and solar energy) in a unique polygeneration system. Polygeneration plants are specially interesting when used to produce potable water, along with power, heating and cooling. This is a very attractive option, in particular for remote and inaccessible communities [2]. An example is represented by small or medium islands, where the availability of fossil fuels and potable water is often limited, whereas renewable sources are often abundant and there is obviously a large availability of sea water [3]. Desalination systems, in fact, are a great way to produce potable water when water scarcity represent a crucial issue, especially if one considers that more than 70% of the word population lives within 70 km from a sea [4].
ACCEPTED MANUSCRIPT 2 of 67
52 53 54 55 56 57 58 59 60 61 62
Polygeneration systems can also contribute to the diffusion of distribute generation (DG) systems. This kind of system is included in the more general concept of distributed energy systems (DES), which consists of the integration of different small-scale DG technologies, instead of a limited number of big, remote power plants (centralized production). In DES systems, the electricity is produced locally near to the final user, and, therefore, the negative effect due to the transmission losses is avoided. At the same time, in such systems, the concept of self-sufficiency, by decreasing the dependence from conventional energy grid, is also promoted. DES systems, fuelled by renewable energy, are usually considered as the best option among all the potential DG technologies. Such systems usually integrate district heating and cooling (DHC) and off-grid electric networks [5]. As mentioned before, the number of possible polygeneration layouts is virtually infinite, since it is possible to combine all the available fossil and renewable conversion technologies. This concept is clearly depicted in Figure 1. Hydrogen
Fertilizer
Methanol
Coal
Gasifier gasificat ion biomn CC
Production unit Syngas / Biogas
Ethanol
Boiler
Biomass
Wind
Electrical desalination unit
Fresh water
ICE Power
Natural gas
Electricity to user Cooling
GT, MGT ST
FC STE
Heat
Electrical chiller
Seawater
Thermal chiller
STC
DHW, Heating
PV Solar PVT
63
ORC
Thermal desalination unit
Fresh water
Geothermal
64
Figure 1. Possible combinations of fossil and renewable conversion technologies in polygeneration systems.
65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80
The topic of polygeneration is not new and several papers tried to analyse it in the past few years. In particular, the majority of the review papers dealing with polygeneration mainly focused on their possible arrangements and integration with renewables and/or buildings [6-8]. However, in authors’ knowledge, none of those papers focused on the simulation techniques implemented in order to analyse such polygeneration systems. In particular, a first attempt was recently performed by some of the authors a recent editorial paper [9], summarizing only a few studies included in a special issue dedicated to the simulation of polygeneration systems. Conversely, the review analysis performed in the present paper provides a comprehensive and detailed state-of-the-art of the available published literature concerning the simulation of polygeneration systems (fuelled by fossil fuels and/or renewable energy source). For these works, the methodology of analysis (exergoeconomic, exergy, environmental, economic, energy and thermoeconomic) as well as the innovative layouts designed to maximize the utilization of the input fuels are also defined. In particular, by collecting a high number of different works, the aim of this review, and therefore, the novelty with respect to the previous work [9], concerns the classification of the polygeneration systems adopting the simulation approach (based on dynamic simulations or steady-state simulations), regarding the following different issues, namely: i) methodology of analysis; ii) energy inputs (renewables and/or fossil fuels);
ACCEPTED MANUSCRIPT 3 of 67
81 82 83 84 85 86 87 88
iii) iv) v) vi) vii) 2.
energy conversion technologies (gas turbine, steam turbine, combined cycle, ORC, internal combustion engine, FC); energy (power, heating, cooling, etc.) and material outputs (liquid and gaseous fuels, desalinated water, etc.); use of polygeneration in buildings; control strategies and optimization techniques; contribution of polygeneration to the development of distribute generation systems. Methodology of analysis
89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106
This paper analyses different research works investigating polygeneration systems from a plurality of viewpoints, namely: energy, exergy, environmental and economic. From all the studies, it resulted that the design and the operation management of polygeneration systems requires a multidisciplinary approach and analysis, mainly due to high complexity of such systems, since they are always based on the integration of different technologies. In particular, such works were developed by adopting several common or advanced analysis methods, such as: i) energy analysis [10-14] of individual components as well as the system as a whole, oriented to estimate the energy saving of a proposed innovative system with respect the conventional technology; ii) economic analysis [2, 10, 11, 14-20], evaluating operation, maintenance and capital costs, in order to estimate the economic feasibility of the systems; iii) environmental analysis [11, 14, 20-26], usually based on a Life Cycle Analysis (LCA) approach [27]; iv) exergy analysis [10, 13, 21, 22, 28-32], exergoeconomic analysis [33-38] or thermoeconomic analysis [39-45], this is an useful tool for calculating the magnitude of the irreversibilities into a system, in order to evaluate possible enhancements of its performance; aiming at determining the economic value for each material flow or energy, by defining their price.
107 108 109 110 111 112 113 114
As above specified, polygeneration systems typically include a plurality of different technologies integrated into a single system; so, the development of optimized control strategies is a crucial issue in order to improve their performance. Therefore, many works concerning the simulation of the polygeneration systems are focused on the selection of the most appropriate operation control strategy [41, 46-51]. In addition, great attention is paid to the design of the system, based on different optimization methods [52-56], aiming at determining the optimal size of each component. The review presented in the next sections includes works based on dynamic simulations [3, 12, 18, 41, 57] and others in which steady-state simulations were performed [15, 58-63].
115 116 117 118 119 120 121 122 123 124 125 126
In particular, the works implementing the dynamic approach investigated several topologies of polygeneration systems, supplied both from fossil fuels (natural gas, coal, etc.) and by renewable energy sources (solar, biomass, geothermal, etc.) for the co-production of desalinated water, heating, cooling, electricity, H2, syngas or others. Dynamic simulations present many advantages, since they allow one to analyse with high accuracy the real operation of the system under evaluation, providing important information about the ways in which their performance can be optimized. Such dynamic simulations are often based on the software TRNSYS, a simulation tool diffusely used by the academic community. It provides a library of built-in components often based and validated vs experimental data [64]. Therefore, the results obtained by the works concerning the dynamic simulations carried out by TRNSYS software are usually considered highly reliable. Table 1 shows a classification reporting the studies only performed by the simulation model and the ones which coupled numerical and experimental analyses. For all the cases, the software implemented in the simulations is detailed.
127
Table 1. Classification studies: Software, experimental validation and modelling Work
Software
Experimental validation and
Only simulation
ACCEPTED MANUSCRIPT 4 of 67 simulation model [2, 3, 12, 37, 39, 41, 65-72] [11, 14, 19, 42, 55, 73-76] [16] [18] [22, 77-85] [86, 87] [4] [10] [88] [15, 17, 27, 34, 58-62, 89-97] [98] [99] [100] [21, 101, 102] [103] [104] [105] [106] [13] [107] [108, 109] [110] [111] [112] [113] [114]
128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143
TRNSYS TRNSYS TRNSYS, Engineering Equation Solver (EES) TRNSYS, EES EES EES EES, LINGO EES, CoolProp EES, BIORAISE, Aspen Plus Aspen Plus Aspen Plus, Dow Chemical Company RO System Analysis (ROSA). Aspen Plus, GE’s GateCycle MATLAB MATLAB, NIST/ASME MATLAB, Simulink MATLAB, Refprop MATLAB TRNSYS DEST ECLIPSE ChemCAD GAMS, CONOPT tool INSEL IPSEpro SIFRE MODEST
model x
x x x x x x x x x x x x x x x x x x x x x x x x x
3. Energy sources Polygeneration systems can be fuelled by fossil fuels (natural gas, coal, diesel etc.) and/or renewable energy sources (geothermal, solar, wind, biomass, etc.). In many cases, a mix of different sources is used. In the next paragraphs, the most significant literature works concerning the simulation of polygeneration systems fuelled by fossil fuels, renewable energy sources and by hybrid energy inputs are described. 3.1 Fossil fuels Several works investigating the simulation of polygeneration plants driven by fossil energy sources are available in literature [11, 17, 48, 73], also due to the fact that coal, natural gas and oil represent the 80% of global energy demand due to their affordability and availability [115]. Coal Currently, the clean and efficient use of coal is one of the most significant goals of the scientific research. In the last years, several advanced coal gasification technologies were developed to maximize the energy efficiency and to achieve a near-zero level of pollutant emissions. An example is the coal partial gasification technology analysed in the work presented by Li et al. [89], where the thermodynamic analysis of this process in a polygeneration system for the methanol and power generation, is carried out. The commercial software Aspen Plus was used to perform the simulation.
ACCEPTED MANUSCRIPT 5 of 67
144 145
Figure 2. Schematic of a polygeneration system based on partial gasification [89].
146 147 148 149 150 151
In contrast to traditional coal gasification technologies featured by total gasification of coal through simple or rough oxidization, partial gasification technology, based on pressurized circulating fluidized bed controls the quantities of oxygen agent and detention time during gasification, by achieving, relatively higher cold gas efficiency (up to 65.5%). The investigated system layout is reported in Figure 2. Here, the syngas from the partial gasifier is synthesized to methanol, whereas the unreacted gas enters gas turbine combined cycle to produce power. The system shows an energy and exergy efficiency equal to 51.16% and 50.58%.
152 153 154 155 156 157 158 159
Another novel coal gasification technology, aiming at separating CO2 emissions during combustion process without extra or with few energies required, is based on the concept of chemical looping combustion. This process is simulated in the polygeneration system investigated by Fan et al. [91]. In this system, the combustion flue gases from both air reactor and fuel reactor are sequentially fed into gas turbines for electricity production; here a heat recovery vapour generator unit for further electricity generation with driving an absorption chiller (ACH) in summer and a heat exchanger for daily heat water production is also simulated. This novel system maintains a maximum energy efficiency of 60.34%, whereas the fossil energy saving ratio of this process of 27.20%.
160 161 162 163 164 165 166 167 168
Utilizing captured CO2 for utility outputs in coal-based polygeneration has twofold advantages: eliminating the subsequent cost of CO2 sequestration and earning additional revenue from utility products using. In the work developed by Bose et al. [90], the performance assessment of a coal-based polygeneration for power with CO2 capture and subsequent use of this captured CO2 for production of a common fertilizer, i.e., urea is proposed. In particular, the syngas produced by gasification, is treated with steam in a water gas shift reactor, after that CO2 is captured. Produced hydrogen is partially utilized for power generation in a combined cycle gas turbine and for urea production. A detailed ASPEN Plus model is developed for this plant. It resulted that, the higher the percentage of captured CO2 used, the higher the production of urea, but a simultaneous decrease of net power output is achieved as both depend on amount of hydrogen of syngas used for these processes.
169 170 171 172 173 174 175
The majority of polygeneration energy systems analysed in literature involving coal gasification, use a single gas source and produce only a unique type of chemical product. Conversely, in the study reported by Li et al. [34] a dual-gas source polygeneration process which uses, besides syngas from coal gasification, coke oven gas as gas sources and co-produces dimethyl ether, methanol and dimethyl carbonate via an integrated catalytic synthesis procedure. The system performance is calculated by numerical simulations, based on detailed chemical kinetics which shows the system feasibility, whereas the carried out exergoeconomic analysis, reported the exergy loss and the production cost of product in each functional block of the process.
176 177 178
Natural gas Over the past few years, combined cooling heating and power (CCHP) systems fuelled by natural gas raised increasing interest [116]. In particular, liquefied natural gas (LNG) is recognized as the best fuel option
ACCEPTED MANUSCRIPT 6 of 67
179 180
in many countries, in terms of global environmental impact [115]. Therefore, CCHP systems fuelled by LNG were recognized as a potential energy system option for sustainable development and low-carbon society.
181 182 183 184 185 186 187 188
An example of this system was investigated by Li et al. [117], for hotels, offices and residential buildings by an energy, economic and environmental analysis. In the proposed polygeneration system, two prime movers are taken into account: a gas turbine and a gas engine. The produced heat is recovered to meet the thermal demands through the thermal exchanger and the cooling load through the absorption chiller ACH. Such system was compared to a reference system, where the space heating and cooling are produced by a boiler and electric chiller. The resulted biggest economic merits are received in offices, 39.21%, driven by gas turbine and 52.83% driven by gas engine. The biggest reduction of emissions and energy consumption are received in hotels, with a maximum value equal to 60.65% and 42.28%, respectively.
189 190 191
The concept of chemical looping combustion and gasification investigated in [91], is also included in a new polygeneration system simulated by Salkuyeh and Adams [58]. The simplified layout of the chemical looping gasification is reported in Figure 3.
192 193
Figure 3. Simplified schematic of the chemical looping gasification system [58].
194 195 196 197 198 199
With respect the previous work using coal, here natural gas and petroleum coke are used as input fuels and different products such as chemicals, olefins, electricity and transportation fuels are co-produced. All process unit operation models were simulated in Aspen Plus 2006.5 simulation software, except for the gas turbines for which a custom model in VBA (Visual Basic for Applications) was linked into the Aspen simulation model instead. The system is not only able to capture 100% of CO2 emissions, but also can be an incentive for those producers to convert stockpiled petroleum coke to more valuable products.
200 201 202 203 204 205 206 207 208
An innovative polygeneration system fuelled by natural gas for the production of methanol and electricity is investigated by Gao et al. [92]. The novelty regards the integration of partial-reforming and partial-recycle scheme (Figure 4) in methanol synthesis which contribute to enhance the system performance (energy and exergy) with respect to the one of full reforming and without recycle, assumed as reference system. In particular, in this new layout, only a part of natural gas entering the reformer will be converted into syngas, which will be cooled by heat recovery unit, and before will be sent to the methanol synthesis process. Here a partial-recycling operation is adopted, which expects that the unreacted syngas is recycled back to the synthesis reactor, and a part will be fed to the power generation subsystem as fuel, and the rest part will be burnt with the GT exhaust gas to cover the heat demand for the reforming process.
ACCEPTED MANUSCRIPT 7 of 67
209 210
Figure 4. Flow sheet of part reforming and part cycling (PR/PC) system [92].
211 212 213 214 215
The system simulations were carried out by using ASPEN PLUS steady state simulation software. In this work, the energy saving ratio is used to assess the polygeneration system energy and exergy performance. The energy results are summarised in Table 2. The reference system has rather bad performance with respect the innovative one which also able to reduce about 6% of energy consumption compared with the single product systems.
216
Table 2. Performance comparison between systems System
Reference
Fuel consumption (LHV, kW) Net work (kW) Methanol production, M (kg/h) Relative energy saving ratio (%) Exergy efficiency
69753 21211 -0.2 61.02
Innovative 315951 81350 21034 5.8 64.14
217 218 219 220
In Table 3, some of the works available in literature concerning polygeneration systems fuelled by natural gas and coal are listed. Whereas, in Table 4 shows several energy indexes (Fuel Energy Saving Ratio, Primary Energy Saving, Primary Energy Rate) and the consequent avoided CO2 emissions of the polygeneration systems fuelled by fossil fuels, when these are compared to the conventional separate production systems.
221
Table 3. Works on polygeneration systems fuelled by natural gas and coal. Work Li et al. [34] Lin et al. [17] Li et al. [89] Bose et al. [90] Fan et al. [91] Adams and Barton [118] Zhu et al. [119] Chiesa et al. [120, 121] Yi et al. [93] Khojasteh Salkuyeh and Adams [98] Sibilio [11]
Input
Soutullo [12]
Natural gas
Calise et al. [39] Arsalis et al. [40]
Technology CC, gasifier
Coal
GT, gasifier CC, gasifier GT, ST ICE Engine generator ICE GT
Output dimethyl ether, methanol and dimethyl carbonate methanol, power, CO2 methanol and power CO2, urea, hydrogen, power power, heating, cooling methanol, diesel, gasoline, power hydrogen, power hydrogen, power and CO2 methanol, power power, ethane and propane power, heat, cool H2 power, heat, cool power, heat, cooll
ACCEPTED MANUSCRIPT 8 of 67
Calise et al. [49] Khojasteh Salkuyeh and Adams [58] Calise et al. [73] Li et al. et al. [117] Tock and Maréchal [122] Rosato et al. [123] Adams and Barton [118] Rubio-Maya et al. [124, 125]
222
CC Reformer
chemicals, olefins, electricity and transportation fuels
ICE GT, ICE GT ICE Gasifier, GT ICE, GT, FC
power, heat, cool H2 power heat and CO2 power, heat, cool methanol, diesel, gasoline and power power, heat, cool and fresh water
Table 4. Economic and environmental indexes of the polygeneration systems fuelled by fossil fuels Value
ΔCO2
Conventional separate production
[13] [11] [14]
Energy index FESR1 PES2 PES
26.6 % 0.2 to 4.3 % 6.6 %
66.7 to 70.5 % 3.9 to 10.6 % 11 to 12 %
Boiler, electric chiller, grid
[17]
PES
8.1 to 15.6 %
52.8 to 88 %
[62] [48] [73] [91]
FESR PES PES FESR
0 to 20.0 % 14.0 to 24.0 % 14.0 to 14.3 % 23.4 to 27.2 %
134 tons/year -
[92]
PES
5.8 %
-
[112]
η
50 % (+11%)
-
[88] [126] [127]
PES PER3 PES
38 % 76. 5 % 20 %
1000 g/kWhe year -
[114]
-
-
225 %
[128]
PES
54 % (maximum)
-
Work
Boiler, air-cooled water electric chiller, grid Synthetic fuel production systems Power generation technologies Standard boiler Boiler, electric chiller, grid Single methanol production Single power generation system Ethanol from wood process Combined heating and electricity system Boiler, cooling absorption machines, grid Vacuum hot water boiler, grid Boiler, grid Single bio-fuel production Single electricity production Gas boiler, compression chiller, grid
223 224
1
Fuel Energy Saving Ratio (FESR): the ratio of the energy saving of the polygeneration system, in comparison with a conventional system with separate productions, to the overall energy consumption of the reference system. 2 Primary Energy Saving (PES): the ratio of the difference between the primary energy of the reference system (separate production) and the primary energy of the polygeneration system, to the primary energy of the reference system. 3
Primary Energy Rate (PER) is defined as the ratio of the primary energy demand to the required output. The system with the lowest value of PER is considered the best with regards to energy consumption.
ACCEPTED MANUSCRIPT 9 of 67
225 226 227 228 229 230 231 232
3.2 Renewable energies During the past few years, renewable energy sources represented an attractive alternative to fossil fuels, due to their potential contribution to fighting global warming, reducing the emission of pollutants in general, slowing down the depletion of fossil fuels, etc. [129], in order to promote a "sustainable" development [23]. Dozens of papers investigated polygeneration systems supplied by renewable energy sources (biomass, geothermal, solar, wind, hydro etc.); so, a wide number of different system layouts was proposed. The present section provides an overview of those systems classified on the basis of the renewable energy source used. Biomass
233 234 235 236
In many cases, biomass-based polygeneration systems use cheap biomass as fuels. Biomass is especially promising in such systems by allowing one to produce a plurality of by-products, along with the conventional energy vectors (electricity, cool and heat). In addition, the possible use of agricultural wastes is extremely promising, since many of these have good potential to be energy resource [130].
237 238 239 240 241
An example of this configuration is presented by Jana and De [15, 24]. They used rice straw, sugarcane bagasse and coconut fibre dust. The polygeneration system (Figure 5) is designed as follows: in a fluidized bed gasifier, the agricultural waste is gasified in syngas, which cleaned, cooled and compressed is utilized for power generation in a combined cycle. An absorption chiller, ACH, is supplied by the heat generated in syngas cooler, whereas an ethanol production unit is supplied by a fraction of the whole generated syngas.
242 243
Figure 5. Schematic of the polygeneration process [15].
244 245 246 247 248 249 250
The whole system was modelled in Aspen and a detailed economic analysis was also developed. In Figure 6, a comparison of net power produced for three different agricultural wastes are shown. Power output is contributed mostly by the gas turbine plant, then, by the low pressure steam turbine and the rest by the high pressure steam turbine. It is noted that net power output for coconut fibre dust is the maximum, and then for sugarcane bagasse followed by that for rice straw. The results of the economic analysis showed that the proposed system is very interesting, specifically for rural communities, by achieving a payback period just above 5 years, when the rice straw biomass is taken into account.
ACCEPTED MANUSCRIPT 10 of 67
251 252
Figure 6. Power outputs for different agricultural wastes per t/year [15].
253 254 255 256 257 258
The same system layout displayed in Figure 5 was proposed by the authors to evaluate the system performance from the techno-economic point of view, by focusing only on rice straw energy source [94]. Here, the analysis is specific for a district in the state of West Bengal of India. Results show that the polygeneration plant using rice straw as energy resource is a feasible option in energy services for developing countries like India. In fact, the estimated payback period of the plant is 4.76 years. The return on investment (ROI) is 15.5 % and the net present value is 97 million USD.
259 260 261 262 263 264 265 266
Based on the results of the carried out simulations in the previous work, a Life cycle analysis (LCA) was developed in ref. [27], where the environmental impact of this polygeneration plant is discussed. Here, the polygeneration system is compared with stand-alone conventional plants with same utilities. This study helps to estimate the life cycle emission, specifically greenhouse gas emission of this plant. Exergy based allocation method is used for this analysis. Results indicate that global warming potential of the standalone generations of same utilities in conventional ways is 100 times more than corresponding straw based polygeneration. Logistics of biomass cause maximum environmental impacts out of the unit processes of the polygeneration. It shares almost 50% of the global warming potential and the acidification potential of this polygeneration.
267 268 269 270
A biofuel-fired trigeneration systems equipped with electrical and thermal energy storages for remote households was simulated by Huang et al. [107]. In particular, in this study wood pellets and willow chips are chosen to supply a Stirling engine and to evaluate the internal combustion engine (ICE) performance, biodiesel is also selected (Figure 7). A thermally-driven chiller is also included.
271 272
Figure 7. Schematic diagram of the trigeneration system, ICE (left) and Stirling engine (right) [107].
273 274 275 276 277 278
The ECLIPSE process simulation package, successfully used to analyse a wide range of energy conversion systems, such as coal fired power plants and biomass energy systems, was adopted to perform the simulations. In Table 5, some of the main energy results of simulations are reported. From the economic point of view, calculations showed that the economic feasibility of this type of plant, evaluated by the breakeven electricity selling price (BESP) is affected on the renewable energy incentive schemes (Renewable Heat Incentive (RHI), for 7 years at a rate of 12.2 p/kWh. and Feed-In Tariffs (FITs) for 10 years at a rate of 13.2 p/kWh.). If the plant
ACCEPTED MANUSCRIPT 11 of 67
279 280
obtains both incentives, the BESP for wood pellets, willow chips and biodiesel are 216, 169 and 310 £/MWh, respectively.
281
Table 5. Technical results. Feedstock Electrical output, kWe (Net) Overall electricity efficiency, % Heat output, kWth (maximum) Cooling output, kWth Overall CHP efficiency, % Overall trigeneration efficiency, % CO2 emissions, g/kWh (CHP) Reduction in CO2 emissions, t/year
282 283 284 285 286 287 288 289 290 291 292
Willow chips
Bio-diesel
3.2 19.3 10.9 4.4 84.3 66.9 409 8.7
3.2 18.8 11.1 4.5 82.4 65.4 413 8.9
5.2 30 7.6 2.2 74.6 64.1 433 6.1
The performance of system for direct conversion of biomass (sawdust) through chemical looping for coproduction of hydrogen and power was evaluated in ref. [108]. A net power output ranging from 400 to 50 MW and a flexible hydrogen output in the range of 0 to 200 MWth were taken into account. A CC is adopted as prime mover. ChemCAD software is used to model and simulate the several investigated plant configurations, including layouts with biomass and / or coal gasification, with and without carbon capture and biomass direct chemical looping. Authors state that the reason to use coal and biomass is due to the fact that currently there are no industrial size gasifiers able to process only biomass. For these gasifiers, the biomass ratio can be up to 30% from total fuel input. Results regarding power and hydrogen production obtained for biomass direct chemical looping case are summarised in Table 6. From the economic point of view, it resulted that the biomass direct chemical looping has a reduced carbon capture capital cost penalty than the gas–liquid absorption carbon capture case (7% points lower specific investment cost).
293
294 295 296 297 298 299 300 301 302
Wood pellets
Table 6. Energy results of biomass chemical looping case. Main plant data
Units
Power only
Sawdust flow rate Gross electric power output Hydrogen output — LHV Gross electrical efficiency Hydrogen efficiency Carbon capture rate
t/h MWe MWth % % %
558.19 0 46.59 0 99.6
Hydrogen and power cogeneration 268.6 500.31 441.52 100 200 41.76 36.85 8.34 16.69 99.6 99.6
A further sector where the concept of polygeneration was often adopted concerns the co-production of power and fuels via bioorganic wastes. An example of this application is the work reported in reference [110]. Here the multi-generation of different fuels such as DME, methanol, ethanol, FT fuels, fertilizers and biogas from biomass wastes is investigated. The layout of this production process is depicted in Figure 8 and it is based on anaerobic digestion and dry reforming processes. In particular, accurate models for production of biogas are written in order to compute the optimal mixture of biomass wastes between cattle and pig slurry, cattle and pig manure, sludge, urban food waste and urban green waste to be digested to obtain the required biogas. The models are written in GAMS® and in order to identify the optimal solution, corresponding to the optimal mixture which maximize the amount of methane, CONOPT tool was used.
303 304
Figure 8. Scheme of biomass waste processing to fuels and chemicals.
ACCEPTED MANUSCRIPT 12 of 67
305 306 307 308 309
It resulted that for a syngas optimal production, achieved by biogas dry reforming, a mixture of biogas consisting of 50% CH4, and 48% CO2 is the most appropriate. When the biogas is intended as fuel, a 70% methane content is targeted. In addition, they state that if the aim is obtaining the most typical fertilizer composition, 65% of cattle slurry and 35% urban food waste is the optimal blend. Solar energy
310 311 312 313 314
Several papers also investigated the integration of polygeneration systems with solar thermal collectors and PV panels. In fact, in polygeneration plants fuelled by solar energy, solar technologies can convert the solar incident irradiation in thermal energy used for solar power plant in case of medium or high temperature heat production or, conversely, for space heating or DHW (by STC), in electricity (by PV) and in cooling energy, by thermally driven chillers or electric vapour compression chillers.
315 316 317 318
In this framework, an interesting work is reported in reference [57], where a solar trigeneration system, based on sheet and tube PVT solar collectors, a single-stage LiBr/H2O absorption chiller, ACH, storage tanks and auxiliary heaters (Figure 9), is dynamically simulated by means of a dynamic simulation model, developed with TRNSYS.
319 320 321 322
In this work [57], an example of integration of solar heating and cooling system with low temperature PVT collectors is reported. System producing electricity, space heating and cooling and domestic hot water for a university building located in Naples (Italy). The system performance is analysed from both energy and economic points of view.
323 324
Figure 9. System layout [57].
325 326 327 328 329 330 331 332
It resulted that the polygeneration system performance is mainly affected on PVT performance that is excellent during the summer, decreasing significantly in winter. The amount of PVT heat used for DHW is 60.7%; the electrical solar fraction is satisfactory since the PVT can cover about 65% of the overall electricity demand. The PVT average electrical efficiency is slightly lower than 10%. This value is not so bad, considering that the nominal efficiency at 25 °C was 16% and taking into account that PVT operates at 50°C and 80°C respectively in winter and in summer. The overall primary energy saving (PES) achieved by the whole system is slightly higher than 70%. The best resulted pay-back period, by the carried out economic analysis, is 3.6 years, when a proper feed-in tariff is considered.
333 334 335 336 337 338 339 340
An example of application of low temperature PVT collectors coupled to a solar-assisted heat pump and an adsorption chiller in a novel polygeneration system based is reported in reference [41]. Here, the dynamic simulation model and a thermo-economic analysis is carried out. The plant produces electricity, space heating and cooling and DHW for a small residential building. During the winter, thermal energy of PVT collectors primarily supplies the heat pump evaporator, whereas in summer, supplies the adsorption chiller (usually activated for lower hot fluid temperatures) providing the required space cooling (Figure 10). The results showed a total energy efficiency of the PVT of 49%, a heat pump yearly coefficient of performance (COP) for heating mode above 4 and a COP of adsorption chiller of 0.55.
ACCEPTED MANUSCRIPT 13 of 67
341 342
Figure 10. Simplified layout of the proposed system including the main components and loops [41].
343 344 345 346 347 348 349 350 351 352 353 354 355
In reference [86] the exergy modelling of a new solar driven trigeneration system was performed. With respect the previous systems, here high temperature PTC collectors are coupled to a high critical temperature efficient ORC, with the n-octane as working fluid. A single-effect ACH from production of cooling energy is also included (Figure 11). Both heating (steam water) and cooling production are achieved by the thermal recovery from the ORC. Different operating conditions, namely power generation, power-heating and powercooling cogeneration and trigeneration were analysed. In particular, by an appropriate integration of heat exchangers and thermal storages, three different modes of operation were taken into account depending on low or high availability of solar radiation and operation during the night time. The exergy analysis was carried out by varying the temperature difference at the pinch point of the ORC evaporator, the inlet temperature of the ORC pump and the turbine inlet pressure. The maximum electrical-exergy efficiency ranged between 7.00% and 3.00% depending the operating condition. In case of trigeneration, global exergy efficiency ranged between 20.0% and 7.00%. Finally, most of the exergy destruction rate occurred in the solar collectors and ORC evaporators.
356 357 358
Figure 11. Schematic of the solar-trigeneration system [86] .
ACCEPTED MANUSCRIPT 14 of 67
359
Biomass and solar energy
360 361 362 363 364 365 366 367 368 369 370 371 372
Several studies examined different configurations of polygeneration plants consisted of both solar (STC, PV and PVT) and biomass technologies. An efficient approach to simultaneously use biomass and solar sources is investigated by Calise et al. [57]. Here, a reciprocating engine, fed by rapeseed oil, was coupled to concentrating parabolic trough solar collectors (PTC) to produce thermal energy and a double-stage LiBr/H2O ACH to produce cooling energy. In the framework of polygeneration systems, this type of plant that integrates reciprocating engines with solar thermal collectors is scarcely investigated [65]. In particular, the solar and biomass source are linked because exhaust gases produced by biomass fuelled reciprocating engine are used to provide additional heat to the fluid heated by PTC. In this study, the whole trigeneration system was modelled and dynamically simulated in transient system simulation tool (TRNSYS). The economic results show that the system under investigation is profitable, especially if properly funded and from the energy point of view the rapeseed oil reciprocating engine is suitable for integration in polygeneration systems based on high temperature solar heating and cooling systems, achieving a PES higher than 93%.
373 374 375 376 377 378 379 380 381 382 383 384 385
PTC are coupled to a willow pellets boiler in a tri/co-generation plant, consisted of an ORC–Vapour Compression Cycle (VCC) systems in the energy–exergy analysis and economic investigation performed by Karellas and Braimakis [10]. Here, ORC produces power consumed partially by VCC compressor, activated only in summer, whereas the heat generated in the condenser is used to meet hot water demand. With respect the previous work, here PTC and boiler are used to provide the request heat of ORC (Figure 12), and the cooling energy is produced by VCC system. The resulted net electric efficiency is 2.38%; maximized thermal efficiency of the ORC is equal to 5.5% when the evaporation temperature is equal to 90°, while the exergy efficiency is 7% when the biomass boiler operates at full load operation. Findings suggested that superheating process and the adoption of a recuperator lead to higher capacities of the heat exchanger areas and of the solar field extensions, without returning any significant increase of the system efficiency. By considering a typical apartment block on a Greek Island, assuming PTC area of 50 m2 a payback period of 7 years is achieved. A sensitivity analysis was carried out by varying the investment cost and the biomass price, analysing their impact on the IRR and SPB.
386 387
Figure 12. The combined ORC–VCC cycle and biomass boiler and PTC heating water circuits [10].
388 389 390 391 392 393 394 395 396
A wood chips gasification system and solar evacuated collectors are coupled in the simulation work carried out by Wang and Yang [25]. Here, an ICE supplied by the produced gas by gasifier and a dual-source powered mixed-effect ACH are also included. Waste heat of ICE and produced heat of solar field are utilized to produce cooling and heating energy. The plant is simulated by using the EES software package, and a comparison of the simulated linear curve of the instantaneous efficiency of the evacuated tube collectors between the simulation data and the experimental one, achieved by the solar energy research institute of Beijing is performed. The results indicate that the primary energy ratio and the exergy efficiency are 57.9% and 16.1%, respectively, and the carbon emission reduction ratio is about 95.7%, at the design condition.
ACCEPTED MANUSCRIPT 15 of 67
397 398 399 400 401
A polygeneration system of generating methanol and power with the solar thermal gasification of the cotton stalk is proposed in ref. [59]. A field of heliostats concentrates the solar radiation on the solar tower where the biomass gasification occurs. Here, the high-temperature solar source is directly employed in the gasification process and, therefore, is considered a promising option for the supply of the process heat (Figure 13).
402 403
Figure 13. Schematic diagram of the proposed polygeneration system [59].
404 405 406 407 408 409
The syngas from the biomass gasification is used to produce the methanol via a synthesis reactor. The unreacted gas is used for the power generation via a CC power plant. The solar assisted polygeneration system is located in Yanqi (China) and is numerically simulated by the Aspen Plus software to estimate the polygeneration system thermal performances. The highest energy efficiency and the exergy efficiency of the polygeneration system approximately are 56.09% and 54.86%, respectively.
410 411 412 413 414 415 416
A hybrid solar-biomass polygeneration system, based on PTC, for the simultaneous production of power, cooling by vapour absorption refrigeration (VAR) and fresh water by a multi-effect humidification and dehumidification (MEHD) desalination unit is investigated by Sahoo et al. [77]. The VAR cooling system operates using the extracted heat taken from steam turbine and condenser heat of the VAR cooling system is used in desalination system for production of drinking water. The system PES is 50.5% and the energy output is increased to 78.12% from this system as compared to simple power plant.
417 418
In Table 7, some of the studies available in literature concerning hybrid solar-biomass polygeneration are listed.
419
Table 7. Works on hybrid solar-biomass polygeneration systems. Work Calise et al. [2] Calise [66] Sahoo et al. [131] Angrisani et al. [132] Rubio-Maya et al. [124] [125] Ray et al. [101] Pantaleo et al. [133] Vidal and Martin [134] Khalid et al. [78]
Solar technology CPVT PTC PTC Scheffler concentrator
Biomass technology Wood-chip auxiliary heater Wood-chip auxiliary heater Boiler Wood pellet fluidized Bed
Prime mover GT, ST Stirling engine (STE)
Evacuated tube collectors
Crops gasifier
ICE, GT, MGT, FC, STE
PV PTC Solar tower
Straw, wood gasifier Wood chips boiler Lignocellulosic gasifier Dry Olive pits combustion chamber
Gas engine GT, ORC GT, ST
Heliostat tower
GT, ORC
ACCEPTED MANUSCRIPT 16 of 67
Ghaith and Abusitta [111]
420
flat plate and evacuated tube collectors
Pellet auxiliary heater
-
Hybrid geothermal and solar systems
421 422 423 424 425 426
Integration of geothermal and solar energy is one of the most investigated configurations. In fact, lowmedium enthalpy sources, as the solar and geothermal ones, represent an interesting input for Thermally Activated Technologies (TAT), or Thermally Driven Technology (TDT), such as multi- effect-distillation (MED) systems, as well as Organic Rankine Cycles (ORC) and absorption chillers (ACH) [39]. Therefore, numerous studies can be found in literature about the investigation of different layout schemes, and subsequently some of them are reported.
427 428 429 430 431 432 433 434 435 436 437 438 439 440 441
A novel trigeneration system obtained by new prototypal flat-plate evacuated solar collectors coupled to a geothermal well is dynamically simulated in the work of Buonomano et al. [18]. The whole system, designed for a hotel building in Ischia Island (Mediterranean Sea), includes a new simulation model for ORC, and produces power, DHW and cooling by a single stage LiBr/H2O ACH driven by geothermal energy only (Figure 14). The ORC is supplied by heat obtained by a geothermal well in which geothermal brine is about at 95°C, but in order to improve system performance, additional heat is provided by solar energy obtained through a 25 m2 solar field. In fact, it resulted that the majority of the thermal energy supplied to the ORC is due to the contributions of geothermal source which is dominant over the solar one. The efficiency of solar collectors, ORC and ACH are 59.2%, 6.4% and 0.68, respectively. Here, a sensitivity analysis in order to investigate the effects of the variation of groundwater temperature, from 90 to 100°C is performed (Figure 15): the higher the well geothermal brine temperature, the higher the ORC inlet temperature. This leads to an increase in the ORC output electricity Eel,ORC, due to the improvement of ORC efficiency, ηORC, from 6.4% to 7.2% and to a decrease of the energy supplied by the auxiliary heater, Eth,AH, since the geothermal source is able to achieve the ORC minimum inlet temperature. Conversely, the produced cooling energy, Ecool,ACH, is not significantly affected by the variation of well temperature.
442
443 444
Figure 14. Sketch of the system layout [18].
ACCEPTED MANUSCRIPT 17 of 67
445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460
Figure 15. Sensitivity analysis: energy, efficiency vs. well geothermal brine temperature [18]. A hybrid solar and geothermal polygeneration system combining CPVT field, a single-stage LiBr/H2O ACH and a MED unit was studied through dynamic simulations in [3]. The plant provides electrical, thermal and cooling energy, DHW and a quantity of desalinated water able to cover the whole request of the Pantelleria Island, assumed as case study. In particular, as depicted in Figure 16, solar thermal energy, at a maximum temperature of about 100 °C, in combination with the thermal energy produced by low-enthalpy (about 80 °C) geothermal wells (GW), is used to supply the MED system. Geothermal energy is also used to produce DHW at 45 °C. The performance predicted through dynamic simulations, supported by appropriate economic models, was excellent. This work is particularly interesting because it presents a comprehensive assessment on combination of different kinds of technologies (SHC, CPVT, GW and MED). Several additional islands in the Italian Mediterranean Sea, Ischia and Aeolian Islands, show a similar potential in terms of availability of geothermal and solar energy are also investigated. For all the investigated locations, a parametric analysis aiming at evaluating the variation of Profit Index as a function of the ratio of DHW produced by the system (QHE4) and demanded by the user is performed. It resulted that Profit Index dramatically decreases in case of scarce DHW demand.
461
462 463
Figure 16. System layout [3].
464 465 466 467 468
Other studies concerning the dynamic simulation of a novel hybrid solar and geothermal polygeneration system capable to produce electric energy, fresh water and space heating and cooling, for the Pantelleria Island was presented and discussed in references [16] and [37]. Here, with respect the previous layout, an ORC supplied geothermal and solar energy is considered. In particular, in reference [3], a low-temperature geothermal well (85°C) is used, mainly for the desalinization process, whereas here a medium-enthalpy
ACCEPTED MANUSCRIPT 18 of 67
469 470 471
geothermal source (160°C) is considered. In addition, in this work, the CPVT collectors are replaced by PTC field coupled with a thermal storage tank (Figure 17). The MED unit and ORC were developed in Engineering Equation Solver (EES).
472 473
Figure 17. System layout [16, 37].
474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490
From the energy analysis it resulted that the solar energy input is much lower than the geothermal one, as highlighted by the low value found for the solar fraction, 9.60%; the ORC efficiency is about 11.6%. During the year the MED unit is able to produce fresh water, equal to 54% of the total seawater flowrate. An accurate energy, economic, exergy and exergoeconomic analysis of the system was performed, from which it was found that the global exergy efficiency varies from 40% to 50% during the thermal mode and from 16% to 20% during the cooling one [37]; besides, the exergoeconomic costs of electricity, cooling water, chilled water and desalinated water resulted very interesting, respectively in the ranges 0.1475–0.1722 €/kWh, 0.01612–0.01702 €/kWhex, 0.1863–0.1888 €/kWhex and 0.5695–0.6023 €/kWhex. The previous works [16, 37] are developed by considering that all the useful products are consumed locally. Instead, in a further study [39], based on a similar layout, the system is supposed to be connected to a district electrical, heating and cooling network. In this case, the energy provided by the system must match the real time-dependent demands of electricity and space heating and cooling of typical buildings of Pantelleria Island (assumed as a case study). The system achieved a SPB equal to 8.50 years, with a potential primary source saving of 37.5 GWh/year and a potential CO2 avoided emission of 9451 tons/year. It resulted capable to cover the energy demands of 800 examined buildings. Moreover, the plant produces 1006 103 m3 of desalinated water and it is capable to cover the fresh water global demand. Average ORC efficiency amounts to 15.30%; the system is mainly powered by geothermal energy in fact, the average solar fraction is only 14.6%.
491 492 493 494 495 496 497
In reference [79], the authors proposed an interesting renewable multi-generation system, which integrates geothermal energy and solar air PVT system, produces electricity, hot water, drying air and capable to perform space heating and cooling. Power is provided by the PVT system and by the ORC; this latter is powered only by medium–high geothermal source. Cold water is provided by a LiBr/H2O ACH, whose thermal input at the generator is represented by the ORC waste heat. The cooling output is used by a dairy farm, in which heating is provided by an electric heat pump. Ambient hot air coming from the solar PVT system is conditioned for the food drying process in the farm. The overall energy and exergy efficiencies of the multi-generation system are
ACCEPTED MANUSCRIPT 19 of 67
498 499 500
calculated to be 11% and 28%, respectively. The exergy efficiency of the solar system is around 12%. The COP and the exergy efficiency of the ACH are 0.73 and 0.21, respectively. COP and exergy efficiency of heat pumps were 4.1 and 0.03, respectively. For the ORC, the energy and exergy efficiencies are 9% and 42%, respectively.
501 502 503 504 505 506 507 508 509 510 511
In reference [80], two novel ORC-based CHP systems powered by solar and low-temperature geothermal (90 °C) sources are presented and compared. Both the configurations are designed to provide 50 kWel, 400 kWth and in both cases three working fluids (e.g. R134a, R236fa, and R245fa) were considered. In particular, one configuration is based on evacuated tube collectors and the power is produced by a single turbine, while the second one is based on a double stage system, in which a field of evacuated solar collectors heats the working fluid up to an intermediate temperature. After this first stage, only a part of the working fluid is heated in a second solar field, composed of direct-steam PTCs, up to the maximum temperature of the cycle. Mechanical work is then produced by two turbo-expanders arranged in series. In both configurations, the best performance in terms of first law and exergy efficiency was obtained with R245fa (Table 8); the best performance in terms of heat recovered was obtained with R134a. Moreover, findings suggested that single-pressure configurations show a better performance, at design conditions.
512
Table 8. Results: efficiencies for first and second configuration for three working fluids. Fluid
R245fa
Configuration System efficiency (%) Cycle efficiency (%) Exergy efficiency (%)
513
Second 6.9 8.9 17.5
R245fa First 9.78 11.3 23.3
Second 7.4 9.5 18.4
R245fa First 13.0 15.1 25.0
Second 10.0 12.7 20.0
In Table 9, some of the works concerning hybrid solar-geothermal polygeneration systems are listed.
514
Table 9. Works on hybrid solar-geothermal polygeneration systems.
Calise et al. [3]
Solar technology CPVT
Geothermal temperature 80°C
Calise et al. [16, 37]
PTC
160°C
Work
Suleman et al. [81]
flat-plate evacuated PTC flat-plate evacuated, PTC PTC
Al-Ali and Dincer [82]
PTC
190°C
Islam and Dincer [83]
PTC
167°C
Buonomano et al. [18] Calise et al. [39] Tempesti et al. [80]
515
First 9.1 10.5 22.7
Prime mover -
Power, heat, cool, fresh water
95°C 160°C 80°C-100°C 110°C
Output
Power, heat, cool ORC
Power, heat, cool, fresh water Power, heat Power, heat, cool, dry wet products Power, cool, heat, industrial process heat Power, heat, cool, dry air
3.3 Mixed energy inputs
516 517 518 519 520 521 522 523
One of the most promising peculiarities of the polygeneration system lies in the possibility of hybridizing existing technologies, based on fossil fuels, with renewable energy sources. Several works available in literature on polygeneration systems, present the integration of renewable technologies (solar, geothermal, wind, etc.) into different conventional systems (trigeneration, DHC, gas-fired boiler, electric heat pump) or are based on the coupling of CCHP and renewable energy systems in a plurality of applications [135]. The main potential applications of hybrid polygeneration systems are represented by hospitals, which have large and constant cooling and thermal loads, but also hotels, residential districts, schools, commercial buildings and so on [136].
524 525 526
In this framework, a theoretical analysis focused on the feasibility of a hybrid plant applied to a real hospital located in Ferrara, Italy, is presented in reference [137]. Several hybrid schemes were investigated and compared: i) the conventional system consisting of electric grid, gas-fired boilers and compression chiller; ii)
ACCEPTED MANUSCRIPT 20 of 67
527 528
PAFCs (phosporic acid fuel cells) integrated with ACHs and gas-fired boilers; iii) solar thermal collectors coupled to Rankine power cycle and ACHs; iv) PV panels coupled to conventional systems.
529 530 531 532 533
Basrawi et al. [26] investigated a hybrid energy system based on a polycrystalline PV and MGT (Figure 18) system able to produce power, heating and cooling of a group of residential buildings under a tropical region. Some of the electricity produced can be stored to the battery. Exhaust heat of the MGT is recovered by the heat exchanger, and it is used to cover water heating demand of the houses, and the rest is supplied to the absorption chiller. The system was compared with another system without PV.
534 535
Figure 18. Configuration of the overall hybrid energy system for all cases [26].
536 537 538 539 540 541
It resulted that when the proposed system operates with power-match operation strategy, the highest Net Profit is obtained. This circumstance occurs mainly because in this case no battery system is needed, and the related high cost can be saved. However, this system presents low environmental performance especially if compared to a combined cycle gas turbine, because larger MGTs in this system operated frequently under partial load. Operation strategy that had smaller MGT that operated under full load can still generate Net Profit but at lower degree because it needs larger PV and battery.
542 543 544 545 546 547 548 549
In reference [67], an existing trigeneration system was hybridized with a solar field equipped with concentrating photovoltaic thermal (CPVT) collectors. The system includes a gas turbine, CPVT collectors, ACHs, tanks and balance of the plant (BOP) devices. The system supplies electricity, heat and cooling and was designed and dynamically simulated to be installed at the District Hospital of Naples, (South of Italy). Energy and economical parameters were also evaluated. The system profitability was found to be acceptable (pay-back period equal to 15 years), even in the case of no public funding. However, if a feed-in tariff of 0.30 €/kWh (similar to the one adopted in Italy for PV systems), the SPB period becomes 6.1 years. Considering also a possible feed-in tariff of 0.10 €per kWh of thermal energy produced by the CPVT, the SPB becomes 4.1 years.
550 551 552 553 554
A polygeneration system including renewable energy sources and a natural gas engine, for residential and tertiary buildings, was analysed by Soutullo et al. [12]. The system, modelled in TRNSYS, includes: PV panels, flat plate STC, biomass boilers, wind turbines, polymer electrolyte membrane fuel cells (PEMFCs), ORC (Figure 19) and a single-effect ACH. A natural gas engine is used as a backup when the renewable technologies cannot supply the thermal district loads. The “design of experiments” methodology was used.
ACCEPTED MANUSCRIPT 21 of 67
555 556
Figure 19. Blocks diagram of the polygeneration plant [12].
557 558 559 560 561 562 563
The results of the dynamic simulations were used to analyse the performance of centralized polygeneration system in three representative buildings, located in different Spanish cities (Oviedo, Seville and Zamora). In the thermal model, two cases (Case 1 that maximizes the operation of the solar thermal field and Case 2 that minimizes the power of the biomass boilers) were evaluated, with renewable fractions equal to 90% and 50%, respectively. In both cases, Seville obtained the lowest solar contribution due to the highest cooling district demand. In the electrical model, three configurations were evaluated: PV + PEM, wind + PEM and PV + Wind + PEM. The last configuration achieved the best results.
564 565 566 567 568 569 570 571 572 573
An interesting work was presented by Arsalis et al. [40], reporting a thermoeconomic analysis of a 1 MW CCHP system driven by liquefied natural gas (LNG) and assisted by PV. The system includes a gas turbine cycle, ACHs, heat exchangers, LNG storage, a PV solar field and BOP components (pumps, valves, etc.); it is designed to serve a complex of households in Nicosia, Cyprus. The system was modelled in EES and simulated at both part-load and full-load conditions. The simulations results, performed on an annual basis, showed that the primary energy ratio is almost constant at all load condition and that the CCHP can completely fulfil the load profile of 436 households. In addition, without PV integration, the system must generate an additional 1959 MWh/year of electricity by LNG conversion. The economic analysis shows a minimum of the life cycle cost (LCC) around a PV capacity of 300 kW.
574 575 576 577 578 579 580 581
The energy and economic feasibility of a solar-assisted heating and cooling system and DHW production for a school building in different Italian climates was assessed by Calise [68]. Here, evacuated solar collectors, a single-stage LiBr/H2O ACH, and a conventional electric-driven reversible heat pump are simulated by using the software TRNSYS. From the energy point of view, results are encouraging, for the potential of energy saving: 64.7%, 52.4% and 61.4% for Naples, Milan and Trapani, respectively. On the contrary, the economic profitability can be achieved only in case of public funding policies (e.g. feed-in tariffs), as always happens for the great majority of renewable energy systems, since the results SPB are 12.1, 15.7 and 12 years for Naples, Milan and Trapani, respectively.
582 583
Some of the studies regarding the simulation of hybrid polygeneration systems, fuelled by fossil and renewable energy sources are listed in Table 10.
584
Table 10. Works on hybrid polygeneration systems. Work
Renewable technology
Fossil fuel
Prime mover
Output
Basrawi et al. [26]
PV
Natural gas
MGT
Power, heat, cool
Calise [19]
PTC
Soutullo et al. [12]
PV, flat plate STC, biomass boilers, wind
Methane, natural gas Natural gas
SOFC PEMFC, ORC
Power, heat, cool, hydrogen Power, heat, cool
ACCEPTED MANUSCRIPT 22 of 67
turbines, Arsalis et al. [40] Marrasso et al. [42] Calise et al.[49]
PV Evacuated tube collectors High-vacuum flatplate collectors
Kieffer et al. [138]
GT
Natural gas
ICE
Natural gas
CC
Natural gas / municipal solid waste Coal, Natural gas
CC
Power, transportation fuel
GT, ST, FC
hydrogen, power, heat and captured CO2
Ondeck et al. [135]
Biomass (wood) gasifier Evacuated tube Collectors, Biomass (crops) gasifier PV
Bizzarri et al.[137]
PV, STC
Natural gas
Buonomano et al. [67]
CPVT Evacuated tube collectors
Natural gas
ICE, GT, MGT, FC, STE GT phosporic acid fuel cells, ST GT
Methane
FC
Tock and Maréchal [122] Rubio-Maya et al. [124, 125]
Calise et al. et al. [74]
585
-
Liquefied natural gas
Coal, Natural gas Natural gas
Meerman et al. [96]
Gasifier (Oil residues, and wood)
Coal
GT, CC
Ng et al. [97]
Bio-oil gasifier
Natural gas, coal, CO2
CC, GT
Ozturk and Dincer [84]
Solar power tower
Coal
CC, FC, ST
Power, heat, cool and fresh water
Power, heat, cool
hydrogen, electricity, FTliquids, methanol and urea Power, hydrogen, methanol, methane, acetic acid Power, heat, cool, hydrogen, oxygen
4. Conversion technologies
586 587 588
Prime movers, converting a fuel into power and heat, usually represents the main components of a polygeneration plant. They can be selected according to different criteria, including input fuels, capacity, investment and operating cost, availability, etc. [139].
589
4.1 Gas turbine (GT)
590 591 592 593 594 595 596 597 598 599 600
The use of gas turbines in CCHP systems is investigated in several researches [140-142]. Usually, in such systems, high-temperature flue gas from the turbine are used for heating purposes; if an absorption chiller, ACH, is included, the waste heat can be also used for cooling purposes. An example of this configuration is reported in the work of Ziher and Poredos [140], who evaluated from an economic point of view a trigeneration system in the Slovenian biggest hospital (Figure 20), with a peakload of 4.2 MW. In particular, the main cogeneration unit is a GT, used to cover the high needs of electrical and thermal energy; gas-fuelled reciprocating engines are also used to provide up to 3 MW of power; over this limit, a GT is used. The steam generated in a heat-recovery steam generator, HRSG, is used for heating and sanitary hot water. A hot-water gas boiler provides additional heat during the winter period. A two-stage absorption chiller with a capacity of 2 MW is included, along with two back-up electric chillers with a capacity of 2 MW and 1.5 MW, respectively.
ACCEPTED MANUSCRIPT 23 of 67
601 602
Figure 20. The concept of energy supply with a trigeneration system [140].
603 604 605
In this work, the author focused mainly on the following aspects: i) economic analysis (Table 9), evaluating payback period (static method), net present value and profitability index (dynamic method); ii) optimization of the cooling production (with and without cold storage).
606
Table 11. Economic evaluation of the trigeneration system analysed in Ref. [112]. Static method
Trigeneration without cold storage Trigeneration with cold storage
Dynamic method
Payback period (years)
Net present value (Mill. Euro)
Profitability index (–)
6.71
2.02
1.07
5.86
2.72
1.09
607 608 609 610 611
Ghaebi et al. [141] performed an energy, exergy and thermoeconomic analysis of a CCHP system with gas turbine. Here, the GT flue gases are used in a dual-pressure HRSG to generate low and high pressure steam; the first one is supplied to an ACH, the second one is used directly. The CCHP has a maximum cooling capacity of 6.96 MW; the thermal capacity is 24.65 MW, the maximum power output is 19.23 MW; first and second law efficiency were 84.01% and 43.92%, respectively.
612 613 614 615
A similar configuration is analysed in reference [142]. The thermodynamic performance of such system was assessed in terms of energy efficiency, electrical-to-thermal energy ratio and exergy efficiency. It was found that the maximum exergy destruction is due to the combustion and steam generation processes, representing over 80% of the overall exergy destruction.
616 617 618 619 620 621 622 623 624 625
In case of small power applications, micro-gas turbines (MGT) can represent an interesting options [85]. In particular, MGT are sometimes preferred to other small-scale prime movers, such as reciprocating engines, mainly due to their good overall efficiency, low maintenance requirements and low emissions [143]. In this framework, the combination of MGT and ACHs in CCHP systems is often investigated. An example is reported in reference [75] (Figure 21); here, a real system, installed in a supermarket, is analysed from both experimental and numerical points of view. The system is based on a 80 kWe recuperated model (MTG 80RC-G), with a built-in hot water heat exchanger. Three MTG units and several absorber chillers are included. The system was simulated in TRNSYS, and the model was validated by comparison with experimental test results. Finally, the model was used to predict the system performance at different operating conditions, varying ambient temperature, fuel flow rate and pressure ratio, etc.
ACCEPTED MANUSCRIPT 24 of 67
626 627
Figure 21. Schematic diagram of the tri-generation test facility [75].
628 629 630 631 632 633 634 635 636
In particular, the Primary Energy Rate (PER) of the system is evaluated, defined as the ratio of the primary energy input (i.e.: the net fuel consumed) to the overall required output (i.e. the sum of the net power generation and cooling capacity). Therefore, the system with the lowest value of PER is considered the best with respect to energy consumption. The results show an interesting trend of PER as a function of ambient air temperature: since both the power generation and cooling capacity decrease in case of a higher ambient air temperature, the PER value, as a result, increases with the rise in ambient air temperature. By varying the pressure ratio from 1.5 to 7.0 (Figure 22), turbine production as well as the compressor consumption increase, with a maximum net power at around 3.0 pressure ratio. Conversely, cooling capacity firstly decreases and gradually rises with a higher pressure ratio. The lowest PER value, about 4.0, can be expected when the pressure ratio is almost 3.0.
637 638
Figure 22. Power and cooling capacity vs. pressure ratio (below) [75].
639 640 641 642 643 644 645
Medrano et al. [87] developed a thermodynamic model of an integrated energy system made-up by a MGT and a single-double effect exhaust fired absorption cycle, which recuperates the heat exchanged from the MTG exhaust gases using two generators at two different levels of temperature. An advantage of this system is that the exhaust gases leaving the high temperature generator of the double-effect cycle can then be further utilized to drive the single-effect cycle, producing an extra cooling effect. The overall exergy efficiency of the system resulted equal to 0.285, higher than the ones resulting from the same MTG and exhaust gas fired single or double effect chillers.
646 647 648 649 650 651
In reference [13], the energy and exergy performance of a small scale CCHP with GT was assessed; the system serves a cluster of buildings, and was analysed through a dynamic simulation, at design and part-load conditions. Here a double-effect ACH is coupled to a MGT with a power capacity of 1748 W. The Fuel Energy Saving Ratio (FESR) is calculated, defined as the ratio of the energy saving of the CCHP system, in comparison with a conventional system with separate productions (SP), to the overall energy consumption of the reference system.
ACCEPTED MANUSCRIPT 25 of 67
652 653 654 655 656
At design conditions, the FESR has a value as high as 26.6%. The lower the power load, the smaller the FESR value. When the load decreases to approximately 30% of the peak value, the FESR is zero, which means that the energy saving performance of the CCHP system is identical to the SP system. So, the CCHP system allows one to achieve energy savings only when the power output of the gas turbine exceeds 30% of the full load. Similar results are shown for the CO2 emissions.
657 658 659 660
A thermodynamic simulation of a trigeneration system consisting of a 28 kWe MGT coupled to a doubleeffect ACH and a heat exchanger to produce hot water is performed in ref. [103]. In particular, a certain amount of the energy of the exhaust gases at high temperature of the MGT are used to supply heat to the generator of the ACH, and the remaining part is used to produce hot water.
661 662 663 664 665 666
The trigeneration system was evaluated in different operating conditions, varying: i) MGT fuel mass flow rate; ii) ambient temperature: the higher the ambient temperature, the lower the MGT power; iii) generation temperatures, showing the sensitivity of the ACH performance to the variation of the temperatures of the MT exhaust gases (Figure 23): for example, if the ambient temperature is equal to 36 °C, the ACH cannot operate at generation temperatures lower than 146 °C; furthermore, at an ambient temperature of 28 °C a generation temperature close to 166 °C is needed to obtain the highest COP of 1.289.
667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686
Figure 23. COP against generator temperature at five different ambient temperatures [103]. The possibility of integrating an externally-fired micro-gas turbine (EFMGT) with a membrane distillation plant in a stand-alone application for the production of electricity, heat and drinking water is investigated by Rahman and Malmsquit [104]. Here, a simulation model of the components is developed (MGT, electrical and thermal loads, membrane distillation unit, and electrical and thermal storage) in Matlab/Simulink environment. A maximum overall efficiency of 72% was found, at the rated electrical output of 2 kW. The authors state that to maximize the overall efficiency, it is required to maximize the use of thermal waste. However, maximizing the overall efficiency subsequently leads to a lower temperature level at the demand side, and thus, the efficiency of the pure water production decreases. Therefore, such a polygeneration system can perform more efficiently in colder regions. Another important conclusion of this work is that the fuel consumption does not decrease significantly with part-load conditions. As a result, such EFMGT system should always operate at the rated electrical load, to obtain a high efficiency. 4.2 Steam turbine (ST) Several research works available in literature are focused on the way to improve the total efficiency of power plant and CHP based on steam turbines [35, 36]; such technology is often included in polygeneration systems [105]. Three different CHP systems designed to be installed in a sugarcane factory were simulated with the EES software in ref. [4]; they are based on a back-pressure steam cycle (configuration #1, Figure 24 b), a condensing steam cycle (configuration #2, Figure 24 c), and a combined cycle (configuration #3, Figure 24 d), respectively.
ACCEPTED MANUSCRIPT 26 of 67
687 688 689 690 691
Figure 24. Scheme of sugarcane production and different configurations of cogeneration systems integrated with the sugarcane factory: (a) diagram of integrated production of electricity, sugar and molasses; (b) configuration #1: backpressure steam cycle; (c) configuration #2: condensing steam cycle and (d) configuration #3: combined cycle (gas turbine+steam turbine) [4].
692 693 694
In this system, the bagasse produced in the extraction system of sugar cane is delivered to the cogeneration system, where it is used as fuel, to produce the electricity and steam consumed by the process. The most relevant results are given in Table 12.
695
Table 12. Performance of the systems simulated in ref. [4].
Configuration
Energy efficiency overall (%)
Bagasse (+) excess (−) shortage (t/year)
Electricity excess (MWh/year)
#1 #2 #3
84.2 65.2 56.9
(+) 316,400 0 (−) 301,600
119,300 306,100 503,100
696 697 698 699 700 701 702
It resulted that the global energy efficiency of the polygeneration system decreases when passing from configuration #1 to #3, but the economic value and the “quality” (exergy content) of the total product increase, due to the share of electricity production which also increases. Configuration #2 produces a considerably higher amount of electricity (186,800 MWh/year) with respect to #1. This increase of electricity production is obtained by consuming 316,400 t/year of bagasse, which represents a virtual electrical efficiency of 28%. In the case of configuration #3, the surplus electricity production is 503,100 MWh/year higher than #1 one; but for this configuration it requires buying more bagasse.
703 704
In ref. [112], a steam turbine is integrated in a lignocellulosic wood-to-ethanol process. The simulation study was carried out using the software IPSEpro™. The system evaluated is based on an existing CHP plant
ACCEPTED MANUSCRIPT 27 of 67
705 706 707 708 709 710 711 712 713
and on the pilot scale plant for the ethanol production. In particular, measured data from operation of the Enköping (Sweden) CHP plant were used for modelling and validating the CHP plant model, whereas the configuration of the pilot scale plant, owned by two universities of Örnsköldsvik (Sweden), was used for modelling the ethanol production. The total efficiency was found to be around 50%, meeting the heating load of the district heating system. The authors concluded that the ethanol production from wood is more efficient when integrated into a CHP plant, in comparison to the case of stand-alone production: in particular, the total biomass consumption is reduced by 13.9% while producing the same amounts of heat, electricity and ethanol as in the stand-alone configurations. In addition, a higher time of operation was obtained for the existing cogeneration system. Thus, the annual production of renewable electricity increased by 2.7%.
714 715 716 717 718 719 720 721 722
Cormos [109] developed a combined polygeneration process for power, hydrogen and carbon dioxide production, consisting of a steam turbine, a bio-methanol steam reformer and carbon capture technologies. In particular, the study is focused on the hydrogen production from bioethanol at industrial scale, investigating three carbon capture designs: one based on pre-combustion capture using chemical gas–liquid absorption (Case1) and two based on chemical looping, using syngas (Case 2) or direct bioethanol looping (Case 3, shown in Figure 25). In all the cases, the heating duties needed for various processes (e.g. reformer, carbon capture by gas–liquid absorption or chemical looping) are recovered from available hot streams within the plant, therefore, the only energy input of the plant is the bioethanol feedstock. The author modelled and simulated the different carbon capture designs with ChemCAD.
723 724 725
Figure 25. Layout of bioethanol conversion for hydrogen production with direct bioethanol chemical looping (Case 3) [109].
726 727 728 729 730 731 732
From the results, it can be noticed that the chemical looping cases (Cases 2 and 3) reach higher energy efficiency (59–63% vs. 53–58%) and almost total decarbonisation of the fuel (>99% vs. 80–92%). Case 3 resulted the most promising, mainly for the feature to capture almost total carbon of the feedstock and to have a lower plant complexity. In particular, the operation of this configuration expects that the bioethanol is vaporised and then it is introduced in the fuel reactor together with the oxygen carrier. The bioethanol is totally oxidised to CO2 and water. The energy efficiency in Case 3 resulted equal to 63.66%, vs. 59–62% of bioethanol reforming without carbon capture.
733 734 735 736 737 738 739 740 741
A new power and refrigeration cycle was simulated in ref. [29], where a Rankine cycle and the ejector refrigeration cycle are combined to produce simultaneously power and refrigeration simultaneously with only one heat source. This system can be driven by the flue gas of gas turbine or engine, solar energy, geothermal energy, industrial waste heat. In the paper, a heat source at the constant temperature of 150 °C was considered. An exergy and parametric analysis and a parametric optimization were performed. In particular, the parametric optimization was performed by means of genetic algorithm, aiming at reaching the maximum exergy efficiency. The results show that the biggest irreversibility is located in the heat addition process and in the ejector. An overall efficiency of 13.72% was found, defined as the ratio between the useful energy output (net power plus refrigeration effect) and the total heat input; the exergy efficiency in optimum conditions was 27.10%.
ACCEPTED MANUSCRIPT 28 of 67
742
4.3 Combined Cycles plants (CC)
743 744 745 746
CC plants represent the most efficient technology for energy conversion, and a common option for new thermoelectric plants all over the world [100]. In combined cycles (CC), the waste heat of a gas turbine (topping cycle) is used to drive a steam turbine (bottoming cycle), so increasing the overall thermodynamic efficiency up to 55% and over [24].
747 748 749 750 751 752 753
An example of integration of CC as prime mover into a trigeneration plant is reported in ref. [21]. Here, the thermodynamic modelling of a trigeneration system based on a gas turbine, a steam turbine and a singleeffect ACH is carried out; an exergy and an environmental analysis are performed. The layout of the system is depicted in Figure 26. The hot gases of GT are utilized in the dual-pressure HRSG with two economizers and two evaporators to generate low-pressure (LP) and high-pressure (HP) steam. The LP steam is used to drive the ACH and the HP steam to generate electricity. Some of the results obtained are reported in Table 13, putting in evidence lower CO2 emissions and higher exergy efficiency with respect to typical CHP systems or GT cycles.
754 755
Figure 26. The schematic of the trigeneration system [21].
756
Table 13. Main results of the energy and exergy analyses in ref. [15].
757 758 759 760 761 762 763
Parameter
Value
Qheating (MW) Qcooling (MW) Wnet (MW) ηI,Tri (%) ExD,Tot (MW) ηex,Tri (%) Normalized CO2 emission (kg/MWh)
46.60 3.50 35.15 75.50 65.80 47.50 158
A similar layout was investigated by Wang et al. [60], who assessed a new CCHP system integrating a gas turbine and a heat-driven cooling/power cogeneration system, using Aspen Plus . In particular, the layout used in this work includes: an ammonia-water Rankine cycle; a single-double effect LiBr/H2O ACH, which better matches the temperature variation of ammonia-water condensation; a Bryton cycle; a hot water exchanger. The novel system was compared to a reference CCHP system including a gas turbine, a double-effect ACH and a heat exchanger. Two evaluation criteria were adopted to evaluate the CCHP performance: exergy efficiency (ηex) and fuel energy saving ratio (FESR). Thanks to the high-temperature heat cascade utilization in the CCHP,
ACCEPTED MANUSCRIPT 29 of 67
764 765 766 767
it resulted that with the same amount of exergy input (7477.78 kW), the exergy output of the system under analysis is 2833.58 kW, and that of the reference system is 2265.09 kW; ηex reaches 37.89%, which is 7.60% higher than that of the reference system, and the FESR is 31.70%, which is 5.19% higher than that of the reference system.
768 769 770 771 772 773 774
Several works are also focused on polygeneration systems coupling CC systems with a gasification technology (coal gasification or biomass gasification), obtaining an integrated gasification combined cycle (IGCC). In an IGCC scheme, the solid feedstock is partially oxidized by oxygen and steam to produce syngas (mainly a mixture of carbon monoxide and hydrogen). An example of this configuration is reported in ref. [119], where the modelling and simulation of a novel polygeneration technology is performed, used for the coproduction of hydrogen and power based on coal gasification integrated to a CC with a dual chemical processes (with carbon capture and storage and with calcium-based chemical looping).
775 776 777 778 779 780
A further simulation study regarding coal gasification, coupled with methanol synthesis and CC for power production with CO2 recovery, is presented in Lin et al. [17]. Here, the proposed system was analysed with and without CO2 recovery. In particular, the polygeneration system proposed is based on the concept of CO2 generation, reaction, transportation, and conversion into energy, with the production of methanol as a clean fuel (Figure 27). In this way, the polygeneration system can recover CO2 with reduced (or even without) energy penalty.
781 782 783
Figure 27. Sketch map of the integration of cascading utilization of chemical energy and CO2 capture [17].
784 785 786 787 788
It was found that the polygeneration system without CO2 recovery consumes less energy for the methanol production (about 76% of the traditional methanol production), and achieves higher power generation efficiency, 15% points higher than that of the traditional power generation technology, i.e. IGCC power plants). In case of CO2 recovery, the energy consumption for methanol production is much less than those of the Fischer–Tropsch fuel and conventional methanol production.
789 790 791 792 793 794 795 796 797
As mentioned above, the biomass gasification represents a further technology which can be used to generate both process heat and significant amounts of electricity. An example where this technology is used in a combined cycle plant is in ref. [61]. Here, the electricity produced is used to match the needs of dry-grind ethanol facilities by utilizing the ethanol process co-products and other biomass sources (corn cobs). The model of the dry-grind ethanol process is simulated by Aspen Plus and, then, used as basis for modelling a gasification system. The authors report that this technology is able to reduce fuel costs for ethanol plants and that the renewable energy ratio of ethanol production could be improved from a typical value of 1.7 up to 5.1; according to this study, a dry-grind ethanol facility with a capacity of 190 million litres per year could produce 30.4 MWe of power while covering all its needs of process heat by using ethanol co-products and corn cobs.
798 799 800 801 802 803
4.4 Organic Rankine Cycles (ORC) The technology of Organic Rankine Cycles (ORC) is very attractive in applications where heat at relatively low-medium temperature is available for conversion into electric energy [62]. In fact, many organic fluids show a performance significantly better than water in Rankine cycles with low maximum temperature, thanks to a higher molecular weight, lower evaporation heat, positive slope of the saturated vapour curve in the T–s diagram and lower critical and boiling temperatures [44]. In addition, for small-size applications, ORC turbines show a
ACCEPTED MANUSCRIPT 30 of 67
804 805 806 807 808
few advantages in terms of maintenance, operation life and part-load efficiency. Anyway, for such systems the selection of an appropriate working fluid is crucial [144]. Several works in literature deal with the simulation of polygeneration plants including ORC systems supplied by low-medium temperature sources (geothermal heat [18], solar energy [43, 44], waste heat [45], biomass products [88], etc.).
809 810 811 812 813
In the work reported in reference [22], an ORC-based trigeneration system is analysed, supplied by the waste heat recovered from a micro GT. The system was modelled in Matlab, whereas the properties of the working fluid (n-octane) were calculated with EES; an exergo-environmental analysis of the system was performed and discussed. The system layout mainly consists of a GT cycle, an ORC, a single-effect ACH and a DHW heater (Figure 28), producing heating, cooling, hot water and electricity.
814 815 816 817 818 819 820
Figure 28. Schematic of the trigeneration system [22]. The exergy-environmental analysis shows that the highest exergy destruction occurs in the combustion chamber and heat exchanger (Figure 29), mainly due to the irreversibilities associated with the large temperature difference occurring in these components. The absorption cycle does not exhibit significant exergy destructions. It resulted that the system with ORC has less CO2 emissions and higher exergy efficiency than the GT alone, providing a significant motivation for the use of trigeneration cycles.
821 822
Figure 29. Exergy-environmental results [22].
823 824 825
The simulation of an ORC system, driven by the waste heat recovered from a diesel engine generator, is studied in ref. [145]. In particular, the aim and the novelty of this work is to investigate the system performance and efficiency of a small scale trigeneration system integrating an ORC system into a diesel engine generator,
ACCEPTED MANUSCRIPT 31 of 67
826 827 828 829 830 831 832 833 834
used as primary mover. The ORC model is developed in Matlab and the diesel engine generator of the laboratory at Newcastle University is used as the case study. R245fa is used as the working fluid. In addition, in order to expand the system supply capacity, the trigeneration system is also coupled to a smart energy storage system. Based on the simulation results, with the energy storage system, the capacity of the electric cycle increased by 47%. The ORC system may generate 0.737 kW of power, by contributing around 8.91% of the total electricity production, with 9.25% efficiency at the generator full-load operation point. The waste heat recovered from the engine is 35.3% of the total energy. Regarding the heating and cooling energy demand, for the selected household, it is demonstrated that the ratio of the recovered heat by ORC satisfies the 41% of the heating energy demand and that part of waste heat used by ACH produce 9.36% of the cooling energy demand.
835 836 837 838 839 840 841 842 843
Solar thermal collectors can be adapted to collect solar energy at low temperature and provide the heat necessary to supply ORC plants. An example of this configuration is reported in the study of Sharaf et al. [30], where the thermal energy produced by a field of PTC is used to drive the ORC via evaporator heat exchanger. In this plant, the exhausted energy from the ORC turbine is also used in the first effect of a MED process, producing power and desalted water (power and solar desalination with MED, PSDMED). This system was compared with a further system, where the solar energy is directly utilized from the via evaporator heat exchanger to the first effect of MED process, producing only potable water (solar desalination with MED, SDMED). The most important results obtained in the both systems are summarised in Table 14, considering a MED plant with a capacity of 5000 m3/day.
844
Table 14. Results for the systems analysed in [30], for a capacity of distillation equal to 5000 m3/day. Parameter
Unit
SDMED
PSDMED
Solar field area Specific power consumption
m2 kWh/m3
1.009×105 2.17
1.32×105 2.67
-
7.56
3
$/m3 $/h MW % MW kg/s
1.645
1.845
155.7 31.82 – 7.65
157.8 33.1 5.381 46.05
Gain ratio* *Distillate
mass flow rate / Steam mass flow rate
Total water price Total investment, operating and maintenance cost Total exergy destruction Overall exergy efficiency Turbine power Steam mass flow rate
845 846 847
Comparing the total cost of the desalinated water, the solar field area and the total exergy destruction rate, it is evident that the PSDMED performs slightly better than the SDMED, with the advantage represented by the availability of a significant power provided by the ORC system.
848 849 850 851 852 853 854 855 856 857 858 859
An ORC-based polygeneration system supplied by biomass, combined with water desalination, was studied by Maraver et al. [62]. The system consists of a small-scale ORC engine, producing power and heat; this latter is recovered for direct heating purposes or DHW production, to generate cooling by a LiBr/H2O ACH and to produce desalted water through a MED unit. The ORC subsystem was modelled using Aspen Plus. The configuration proposed is compared to stand-alone systems producing the same energy and material outputs. The authors investigated the behaviour of the plant by varying the main design and operating parameters of each components, including the working fluid for the ORC unit; among 33 working fluids analysed, it was found that Fluorobenzene and octamethyltrisiloxane offer good cycle efficiencies as well as acceptable densities and specific power capacities. The authors also determined the optimal distribution of the heat generated by the ORC, finding out that the highest savings were obtained when the amount of heat used for heating or DHW production was around 60%. Moreover, an economic analysis was performed, and it was found that the SPB ranged between 4.00 and 20.0 years, depending on the price of biomass and on the capital cost of the MED unit.
860
4.5 Internal Combustion Engine (ICE)
ACCEPTED MANUSCRIPT 32 of 67
861 862 863 864 865 866 867 868 869 870
Internal combustion engines (ICEs) represent a very mature, largely available and well-known technology [146], and, depending on their design, can use different fuels: natural gas, biogas, biodiesel, etc. [11]. Such engines are the most widely used prime movers in small- and medium-scale distributed energy resource (DER) systems [147], due to their low capital cost and robustness [148]. Some limitations are the need for frequent maintenance, vibration, noise and emission issues [147]. It must be noted that in an ICE, only 30–40% of an engine's fuel combustion energy is converted into electricity [149]. So, a significant amount of waste heat must be dissipated, mainly through expulsion of exhaust gases and engine cooling system. In order to recover such waste heat, conventional ICEs are often coupled to heat exchangers. In addition, ICEs can be easily coupled with other devices, such as adsorption [126] and absorption chillers [127], heat pumps and electric chillers [63, 150, 151].
871 872 873 874 875 876 877 878 879 880 881 882 883 884
An example of this application is investigated in the simulation and experimental work carried out by Wu et al. [126]. In this work, a micro-CCHP based on a 16 kWe ICE is investigated, whose exhaust heat is used to drive a 7.5 kW adsorption chiller. A thermal management controller (TMC) is also included to manage the waste heat: the heat is sent with the highest priority to a buffer tank, then to a hot water cycle, and finally to the cooling water cycle (Figure 30); this depends on the set point temperatures of buffer and hot water tanks. The heating load is controlled by fan coil 2 and fan coil 3, the cooling load by fan coil 1. The model of ICE is based on the statistical data provided by ASHRAE for naturally aspirated, small ICE. Experimental results were found to be in good agreement with the simulations. It resulted that the system can provide simultaneously a power output of 16 kW electric, up to 17.7 kW for heating purposes and 6.5 kW for cooling (Figure 31). In particular, as shown in this figure, when the electric output is low, the heating output is nearly 0, namely all the waste heat is absorbed by the adsorber. When the electric output is over 6.4 kW, the heating output starts to increase while the cooling output and COP of adsorber both change slightly. Namely, when the electric output is over 6.4 kW, the waste heat recovered by TMC exceeds the heat transfer capacity of the first stage and part of waste heat is assigned to the second stage.
885 886
Figure 30. Layout of experimental micro-CCHP system [126].
ACCEPTED MANUSCRIPT 33 of 67
887 888
Figure 31. Influence of electric output on heating and cooling output [126].
889 890 891 892 893 894 895 896 897
The simulation of a medium size ICE (power capacity of 300 kW) with Aspen Plus was presented by Chen et al. [63]. Here, the waste heat recovered from the jacket water circuit (at 90 °C) and from the exhaust gases (at 500 °C) is used to generate additional power, by means of a turbine, and cooling energy, by an ammoniawater absorption chiller. In design conditions, the heat recovered from the exhaust gas and jacket water are 289.56 kW and 180.47 kW, respectively. The net power output of the turbine is 14.76 kW and the refrigeration output is 78.10 kW, calculated as the ratio of the cooling output (225.72 kW) to a reference COP of 2.89. So, the corresponding total equivalent power output is 92.86 kW. An equivalent heat-to-power efficiency of the combined production of power (through the GT) and cooling is defined and resulted equal to 19.76%. An exergy efficiency of 33.69% and a payback period of 2.84 years were estimated.
898 899 900 901 902 903 904 905 906 907
An internal combustion engine included in a CHCP system for a small supermarket was simulated by Maidment et al. [127]. The engine produces: i) electricity to drive a low-temperature vapour-compression refrigeration cycle and for lighting, HVAC and other equipment; ii) heat, recovered from engine jackets and exhaust gases, used to satisfy the space-heating and hot-water demands; iii) cool, by an ACH used to refrigerate propylene glycol to −10°C for cooling the chilled-food cabinets in a supermarket. A gas-fired boiler is also included in the system to supplement the heat provided by the combustion engine when necessary and a wasteheat exchanger is used to reject the excess heat to the atmosphere. The investigation has shown that this configuration is extremely efficient, with Primary Energy Savings (PES) up to 20%. A payback period of 6 years has been calculated. For larger stores, even better payback periods could be achieved, as the unit cost of the ACH significantly decreases.
908 909 910 911 912
Sibilio et al. [11] simulated an innovative building-integrated micro-trigeneration system, based on a 6.0 kWel / 11.7 kWth natural gas-fuelled ICE feeding an electric air-cooled vapour compression water chiller for heating, cooling and DHW purposes, in a multi-family house. (Figure 32). The same plant configuration is simulated also in ref. [14], evaluating the energy, environmental and economic effects of electric vehicle charging.
ACCEPTED MANUSCRIPT 34 of 67
913 914
Figure 32. Scheme of the proposed system in [11].
915 916 917 918 919 920 921 922 923 924
The system also includes a hot water tank with three internal heat exchangers, an auxiliary boiler, an electric air-cooled water chiller, a heat exchanger, thermostats, pumps, diverters, and a group of fan-coils installed in the building. The plant was dynamically simulated in TRNSYS, comparing it with the conventional separate energy production. The model of the ICE unit was calibrated and validated based on measured data. As for the operation strategy, a thermal load tracking strategy was adopted. The system was analysed for three different Italian cities (Palermo, Naples and Milan), evaluating the CO2 emissions, Primary Energy Saving (PES) and operating costs (OC). The Figure 33 highlights that PES ranges from 0.2% to 4.3%, ΔCO2 from 3.9% to 10.6%, ΔOC from 11.3% to 19.9%, with the best results in Milan. Although the PES and ΔCO2 are negative during the summer, the authors highlight that in such season the system can reduce the load peaks on the electric grid, alleviating the risk of network congestion and failure events.
925 926
Figure 33. Values of (a) PES, (b) ΔCO2 and (c) ΔOC for the three investigated city [11].
927 928 929
Rey et al. [152] designed, built, tested and dynamically simulated in TRNSYS a gasoline ICE micro CHP system, whose exhaust gases and engine cooling fluid are recovered for a recreational sailing boats used as mobile homes. The system, producing heating, DHW and electricity, was designed to work independently, and
ACCEPTED MANUSCRIPT 35 of 67
930 931 932 933 934
therefore no auxiliary system was considered, whereas an electricity and thermal energy storage to match the demand and production were included (Figure 34). The performance of the ICE was compared with similar system based on a Stirling engine (SE). The transient simulations were performed for three different European climates (Helsinki - Finland, Breskens -. The Netherlands and Malaga - Spain) and varying the size of the electricity storage.
935 936
Figure 34. ICE-based micro CHP unit layout [152].
937 938 939 940 941 942 943 944 945 946
The authors present their experimental results showing electrical and thermal capacities of ICE system equal to 0.653 kW and 5.414 kW, respectively. From the simulations, it resulted that the system guaranteed the heat demands in all climates. The results of SE and ICE were similar as for electric and thermal energy production, for all the locations, because the simulations are performed assuming that a thermal-load-match control strategy. The slight differences in heat production are due to the small differences in the heating capacity of the two systems (5.93 kW for SE). Annual electrical coverage never reaches 100% (Figure 35). Nevertheless, the ICE supplies less electricity because of its lower electric/thermal capacity ratio: 13%, compared to 15% with SE. The parametric analysis showed that the optimal capacity of the electricity storage system was double with respect to the initial configuration (24 V/100 Ah), leading to a significant increase (up to 11% for Malaga) of the electrical production.
947 948 949 950 951 952
Figure 35. Annual thermal production in kWh (left) and electrical coverage (right) [152]. 4.6 Fuel cells (FC) Fuel cells are based on the direct conversion of the chemical energy of fuels into electricity, without any combustion nor mechanical drive. Heat, water and other outputs are also available, so that these systems are very attractive for polygeneration applications (Figure 36).
ACCEPTED MANUSCRIPT 36 of 67
953 954
Figure 36. Polygeneration system based on FC technology.
955 956 957 958 959 960 961 962 963 964 965 966 967 968
FC are known as the most efficient small-scale power plants, featuring high electric efficiency (up to 60%) and overall efficiency (up to 90%) even at mini or micro-scale sizes [11]. They have several benefits, such as: almost zero emissions, a good match with the residential thermal-to-power ratio, reliability, quiet operation, potential for low maintenance, excellent part load behaviour [153, 154]. The work carried out by Soutullo et al. [12], reported above, is not the only one considering the integration of FCs into polygeneration systems. In fact, a number of possible configurations were investigated in literature, such as: Solid Oxide Fuel Cells (SOFC) [19] or PEMFCs [74] integrated with solar heating and cooling systems, as well as PEM integrated with CPVT and an electrolyzer. This last configuration was dynamically simulated in TRNSYS environment and analysed from the energy and economic points of view in the work of Calise et al. [69]. Here, the system proposed consists of a 600 m2 CPVT field (PTC with triple-junction solar PV cells), PEMFCs (180 kWe and 153 kWth), a 375 kW single-stage LiBr/H2O ACH, a 216 kWe alkaline water electrolyzer and an auxiliary gas boiler (Figure 37) and produces space heating / cooling, DHW, electricity for a small real university building. In particular, the electrolyzer system is powered only by the CPVT electricity and produces hydrogen, used to supply the FC and oxygen, which is sold.
969 970
Figure 37. System layout analysed in ref. [69].
ACCEPTED MANUSCRIPT 37 of 67
971 972 973 974 975 976 977 978 979 980
Some of the results of a dynamic simulation performed in TRNSYS, for the climate conditions of Naples (South Italy), are reported in Table 15. The CPVT thermal efficiency, calculated with respect to the beam radiation, is 63.4 %. The electrical one is 19%, due to the use of a multijunction PV cells. The overall efficiency at full load of the FC is 78.4%. 26.8% of the electrical energy supplied by CPVT to produce hydrogen is dissipated due to the EL inefficiencies. The yearly mean ACH COP is about 0.70. In case of a capital investment subsidy of 50%, SPB significantly decreases to about 6 years, in such a way the system can became profitable. A parametric analysis on the PEMFC electrical power variation is also carried out, and from the economic point of view it was observed that when an optimal FC nominal power of 100 kW is selected, the SPB is about 5 years, in presence of incentives. Therefore, this work confirmed the technical and economic feasibility of such innovative polygeneration system.
981
Table 15. Efficiency and economic performance for the system analysed in ref. [69] - yearly results.
982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997
Parameter
Value
ηth,CPVT
0.634
ηel,CPVT
0.189
ηth,FC
0.434
ηel,FC
0.35
ηEL
0.732
COPACH
0.729
SPB
12.5
SPB with 50% subsidy
5.84
Unit
–
years
An interesting polygeneration microgrid was simulated in TRNSYS 16 in ref. [55], including a lead acid battery bank, wind turbine, monocrystalline PV panels, a Proton Exchange Membrane (PEM) fuel cell, a PEM electrolyzer, a metal hydride tank, a reverse osmosis (RO) desalination unit using energy recovery, a H2 vehicle and control system. In particular, the hydrogen vehicle is a hybrid Fuel cell – battery scooter with an average hydrogen consumption of 2.4 Nm3 /100 km. The polygeneration microgrid covers the electricity and potable water needs for two households; in addition, the hydrogen produced is used to covers the energy consumption related to transport of the same final users. The system is considered to be installed on a small island in the Aegean Sea, Greece. In this work, the system was evaluated from an economic point of view, evaluating NPV and SPB. A detailed optimization procedure aiming at minimising investment and operational cost was carried out, too. The results showed that, due to current high H2 operating cost, it was preferable to store the produced electricity in a bigger battery bank (1000 Ah at 48 V), by reducing the size of the H2 electricity storage. The FC optimal size resulted equal to 300 W; therefore, the FC is used as a backup energy producer, since its produced energy is significantly lower with respect the ones produced by PV panels and wind turbine (Figure 38). In any case, the authors state that this result does not suggest removing FC in their configuration, since the use of FC electricity avoids to completely discharge the battery and the problems connected with inverter synchronization. The NPV for the polygeneration microgrid is 21,236 €.
Energy (GWh)
PV Produced Energy Wind Turbine Produced Energy
Electrolyzer Unit Consumed Energy Desalination Unit Consumed Energy
Electrical load
998 999
Figure 38. Yearly produced and consumed energy [55].
ACCEPTED MANUSCRIPT 38 of 67
1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012
The previous papers investigated FC polygeneration systems as hybrid systems, that is, the FC integration into another electricity generation system. An application regarding polygeneration systems based on SOFC as the unique prime mover is carried out in reference [155]. Here, a planar 5-kW low-temperature SOFC polygeneration system, based on nanocomposite materials, is analysed. The SOFC operating temperature reaches 550°C and its power density is 0.4 W/cm2 or 0.7 W/cm2, when supplied by hydrogen or syngas, respectively. In this work, the system design and the energy and mass balances, based on lumped-parameter model, are presented; a simulation based of the use of syngas as the input fuel is performed, too. The results show that the electrical and thermal efficiency of the SOFC are equal to 19.5% and 57.6%, respectively. In addition, the authors state that there is a considerable amount of heat at the FC outlet, which can be harvested for further use, despite the fact that this SOFC has lower operating temperature than common SOFC systems. Therefore, there is a high potential in low temperature SOFCs for polygeneration application, because they do not have the common problems of high temperature SOFCs and are less expensive. Besides, the use of waste heat coming from the SOFC stack enhanced the efficiency to about 75%.
1013 1014 1015 1016 1017
Becker et al. [95] presented the simulation, design and techno-economic performance analysis of a 1 MW SOFC polygeneration system for combined production of heat, hydrogen, and power. The system was simulated by Aspen Plus in order to evaluate the thermodynamic performance of the system and its components. The electrical efficiency of the SOFC at the rated power was estimated equal to 48.8%; the overall plant efficiency was found equal to 85.2%.
1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030
In references [156, 157], the trigeneration system proposed is based on SOFC, which was completely analysed from energy, exergy and exergoeconomic points of view. Energy for heating and cooling is obtained by recovering heat from the gases exiting the SOFC. The authors investigated the influence of two significant SOFC parameters, such as current density and the inlet flow temperature, on several output variables. Findings show that the minimum energy efficiency of the trigeneration system is 33%, higher than for the SOFC alone. Furthermore, the maximum energy efficiencies are 79% in case of trigeneration, 69% in case of combined power and heat production, 58% in case of power and cooling cogeneration and 46% in case of electricity production. The maximum exergy efficiency is about 47%, obtained in case of trigeneration. The exergy analysis showed that the main exergy destructions occur in the air heat exchanger, the SOFC stack and the after burner, which was supposed to oxidize all the non-reacted gases exiting the SOFC. As regards the exergoeconomic analysis, the authors investigated the effects of three important SOFC parameters (current density, inlet flow temperature and fuel utilization factor), on several output variables, such as: unit cost of the electrical power, unit cost of the energy for cooling and heating, and total unit cost of the products.
1031 1032 1033 1034 1035 1036
5. By-products of polygeneration systems In the majority of the works described above, polygeneration systems mainly focused on the production of energy vectors were analysed. However, several other studies presented novel systems, purposely designed to produce other outputs. In general, polygeneration systems are potentially able to provide power, space heating and cooling, DHW, desalinated water, but also chemical products, synthetic fuels and other materials (Figure 39).
ACCEPTED MANUSCRIPT 39 of 67
Space heating, DHW Synthetic fuels
Space cooling
Outputs Chemical products
Power
Others
Desalinated water
1037 1038
Figure 39. Outputs of polygeneration systems.
1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052
In particular, several papers are available in literature concerning the use of polygeneration systems that include desalination facilities. In fact, thermally-activated desalination processes can be economically profitable when fed by waste heat, for example in cogeneration plants [158]. However, even electric-driven desalination facilities, based on Reverse Osmosis (RO), are often considered. A polygeneration scheme based on a 24.7 MW natural-gas GT was adopted in [99]; here, the GT cycle is coupled to a bottoming ORC turbine, that is mechanically coupled to the RO unit.The system was developed for on-site hydrocarbon production in the Arabian Gulf. The thermodynamic performances of the combined power cycle were calculated by Aspen Plus V8.4, those of the RO unit by the software Dow Chemical Company RO System Analysis (ROSA). The energy, exergy and economic analyses were carried out for four different organic fluids, but are mainly focused on octamethyltrisiloxane, which yields 6 MW of additional net power output. The RO unit operates with a specific energy consumption of 4.1 kWh/m3 and an exergy efficiency of 29%. The exergy efficiency of the polygeneration system as a whole is estimated equal to 32%, thereby enhancing the efficiency of the GT alone by 6%. The system becomes profitable after approximately 3 years, provided that the desalinated water can be sold at a subsidized price.
1053 1054 1055 1056 1057 1058 1059 1060 1061 1062
The use of renewable energy sources (solar, biomass, geothermal) to power desalination processes in polygeneration systems is also widely investigated. An interesting dynamic simulation was presented in ref. [106], regarding a solar MED polygeneration system. In particular, in this study, a PTC field coupled to a 50 MWe Rankine cycle is considered. The heat needed to drive the MED unit is provided by the Rankine cycle condenser. The system was dynamically simulated in TRNSYS; only the MED unit was modelled with Matlab. Both thermodynamic and economic analysis are performed, and two case studies are presented, for Venezuela and northern Chile. It was found that the polygeneration system can provide electricity and water for more than 85,000 inhabitants at favourable prices, increasing the total plant annual cost of only 6%÷12%. The plant was also optimised, and it resulted that the desalination capacity in the optimum configuration was 29,323 m3/day for Venezuela and 22,160 m3/day for Chile.
1063 1064 1065 1066 1067 1068 1069 1070 1071
A polygeneration system integrating solar collectors and membrane distillation was investigated by Mohan et al. for United Arab Emirates locations [76]. Here, the solar energy is directly provided to the desalination unit, as shown in Figure 40. A 35.2 kW single stage LiBr/H2O ACH is also included in the system to meet the cooling load of a group of portacabins, modelled in TRNBUILD and located in Rakric. The production of DHW is also included. The whole system is modelled in TRNSYS. The model is validated by on-site experiments. In this study, three different solar technologies are compared: flat plate, evacuated tube and compound parabolic collectors. The dynamic simulations show that the best pay-back value, equal to 6.75 years, is achieved in case of evacuated tube collectors having gross area of 216 m2. The best performance in terms of daily production of membrane distillation and DHW production are 92.8 kg/day and 301.67 kg/day, respectively.
ACCEPTED MANUSCRIPT 40 of 67
1072 1073
Figure 40. Schematic layout of the solar polygeneration system analysed in ref. [76].
1074 1075 1076 1077
Further interesting configurations analysed in literature are based on the use of geothermal energy, eventually coupled with solar energy. Such kind of polygeneration systems can be ideal in small volcanic islands, where the availability of solar and geothermal energy is high whereas the availability of fresh water is scarce [3]. For example, this idea was investigated in the works reported in references [3, 16, 37, 39].
1078 1079 1080
In the framework of polygeneration systems, the co-production of energy and chemical products has recently attracted a great interest. In particular, several works are focused on the production of methanol [17, 59, 89], ethanol [15, 24, 61, 112] and hydrogen [69, 90, 109, 119].
1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092
The integration of a methanol-producing biorefinery with an existing CHP unit and a local industry (a butchery) in the Danish city of Horsens is studied in ref. [113]. In this work, the concept of flexible multigeneration system (FMG) is highlighted, defined as integrated systems that generate multiple energy services and are able to adjust operation in response to fluctuating demand patterns and varying price schemes in the overarching energy system. The FMG under evaluation was developed to produce methanol from renewable biomass, in particular wood chips, and to match the local district heating demand and the thermal demand of the butchery. The layout includes a two-stage biomass gasifier, a solid oxide electrolysis cell, a methanol production facility, industrial heat pumps, and novel heat and gas infrastructures. To simulate each system component, the energy system modelling tool SIFRE is used; a detailed procedure of optimization is also described, aiming at maximizing NPV and minimizing the total CO2 emission. For the optimal design, the NPV was estimated to vary within the range 252.5–1471.6 M€, mainly depending on the capital cost and on the price of methanol.
1093 1094 1095 1096 1097 1098 1099 1100 1101
A techno-economic analysis was presented by Adams and Barton [118] for a polygeneration plant fed by coal and natural gas and co-producing methanol through methanol synthesis, as well as diesel fuel and gasoline through the Fischer–Tropsch process, and electricity. In this work, a new strategy regarding the natural gas reforming for efficient polygeneration systems is also proposed. Such strategy, defined ‘’internal reforming’’, is based on the fact that the gasifier cooling is provided by the reforming of natural gas, rather than by steam generation. The simulations show that the ’’internal reforming’’ provides increased energy efficiency and can be the optimal design choice in many market scenarios. In particular, th einternal reforming is up to 2 percentage points more efficient than the external one, due to the improved heat integration when about half or more of the output was liquid fuels.
1102 1103 1104 1105 1106 1107 1108 1109 1110
Regarding the ethanol production, a research work aiming at analysing the effects of introducing a lignocellulosic ethanol polygeneration plant into the district heating system in Stockholm, Sweden, was carried out in ref. [114]. The plant has capacity in terms of an ethanol production corresponding to 95 MWth, with biogas, electricity and heat as co-products. The system was studied using MODEST, a model framework based on linear programming, developed for optimisation of dynamic energy systems with time-dependent components and boundary conditions. The results show that such system would produce 110 GWh of electricity annually and furtherly, income from the sale of the biofuels and electricity produced would be about €76 million and €130 million annually, respectively, which is an increase of 70% compared to the income from the electricity produced in the system today. A reduction in global CO2 emissions of about 0.7 million tonnes
ACCEPTED MANUSCRIPT 41 of 67
1111 1112
annually, would be also reached. If the electricity price would increase by 20%, the system global CO2 emissions would be even lower and the income from the sold co-products would be about €232 million annually.
1113 1114 1115 1116 1117
Hydrogen production is another interesting process included in polygeneration systems [159, 160]. In fact, it can be used to produce electricity with low carbon emissions in a GT [161], or in SOFCs [162, 163]. Alternatively, H2 can be properly used for PEMFCs. In this case, high purity H2 is needed. Pure hydrogen is generated from syngas by the pressure swing adsorption (PSA) technique, with 85% hydrogen separation efficiency [120, 121].
1118 1119 1120 1121
An interesting polygeneration system, aiming at producing hydrogen, oxygen, power, heat and cool, involving the application of both fossil and renewable energy sources is investigated in [84]. Here a solar power tower is combined with a coal gasification system, a GT fuelled by syngas, a ST supplied by wasted heat of GT cycle, a single effect ACH, a hot water production unit, and a hydrogen production unit (Figure 41).
1122 1123
Figure 41. Schematic diagram of the multi-generation energy production system [84].
1124 1125 1126 1127 1128 1129 1130 1131
In particular, the heat produced from the solar tower is converted by a Rankine cycle into power, in turn used to supply an electrolyser which produces hydrogen. The thermodynamic properties of the working fluids involved into system are calculated by using EES software. The energy efficiencies for the different subsystems, i.e. coal gasification system, Rankine cycles, GT system, ACH and heating unit, hydrogen production unit, and of the overall multi-generation system, were evaluated, from both the energy and exergy point of view. In particular, regarding the hydrogen production, the reached energy and exergy efficiencies were equal to 19.43% and 14.41%, respectively; the obtained energy and exergy efficiencies of multi-generation system were equal to 54.04% and 57.72%, respectively.
1132 1133 1134
In Table 16, some of the works cited in this section are summarised, classified on the basis of possible inputs (renewable, fossil and hybrid), technologies used and outputs, are reported, along with other information regarding their economic performance (Simple Pay Back or Net Present Value).
1135
Table 16. Economic performance of some of the systems described in this section. Work
Input
Technology
[99]
Natural gas, seawater
RO, ORC, GT
[67]
Solar, Natural gas
[18]
Geothermal and solar
CPVT, GT Evacuated collectors, ORC
Output Fresh water, power, heat power, heat, cool
Economic index SPB=3 years SPB=4 years SPB=3 years
ACCEPTED MANUSCRIPT 42 of 67
1136
[10]
Biomass and Solar
[57] [15, 94]
Solar Agricultural waste
PTC, ORC, willow pellets boiler PVT gasifier, CC
[98]
Coal
GT, ST
[11]
Natural gas
ICE
SPB = 7 years
power, ethanol power, ethane and propane power, heat, cool
SPB = 4 years SPB = 5 years NPV= 353 $ million SPB= 8 years
6. Polygeneration systems for building
1137 1138 1139 1140 1141
The application of trigeneration / polygeneration technologies within buildings, also known as building cooling heating and power (BCHP), is widely studied in literature [147], due to its potential contribution to reduce the energy consumption in buildings, that is responsible for about 40% of global energy consumption and more that 30% of greenhouse gases emissions [164]. Small-scale BCHP systems can be important to promote the transition toward Net Zero Energy Buildings (NZEBs).
1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156
In this framework, the integration of solar technologies within the building in polygeneration systems is investigated in several literature works [18, 39, 67, 68], evaluating the system performance with dynamic simulation tools, by taking into account the hourly power, heating and cooling loads. An interesting example is represented by the work by Marín-Sáez et al. [165], where a building integrated PVT (BIPVT) system is analysed, designed to be superimposed on the blinds of a solar louvre shading system. The system is equipped with cylindrical holographic lens. The simulations are performed calculating the direct normal irradiance spectrum by the SMARTS radiative model, which relies on the following parameters: air mass, aerosol optical depth, precipitable water, and Angstrom exponent. This model is subsequently validated versus experimental data, showing a good agreement. The optical model is based on the Kogelnik’s coupled wave theory and the approximate scalar theory established by Syms [166]. Finally, energy simulations are carried out in TRNSYS; in particular, the system was evaluated under the weather conditions of Sde Boker (Israel) and Avignon (France), for specific case studies. In order to perform a comprehensive simulation, the optimal model, developed in MATLAB, was also integrated in TRNSYS environment. In addition, the developed PVT collectors were coupled with a conventional solar thermal system, obtaining a “combysolar” system able to supply the building with space heating, DHW and electricity (Figure 42).
1157 1158
Figure 42. Simulated system topology [165].
1159 1160 1161 1162
The authors found that the optical efficiency in Avignon is higher than that calculated for Sde Boker. In addition, for both cases, optical efficiency dramatically decreases in summer. The maximum monthly average optical efficiency is about 81% (Avignon, December). A similar trend is also detected for the efficiency related to space heating. However, in this case, Sde Boker exhibits higher values. Solar fractions, related to DHW
ACCEPTED MANUSCRIPT 43 of 67
1163 1164 1165
production, were 79.3% and 95.5%, for Sde Boker and Avignon, respectively. The solar fraction for space heating was above 15% in both locations. The solar fraction for electric energy was slightly below 10% in both cases.
1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177
The hybridization of trigeneration systems with solar systems was also investigated by Marrasso et al. [42]. The authors developed such system by considering that the increase of energy within building is often due to the growing demand for cooling and solar cooling systems can avoid that such needs are covered by conventional systems. However, for such systems, large roof areas are typically required in order to match the cooling demand. This circumstance makes solar heating and cooling poorly feasible in historical buildings, where the roof area available is often small. Therefore, the authors of this work proposed a novel arrangement, where the solar thermal collectors are coupled with a micro CHP system based on a reciprocating engine and an ACH (Figure 43). The system was used to supply energy to a three-storey office building located in Naples (Southern Italy), dynamically simulated in TRNSYS. The system primarily uses solar energy: the CHP unit is activated only when the tank top temperature falls below a certain set point. An electrical backup heater is also considered during peak heat demands. During the winter, heat is supplied to the zones providing space heating. Conversely, in summer heat is used by the ACH to provide space cooling.
1178 1179
Figure 43. System configuration in cooling mode [42].
1180 1181 1182 1183 1184 1185 1186 1187
The results show that, during the heating season, the solar fraction increases as a function of the tilt angle. Conversely, in summer, the opposite trend is detected. The tilt angle, along with other design parameters, were determined through a thermoeconomic optimization. Then, in the optimal configuration, the average efficiency of the solar collectors was 49%. The electrical efficiency of the cogenerator was approximately 26%; the thermal efficiency was higher than 60%. The electrical balance shows that the electricity drawn from the grid is higher than that provided by the CHP, both in winter and summer. Finally, the authors performed a parametric analysis varying the storage tank volume. Such analysis showed that solar fraction increases as a function of the storage capacity.
1188 1189 1190 1191 1192 1193 1194
The modelling and simulation of 125 m2 BIPVT collectors coupled to a 3-floor office building, for the production of solar heating by radiant floor system, solar cooling by a low-temperature adsorption chiller, solar DHW and electricity is carried out in [70]. The properties of opaque and transparent elements of the modelled building are well-representative of the traditional Italian buildings. The use of an electricity storage system, obtained by lead-acid batteries, coupled to an inverter/regulator system, is also taken into account (Figure 44). System energy, economic and environmental performances for different Italian weather zones (Turin, Florence, Naples and Palermo) are estimated.
ACCEPTED MANUSCRIPT 44 of 67
1195 1196 1197
N
Figure 44. Innovative system: building-system layout with roof BIPVT collectors (with electricity storage) and adsorption chiller [70].
1198 1199 1200 1201 1202 1203 1204
It resulted that, for the weather zone of Naples, the total PES for both electricity consumptions and DHW production resulted about 64%, whereas the PES for DHW only is equal to about 67%. An overall PES for space heating and cooling about 24% is reached. The simple payback period linearly decreases as a function of the heating degree days and increases with the incident solar radiation. The minimum is obtained in Turin (10.6 years), the maximum in Palermo (11.3 years); the avoided CO2 emissions linearly increase as a function of the incident solar radiation, reaching the maximum of about 90% in Palermo. In the remaining zones, the CO2 saving ranges from 76 to 84%.
1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216
The energy and economic performance of roof and/or façades BIPVT collectors for a multi-storey residential building consumptions is carried out in ref. [71], in order to assess the active and passive effects due to the integration of solar technologies within the building. In particular, a comparison among innovative building-plant system configurations is carried out, based on BIPVT collectors for the multi-generation of electricity, thermal energy, and DHW. The simulation models of the proposed system layouts are designed and implemented in TRNSYS simulation environment for the dynamic assessment of their energy and economic performance. Different European climates are investigated and a detailed parametric analysis by varying the thermal resistances and capacitances of the building envelope is performed. The BIPVT collectors produce a decrease of the primary energy demand ranging from 67% to 89%, depending on the weather zone and the building-plant configuration. The economic profitability resulted slightly better for roof BIPVT collectors than for roof and façade applications. For the case studies investigated, the SPB values appear quite high, varying from 11 years for South European weather zones to 20 years for the North European ones.
1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231
In ref. [72], a technical and economic evaluation is carried out for PVT systems coupled with small-scale thermally-driven solar-cooling systems (absorption refrigerator or heat pump) and thermal energy storage, to increase system autonomy. In particular, four scenarios are investigated: 1) PVT thermal output is used for space heating and DHW demands and the electricity is used to cover the space cooling demand; 2) PVT thermal output is used for DHW demand, while space cooling and heating demands are covered by means of a refrigeration cycle; 3) PVT thermal output is used for DHW demand, space heating and cooling needs are covered by an electrically-driven water-to-water HP unit; 4) PVT thermal output is used for DHW, space cooling and heating demands. The systems are investigated in ten European locations, using yearly and monthly averaged solar-irradiance and energy-demand data related to houses with a total floor area of 100 m2 (4–5 persons) and a 50 m2 roof area. Special attention is paid to the total levelized cost of combined energy generation (LCOE) in the housing sector, defined as the net present value of the unit-cost of energy provided over the lifetime (over 20 years) for a given configuration. The yearly result show that: i) Seville, Rome, Madrid and Bucharest are the most promising locations, among those examined; ii) the most efficient system configuration is achieved by coupling PVT panels to water-to-water heat pumps that use the PVT heat to increase the system COP; iii) PVT systems are able to cover 60% of the combined space heating and DHW and almost 100% of the
ACCEPTED MANUSCRIPT 45 of 67
1232 1233
cooling demands of houses for all of the four aforementioned locations; iv) the overall LCOE of these systems ranges between 0.06–0.12 €/kWh, which is 30–40% lower than that of equivalent PV-only systems.
1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246
In recent years, the problem of reducing energy consumptions and greenhouse gases emissions in buildings was also addressed by considering the technology of natural-gas-driven BCHP systems, that show high energy efficiency and low greenhouse gases emissions [128]. Such kinds of systems are investigated in the works reported in references [20, 39, 73, 123, 128]. In particular, Zhang et al. [128] studied the feasibility of thermal energy storage in BCHP system consisting of a natural-gas fired GT, an ACH, a heat exchanger and an auxiliary gas boiler (Figure 45). A hotel and an office building were considered as case studies, both located in Beijing. The authors investigated the use of a thermal storage aiming at improving the efficiency of GT and ACH, at part load conditions. In their analysis, they considered the degree of mismatch (DM), based on the relationship between user load demands and the supply thermal power ratio of the GT. In addition, the analytical relationship between PES of whole system and DM was evaluated, using two control strategies: following thermal load (FTL) and following electrical load (FEL). Three typical working conditions were considered: Heating & Electricity, Cooling & Electricity, Heating & Cooling & Electricity.
1247 1248 1249
Figure 45. Schematic diagram of building energy supply system: (a) traditional separated generation system; (b) BCHP system [128].
1250 1251 1252 1253
The results show that the more DM approaches to 1 (which means that the energy supplied by the system just equals the demand), the higher is the PES, and that the exportability of the electricity produced by the CHP unit is crucial for achieving good energy savings. The system is found more suitable for cases in which the heating demand is dominant.
1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269
Rosato et al. [123] presented a dynamic simulation, developed to evaluate the thermoeconomic and environmental performance of a building-integrated micro-cogeneration system in an Italian residential application . In this work, the transient nature of building and occupant loads, the part-load characteristics of the cogeneration system, the interaction between the loads and the system output, and system energy management and control were considered. In particular, the CHP unit is based on a 6.0 kWe natural gas-fuelled ICE producing electricity, space heating and DHW for a multi-family house consisting of three floors, located in Naples (south of Italy). The CHP unit and a natural gas-fired boiler co-operate in order to guarantee a given water temperature level (55 °C) within the tank, that has three internal heat exchangers. A group of fan-coils, supplied by the hot water storage, is installed in the building in order to balance the sensible heating load. The electric load-following and the heat demand-following control strategies were compared: it resulted that the system determines a significant reduction of primary energy consumption, CO2 emissions and operating costs. The heat demand-following control strategy appeared more profitable from both environmental and energy point of views. In particular, the simulations results revealed that, by taking into account the Italian electricity grid mix, the thermal load-following approach can offer a primary energy saving of about 6.5%, a reduction of CO2 emissions of about 12.2% and a saving on operating cost of about 20.5%; under electric load-following operation, these values become 5.1%, 10.6% and 31.0%, respectively (Figure 46).
ACCEPTED MANUSCRIPT 46 of 67
1270 1271 1272
Figure 46. Comparison between the proposed and conventional system: primary energy consumption (a), ΔCO2 equivalent emissions (b), and operating cost (c) [123].
1273 1274 1275 1276 1277 1278 1279 1280 1281
An economic and environmental evaluation of two micro CHP systems (namely: gas engine and fuel cell), for residential buildings in Japan was performed by Ren and Gao [20]. The CHP plant is driven by city gas to meet a part of the electrical demand (including cooling load with the use of air conditioning); the deficiency is served by the utility grid. The heat recovered from the micro CHP units is used for heating and hot water requirements. An auxiliary supplementary burner and a storage tank are included. Two different operating modes are taken into consideration, aiming at minimum-cost and minimum-emissions, respectively. The results show that the fuel cell is represent the best option from both economic and environmental points of view. When the operation is aimed at optimising the economic benefits, the annual energy cost is reduced of about 26%. When maximizing the environmental merits, annual CO2 emissions are reduced of about 9%.
1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298
7. Operation control strategy and optimization In many recent works on polygeneration, a special attention is paid to the comparison of different control strategies, often related to the design optimization of the system proposed [41, 50, 55, 58, 167, 168]. To address such problems, dynamic simulation tools are crucial. For example, in references [11, 73, 123, 169], different operation strategies of CHP systems were analysed, namely: rated power strategy (base-load strategy), electrical load-following and thermal load-following; hybrid strategies were also proposed, matching both thermal and electrical requests (on the basis of economic criteria). For example, in the work of Calise et al. [39], three different operating strategies were compared in order to reduce the capital cost and maximize the performance of a trigeneration system consisting of a natural gas fired reciprocating engine and an ACH, producing DHW, space heating / cooling and electricity for a real hospital building. Here, a detailed dynamic simulation model, developed in TRNSYS environment, is used to predict the system real time performance. The investigated operating strategies were: the Thermal Load Tracking strategy (TLT) - the engine partializes to follow the thermal demand, by adopting a proportional controller; the Maximum Power Thermal Load Tracking strategy (MPTLT) - the engine does not partialize and it works always at maximum power, by adopting an On/Off hysteresis controller; the Electricity Load Tracking strategy (ELT) - the engine partializes to follow the electrical request. The MPTLT control strategy, regarding the engine operation, is depicted in Figure 47.
ACCEPTED MANUSCRIPT 47 of 67
1299 1300
Figure 47. Cogenerator control strategy [48].
1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340
In particular, when the temperature coming from the sanitary hot water heat exchanger is higher or equal than 72 °C, the engine, managed by the On/Off hysteresis controller, turns off; it turns on again when such temperature is lower than 67.5 °C. If this temperature is too low, the system is managed by a suitable feedback controller that bypasses a part of the heating water flow rate that comes out from the jacket water heat exchanger. Finally, a proportional control allows the cogenerator to work at part-load operation from 50% to 100% of the maximum power, as a function of the engine inlet temperature [39]. From the carried out simulations, it resulted that the best strategy from the energy point of view is the ELT one with the higher electrical production equal to 1203 MWh/year, but the worst value of the thermal efficiency (equal to 38.9%) and global efficiency (equal to 75.6% vs 89.9% for TLT strategy and 90.3% for MPTLT strategy). PES values for the TLT, MPTLT and ELT strategies were respectively: 23, 24 and 14%. ELT strategy is the best one also for the achieved economic results being the SPB values for the ELT, MPTLT and TLT strategies respectively 4, 4.2 and 4.4 years. After the comparison of three strategies, in this work, an optimization, by the TRNEdit tool of TRSNSYS and computer Design of Experiment (DoE), selecting as objective functions the SPB and PES, for the ELT strategy (the best one) was also carried out. From the optimization results it was found out that the lowest SPB value is equal to 3.9 years and the highest PES value is 20.6%. A similar system layout is analysed in reference [73], but in this study the trigeneration system is designed to produce electricity, space heating and cooling for a real industrial building. Here, three control strategies are compared: base-load operation, electric load tracking and a new hybrid strategy based on the simultaneous tracking of electric and thermal-loads. The authors developed several detailed control strategies, including several temperature controllers in the model, able to manage the temperatures of heat recovery exchangers (jacket and flue gas loops), the activation temperature of ACH and fan coil units, the production of DHW, the activation of the engine, etc. Special attention was also paid to the new, above mentioned hybrid strategy: here the engine follows the power demand until the temperature of the DHW outlet water reaches a set-point temperature equal to 75.5 °C; then, the system is partialised at a fixed point, defined by the ratio between the power actually produced by the prime mover and its rated capacity, Pel,PM/Pel,rated,PM, until the temperature reaches a value of 69.5 °C; the fixed value of Pel,PM/Pel,rated,PM is selected to simulate the operating strategy of a real prime mover. When the partialization is activated, the electrical demand (Pel,PM,dem) is also checked, and it must be higher than a fixed ratio, defined as Pel,PM,dem/Pel,rated,PM. This last condition is set in order to obtain a minimum electrical load. Conversely, with the present energy price and legislation, the economic feasibility would be negatively affected. The main results obtained with this new control strategy are reported in Figure 48, showing, in particular: SPB, electrical economic saving (ΔCel,user), global efficiency (ηglob) and economic gain due to excess electricity sold to the grid vs. the ratio Pel,PM/Pel,rated,PM. The lines in the graphs represent the minimum value of Pel,PM,dem/Pel,rated,PM that has to be satisfied in order to activate the prime mover. For low values of the Pel,PM,dem/Pel,rated,PM, ηglob decreases, due to the longtime of part-load operation of the engine; the electricity sold to the grid significantly increases, because the production is greater than the demand. A good value of ηglob - about 0.8 - is reached for Pel,PM/Pel,rated,PM = 0.5. Conversely, for the same value a reduction of economic saving and an increase of SPB is observed. This result is mainly due to adoption of rules of the Italian electricity energy market which boost more electric – driven strategies than efficiency ones. Anyway, it is noted that with this new hybrid control strategy the best SPB is equal to 3.8 years with respect the value of 4.1 and 4.3 years of the electricity tracking and base-load operation strategy, respectively.
ACCEPTED MANUSCRIPT 48 of 67
1341 1342
Figure 48. Efficiency and economic results [73].
1343 1344 1345 1346
In the work of Taljan et al. [102] the simulation of a biomass-fired ORC CHP system with heat storage is presented; the system was optimized, too, aiming at maximizing the profits from heat sales and electricity. The objective of this study was to use optimal operation strategies to size optimally the whole plant. The optimization methodology consisted of three main steps (Figure 49).
1347 1348 1349 1350 1351 1352 1353 1354 1355
Figure 49. Methodology flowchart [102]. In particular, the optimal sizing routine calls the dispatch optimization routine which sets the hourly optimal operation and economic parameters of the system. After selecting the optimal operation parameters, an economic analysis evaluating a Modified Internal Rate of Return (MIRR) for each iteration is carried out. The MATLAB optimization toolbox and the AMPL mathematical modeling language (based on mixed-integer linear programming optimization models) are used for optimal sizing and optimal operation routine, respectively. MATLAB is also used for the calculation of economic indexes. The results show that the evaluated MIRR at the ORC rated power (100 kW), is equal to 7.80% and 8.74%, respectively, in the case without and
ACCEPTED MANUSCRIPT 49 of 67
1356 1357 1358
with thermal storage (50 m3). Therefore, more favourable economic parameters are achieved if the heat storage is not included in the system. In addition, for heat demands higher than 5 GWh/year, a MIRR of 10% was calculated, without thermal storage and at a biomass price lower than 17 €/MWh.
1359 1360 1361 1362 1363 1364 1365 1366 1367 1368
A multi-objective optimization of a multi-generation energy system, based on an evolutionary algorithm, is presented by Ahmadi et al. [170], aiming at minimizing the total cost (fuel, components purchasing and environmental impact) and maximizing the exergy efficiency. The system includes: a micro GT, a dual pressure heat recovery steam generator, an ACH, an ejector refrigeration cycle, a domestic water heater and a proton exchange membrane electrolyzer, producing power, heating, cooling, hot water and hydrogen. An equation for the Pareto optimal curve is also found, reporting the relationship between exergy efficiency and total cost rate. From this curve it resulted that the total cost rate increases moderately as the total exergy efficiency of the cycle increases to about 65%. An increase of the total exergy efficiency from 65% to 68% leads to a significant increase of the cost rate. The maximum exergy efficiency, equal to 67.89%, corresponds to the highest total cost rate, 615.75 $/h.
1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379
The polygeneration system reported in ref. [2] was also optimized in reference [56], in order to maximize the exergy performance and the economic feasibility of a solar MED polygeneration system. The plant, called Renewable Polygeneration System (RPS), consists of a solar heating and cooling system, CPVT collectors, a biomass heater and a MED unit, and provides electricity, cool and heat and desalted water. A detailed energy, exergy and economic dynamic simulation model of the whole RPS was included in the work. The optimization, based on the computer-aided Design of Experiments (DoE) technique, aims at assessing the optimal configuration by the optimization of economic and exergy objective functions. By DoE technique the sensitivity of the selected design variables the shape of the optimum response surface was also analysed. In particular, the analysis was carried out by varying the number of solar collectors and MED effects, thermal storage specific volume and aims at maximizing the profit index and the exergy efficiency and minimizing the global exergy destruction and operation costs.
1380
Table 17. Optimization results [56]. Design variables OFs PI Cop ηex Exd,total
1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397
Goal Max. Min. Max. Min.
NSC
Optimal
NEFF
vTK1
qp4
-
-
L·m−2
kg·h−1
760 1000 1000 250
14 14 9 9
25 25 25 25
475 620 100 100
Initial
Value
Unit
OFs
Value
Unit
3.43 −2480 12.80 9.46
k€ % GWh/y
PI Cop ηex Exd,total
2.64 −541 7.20 10.9
k€ % GWh/y
The optimization results (Table 17) suggest that storage tank volume should be relatively small (25 L for unit area of CPVT collectors); in addition, the optimal number of effects in the MED unit was found equal to 14, and an optimal number of 760 CPVT collectors was calculated if the Profit Index was to be maximised (with an optimal value of 3.43); differently, 1000 CPVT are required to minimize the operation costs (-2480 k€) and to maximize the global exergy efficiency (12.80%). Therefore, despite the high capital cost of the CPVT collectors, the economic optimization suggests increasing the solar field area. Anyway, for large solar fields, a high value of exergy is destroyed (87.20% of the fuel exergy entering). Consequently, it resulted that a reduction of the number of CPVT collectors minimizes the global exergy destruction. In references [124, 125], a mathematical optimization process was implemented in order to find the optimal size of a polygeneration plant fuelled by natural gas, solar energy and gasified biomass, which can provide fresh water in a tourist resort in Spain. In particular, a two-step optimization procedure was implemented. In the first one, an optimization process based on a MINLP (Mixed Integer Non-linear Programming) problem was designed to select the technology (between ICE, GT, FC, STE, RO, MED, single or double effect ACH) and the capacities of the components to be included in the polygeneration plant, by evaluating the NPV economic objective function along with energy (PES) and environmental constraints (greenhouse gases emissions). In this step, the monthly averaged requirements were taken into account. It was
ACCEPTED MANUSCRIPT 50 of 67
1398 1399 1400 1401 1402 1403 1404 1405 1406
found out that natural gas fuelled engines were the best solution from an economic point of view (NPV=347 k€), but the best value of CO2 emission reduction was obtained using solar and biomass renewable resources (3114.1 ton/year) [124]. The optimized configuration included an ICE as prime mover, a single effect ACH to cool the resort and a MED unit to produce desalted water, as well a plate-fin heat exchanger for the heating demand. In the second step of the optimization procedure [125], instead, the hourly load variability was considered and, therefore, energy storage systems were also taken into account. CONOPT module (by GAMS) was used to solve the operating optimization of the plant. In the optimized configuration, a NPV of almost 300 k€ (reduced with respect the NPV of first step one for the storage system cost), an 18% of PES and more than 850 tonCO2/year avoided emissions were achieved.
1407 1408 1409 1410 1411 1412 1413 1414 1415
A complete method regarding the design of trigeneration plants which takes into consideration operation mode (control strategy and tariff selection), energy performance (system load coverage, PES, grid utilization factor), and economic aspects (operational profit, return on investment) was presented in ref. [171]. Here, a detailed mathematical model was developed to compare different alternative solutions and to show the magnitude of the selected parameters variation on the plant performance. A new operation strategy, ‘’electrical equivalent load following (ELF)’’, was introduced. In particular, in this strategy, the equivalent required electricity Eleq is the electricity needed to cover electric demand and cooling demand that is not covered by the absorption chiller, which produces cooling energy only if waste heat from the prime mover is available. An example of this strategy is described in Figure 50.
1416 1417 1418 1419 1420 1421 1422 1423 1424 1425 1426 1427 1428 1429
Figure 50. Example of electrical-equivalent load following strategy [171]. In particular, Eleq and thermal energy produced are plotted versus the prime mover nominal size, for fixed values of the energy demands and the prime mover heat to power ratio (HPR). In summer months, Eleq decreases when the prime mover size increases, because the prime mower waste heat is used to produce cooling. During winter the thermal demand is higher than the cooling one, therefore, the proposed strategy is analogous to electrical load following strategy, since no waste heat from the prime mover to supply the ACH is available. The ELF strategy is compared with three conventional strategies, adopting the following operation assumptions: i) ELF with a minimum electricity load bought from the grid of 0%, 50% or 100% kW of the annual base load (i.e. ELFset = 0, 175 or 350 kW); ii) Continuous Operation (CO) for 8, 16, or 24 h simulating 1,2, or 3 shifts; iii) Peak Shaving (PS) of PSset = 70%, 80% or 90% of the monthly electricity load; iv) Base Load (BL) operation on BLset = 175, 350, 500 kW (50%,100%, 150% of the base load). The calculation of the PES and annual profit is performed in order to carry out the comparison among the different control strategies and adopted tariffs. The case study assumed to compare such strategies regards a trigeneration system consisting
ACCEPTED MANUSCRIPT 51 of 67
1430 1431 1432 1433 1434 1435 1436
of an ICE, ACH, auxiliary units (boiler and electrical chiller) and a heat storage. A 350 beds hospital complex located in Greece was considered. The following results were found: more than 80% of electrical, thermal and cooling demands are covered only in ELF, ELFset = 0 and CO, for 24 h with a prime mover nominal power of 0.75 MWe. A 100% coverage of thermal energy can be achieved only with the CO strategy, for 24 h. Strategies that do not achieve a PES of 10% (desired for high efficiency CHP plant) as PS or CO (one shift continuous operation) are not suitable. The lower annual profit (lower than 50 k€) is obtained for CO strategy. The higher PES value (more than 30%) and best annual profit (about 200 k€) were achieved with such ELF strategy.
1437 1438 1439 1440 1441 1442 1443 1444 1445 1446 1447 1448 1449 1450 1451 1452 1453 1454
8. Distributed generation Polygeneration systems, fuelled by renewable sources and/or conventional fossil fuels, can be furtherly classified as centralized or decentralized systems. Centralized systems are usually characterized by high capacities and are mainly supplied by fossil fuels, due to their high energy density, requiring a constant availability of the energy sources used to drive the system. Decentralized systems are often designed for isolated and remote communities (islands, rural areas), that have no connection with the utility grid, or for distributed generation (DG) coupled with advanced smart-grid systems (Figure 51). These could be on-grid (i.e.: the system operates in combination with the normal power grid to meet the demand of different energy products), or off-grid (stand-alone independent operation), for single house or communities. In DG systems, developed in contrast to the conventional centralized power systems, the input energy sources are locally available (as natural gas, solar, wind, geothermal), and the systems are located near the final users [172]. Therefore, DG is a way to promote sustainable energy development. In this case, factors like land availability, density of population and other socio-economic aspects play a crucial role in the selection of the capacity. DG implies the integration of several small-scale technologies, both fueled by renewable energies (solar power, wind power, hydro power etc.) and fossil fuels (ICE, GT, STE, FC); however, renewable sources are obviously preferred. Polygeneration systems
Decentralized
Centralized
(including or not RES)
(including or not RES)
Distributed Generation
On-Grid (including or not storage)
1455 1456
Fossil fuels only
RES only
Off-Grid Smart Micro Grid
Figure 51. Decentralized and centralized polygeneration systems.
1457 1458 1459 1460 1461 1462
Isolated or rural areas
Advantages of such systems are [173]:
self-sufficiency; reduction of the dependence on the external energy supply; local production, low transmission losses, by avoiding transportation costs; no need for expansion of costly grid interconnection; lower capital costs with respect to centralized plants;
Hybrid Fossil/RES
ACCEPTED MANUSCRIPT 52 of 67
1463 1464
1465 1466 1467 1468 1469 1470 1471 1472 1473 1474 1475 1476 1477 1478 1479 1480 1481 1482 1483 1484 1485 1486 1487 1488 1489 1490 1491 1492 1493 1494 1495
An important aspect of DG is that the system has to be able to match the energy request and production. This is the basic concept of the ‘’smart grid’’ (computerized power grid) that, equipped with advanced control and modern communication devices, facilitates the coordination between the energy supplier and the end user [7]. Significant aspects for the correct operation of smart grids are real-time measurement and monitoring, automatic energy loads adjustment, innovative control strategies, accurate assessment, proactive coordination and integrated design. Smart grids are designed to improve the current power grid and to achieve the goal of sustainability, characterized by cost and energy efficiency, reliability, and environmental friendliness [174]. This occurs for the electrical, thermal (heating, DHW) and cold production, distributed separately via local power grid, the district heating (DH) and district cooling (DC) network, respectively, although production of thermal energy and electricity are coupled within the polygeneration plant. Key components for the correct operation of the DG systems are the electrical and thermal storage facilities (electrical vehicle recharge station, electrical battery storage, storage hot and cold tanks), included in order to balance the user demand and energy production, to store the eventual extra-production, by mitigating the effects of renewable energy intermittency. In this framework, by considering the power production by the renewable energy sources as solar and / or wind energy, which are definitely intermittent and variable, several studies were focused on the dynamic nature of the problem caused by charging and discharging of different types of electricity storage facilities, which are important components in the smart polygeneration micro grid. The same importance is assigned to the thermal energy storage, in case of cooling and heating energy production, which can be prepared and stored in storage at off-peak hours and released into district heating and cooling (DHC) network from storage during peak hours [5]. In order to improve the DG systems costs and overall efficiency, the connection to normal long-distance power transmission network (on-grid DG systems) is also possible. This configuration also allows one to overcome issues as the excess production typical of some countries, which are featured by a considerable production due to the large solar field or wind farms. In this way, the power exchanging of DG systems with the normal network of any energy surplus is a feasible and sustainable solution. A further aspect to improve the overall efficiency of DG systems is represented by the proper selection of DG systems scale. Generally, DG polygeneration systems using renewable energy are characterised by small capacities, mainly depending on the availability and distribution of the local resources. From this point of view, DG systems should have a suitable and not too small capacity; however, the final global efficiency could be not high enough if compared with the conventional fossil or renewable energy polygeneration systems or with the separate production plants.
1496 1497 1498 1499
In the following, some of the works reporting analysis performed on the DG systems, carried out by adopting different optimization techniques [175-179] aiming at contributing their spread in the sector of the sustainable polygeneration, are reported. In these works, the importance of the simulation models, as key tools for evaluating the performance of complex systems, as are the DG systems, is also underlined.
1500 1501 1502 1503 1504 1505 1506 1507 1508 1509 1510 1511 1512 1513 1514
As many researchers highlighted, the most widespread DG plants are based on the trigeneration systems integrated into DHC networks, by focusing on the optimisation of the energy conversion technologies and their operational strategies. With this kind of system, the layout may be trivial if compared with complex polygeneration systems consisting of a plurality of different technologies. However, the control of such systems is not simple, since thermal, cooling and power demands may considerably vary during the day. This issue was investigated by Conte et al. [51], who addressed the optimal management of trigeneration systems including ACH for cooling production. Their study develops from the literature analysis and shows that, in most cases, flexibility and modularity of efficient and recent trigeneration systems are not adequate to correctly work in DHC at very low partial load ratios. Trigeneration plants usually integrate several chiller devices (double and single effect ACHs, vapour compression): the time-dependent cooling load should be redistributed among such types of chillers aiming at achieving the lowest operating cost and the maximum efficiency. In traditional systems, the management of such chillers is obtained by a sequential approach. This approach is normally employed to obtain a constant chiller flow rate and to reach the same outlet chilled water temperature among the chillers. Nevertheless, this control strategy is rather simple, and it may be considerably improved adopting optimized sharing load methods, as shown in several available literatures works. The novelty
reduction of the fossil fuels dependency and the mitigation of environmental impact; reduction of breakdown effects of a single unit on the whole energy network.
ACCEPTED MANUSCRIPT 53 of 67
1515 1516 1517 1518 1519 1520 1521 1522 1523 1524 1525 1526 1527 1528 1529
of the study of Conte et al. [51] is represented by the use of real data taken from an existing trigeneration plant located near Barcelona (Spain). Cooling and thermal capacities are 8.75 MW and 1.40 MW, respectively. The plant consists of three ICEs, each one with an electric capacity equal to 3.35 MW. The heat produced by waste gases supply double-effect ACHs, whereas the jacket water recovered heat is used to drive a single-effect ACH. The operation of system is about 5000 h / year. The system was carefully monitored by flow rates, temperatures and heat flows measuring. A mixed integer linear programming (MILP) approach was adopted to optimize system operation, whereas a general algebraic modelling system (GAMS) is used to model the whole trigeneration system. The authors examined the load sharing strategy for a typical day, by changing the distribution of load as a function of the different chillers part-load performance. Results showed that the single-effect ACH has the lowest efficiency, but its operation is useful for reducing the cost of engine cooling energy rejection and maximizing engine heat recovery. Double and single-effect ACH COP were 1.0 and 0.65, respectively. Such results are satisfactory, by taking into account that such devices usually operate at 40% of their rated capacity. The optimization performed by adopting economic objective functions indicated that the double-effect ACH operation should be preferred to that of the single-effect one.
1530 1531 1532 1533 1534 1535 1536 1537 1538 1539 1540 1541 1542 1543 1544 1545 1546 1547
The optimization of a distributed cogeneration system is performed in ref. [52]. The plant consists of a DH network, a solar thermal plant, a long-term heat storage, a set of CHP units and conventional components, such as boilers and compression chillers. The considered CHP units were a centralized ICE and small-scale (ICE or MTG) system, properly located close to, or inside, nine industrial facilities located in the North-east of Italy. The optimal energy production system and operation strategy needed to satisfy the energy demand and to minimize the total annual cost were identified, using a MILP model (developed with the commercial software X-press). Different configurations were considered, namely: i) conventional solution; ii) isolated solution with CHP and CCHP systems for partially replacing conventional boilers and compression chillers; iii) distributed cogeneration solution with all users connected each other through a DH network and by adopting CHP and CCHP systems; iv) distributed renewable solution, like the previous one, but with the addition of a centralized solar thermal plant and a thermal storage. Note that in this analysis the objective function, total annual cost, simply corresponds to the heat cost because the cost of electricity and cooling energy are considered constant for the specific set of analysed users. The annual cost for the configuration without solar field and the DH network, with both boilers and cogenerators, allows a 4.5% reduction of the total annual cost, with respect to the conventional solution (5.47 c€/kWh). A 5% reduction is obtained by adopting a distributed solar system, which includes the DH network, the thermal storage and the solar field. In this configuration, a 15% reduction of the primary energy consumption is also achieved. In addition, the optimally sized solar field produces about 55–60% of the user annual thermal demand.
1548 1549 1550 1551 1552 1553 1554 1555 1556 1557 1558 1559 1560 1561 1562
A main aspect is represented by the allocation of DG sub-systems. This issue is dealt with by Peng et al. [53] presenting a novel approach to resolve the problem of the optimal allocation of DG systems. In this work, a modelling of a PV panels field and wind turbines by using Beta and Weibull distributions, respectively, was carried out. The load was modelled using a normal distribution. The optimal DG allocation (ODGA) was investigated using a multi-objective approach integrating as objective functions total DG costs, pollutant emissions and system power losses. Obviously, the investigated system is subject to both chance and deterministic constraints. This problem was analysed by using the crisscross optimization algorithm with Monte Carlo simulation. The crisscross optimization is used to detect the optimal solution among a set of feasible solutions, whereas the Monte Carlo method is used to deal with system uncertainties and to solve the power flow problem. This novel approach was tested on two benchmark cases widely used in ODGA problems, namely IEEE 33-bus and IEEE 69-bus systems. The novel algorithm is shown to be about 25% faster than particle swarm optimization (PSO). The novel approach also ensured a better stability of the voltage with respect to the case of the PSO technique. In summary, the development of such optimization techniques is very useful to design polygeneration systems including many different power plants, especially when powered by renewable energy sources.
1563 1564 1565 1566
The optimal design was studied by Melheri et al. [180], taking into account both the capacity and allocation of DG systems for the distribution of power and heating demand at the level of a small neighbourhood. The optimization of a micro-grid configuration in combination with the design of a heating pipeline network, by investigating different distributed energy technologies (PV, FC, STE, gas engine, boiler), is carried out aiming
ACCEPTED MANUSCRIPT 54 of 67
1567 1568 1569 1570 1571 1572 1573 1574 1575 1576
at minimizing the investment and operational cost of the overall DG system. Special interest was paid to the design of the heating pipeline network that serves the transportation of heat between the different nodes, where the buildings are located. This was performed by a MILP model, solved by using the GAMS CPLEX solver. The developed optimization model was solved for several cases study (i.e. ten and twenty buildings with and without heat exchange between the buildings, ten buildings with centralized CHP, ten buildings without distributed technologies and heating networks) where the neighbourhood is assumed to be located in Greece. Considering ten buildings and heating distribution networks, the total annualised cost of optimal solution is € 4,742.7, with CHP units of capacity of 4.7 kWe installed in buildings i2, i3, i6, i10, without back-up boilers and PV arrays installed in all buildings. Instead, the optimal heating pipeline network consists of 2 sub networks (Figure 52).
1577 1578 1579 1580 1581 1582 1583 1584 1585 1586 1587 1588 1589 1590 1591 1592 1593 1594 1595 1596 1597 1598 1599 1600 1601 1602 1603 1604 1605
Figure 52. Heating pipeline network for the “10 Buildings” case. [180]. In ref. [181], several detailed simulation models, aiming at showing their accuracy and influence on optimal operation cost and strategy, with respect the simplified models, for a distributed CCHP system are proposed. The configuration adopted to perform the comparison consists of a natural gas ICE, ACH, and auxiliary devises as an electrical chiller and a boiler. If the ICE electricity production is not sufficient to satisfy the electrical demand (an office building in Tianjin, China), the possibility to use electricity public grid is simulated. No sold for surplus electricity is expected. Results show that 2240 ¥/day, accounting for 2.5% of the total daily cost, is reduced when the detailed model is used to formulate the operation strategy, demonstrating that the detailed model is superior to the simplified model when the economy and rationality of the optimal operation strategy are considered. 9. Conclusions Due to the recent development of small-scale energy technologies and renewable energy sources, energy systems are rapidly changing from the conventional centralized model to new, de-centralized configurations, mostly focused on the use of renewable sources and, therefore, including energy storage devices, needed to face the variability of most renewable sources. In this framework, a very promising solution is represented by polygeneration systems, using multiple energy sources to generate multiple products (electricity, thermal energy, hydrogen, ammonia, etc.). In fact, such systems are potentially able to ensure high energy efficiency, environmental compatibility and high flexibility, and can be profitable even from an economic point of view: for example, they can be more efficient in facing fluctuations of energy prices and changes in end-product demand. In other terms, the systems for DG, that are often intended only for power production, can become much more energy-efficient and interesting when many technologies are combined to simultaneously provide heat, power and other energy products within a single, integrated process. Due to their intrinsic complexity and to the large amount of different possible configurations, an important research effort is needed on polygeneration systems, in order to correctly address their design, optimization and control and to develop efficient decision support tools for designers, policy makers and end users. This is one of the reasons for the recent, significant growth in the number of scientific papers dealing with this type of system. This paper intends to contribute to such research effort.
ACCEPTED MANUSCRIPT 55 of 67
1606
Conflicts of Interest: The authors declare no conflict of interest.
1607
References
1608 1609 1610 1611 1612 1613 1614 1615 1616 1617 1618 1619 1620 1621 1622 1623 1624 1625 1626 1627 1628 1629 1630 1631 1632 1633 1634 1635 1636 1637 1638 1639 1640 1641 1642 1643 1644 1645 1646 1647 1648
1. 2.
3.
4. 5.
6. 7. 8.
9. 10.
11.
12.
13.
14.
15.
Agency, I.-I. E., World Energy Outlook 2013. 2013. Calise, F.; Dentice d'Accadia, M.; Piacentino, A., A novel solar trigeneration system integrating PVT (photovoltaic/thermal collectors) and SW (seawater) desalination: Dynamic simulation and economic assessment. Energy 2014, 67, 129-148. Calise, F.; Cipollina, A.; Dentice d’Accadia, M.; Piacentino, A., A novel renewable polygeneration system for a small Mediterranean volcanic island for the combined production of energy and water: Dynamic simulation and economic assessment. Applied Energy 2014, 135, 675-693. Serra, L. M.; Lozano, M.-A.; Ramos, J.; Ensinas, A. V.; Nebra, S. A., Polygeneration and efficient use of natural resources. Energy 2009, 34, (5), 575-586. Rong, A.; Lahdelma, R., Role of polygeneration in sustainable energy system development challenges and opportunities from optimization viewpoints. Renewable and Sustainable Energy Reviews 2016, 53, 363-372. Murugan, S.; Horák, B., Tri and polygeneration systems - A review. Renewable and Sustainable Energy Reviews 2016, 60, 1032-1051. Rong, A.; Su, Y., Polygeneration systems in buildings: A survey on optimization approaches. Energy and Buildings 2017, 151, 439-454. Jana, K.; Ray, A.; Majoumerd, M. M.; Assadi, M.; De, S., Polygeneration as a future sustainable energy solution – A comprehensive review. Applied Energy 2017, 202, 88-111. Calise, F.; D’Accadia, M. D., Simulation of Polygeneration Systems. Energies 2016, 9, (11). Karellas, S.; Braimakis, K., Energy–exergy analysis and economic investigation of a cogeneration and trigeneration ORC–VCC hybrid system utilizing biomass fuel and solar power. Energy Conversion and Management 2016, 107, 103-113. Sibilio, S.; Rosato, A.; Ciampi, G.; Scorpio, M.; Akisawa, A., Building-integrated trigeneration system: Energy, environmental and economic dynamic performance assessment for Italian residential applications. Renewable and Sustainable Energy Reviews 2017, 68, Part 2, 920-933. Soutullo, S.; Bujedo, L. A.; Samaniego, J.; Borge, D.; Ferrer, J. A.; Carazo, R.; Heras, M. R., Energy performance assessment of a polygeneration plant in different weather conditions through simulation tools. Energy and Buildings 2016, 124, 7-18. Chen, Q.; Han, W.; Zheng, J.-j.; Sui, J.; Jin, H.-g., The exergy and energy level analysis of a combined cooling, heating and power system driven by a small scale gas turbine at off design condition. Applied Thermal Engineering 2014, 66, (1), 590-602. Rosato, A.; Sibilio, S.; Ciampi, G.; Entchev, E.; Ribberink, H., Energy, Environmental and Economic Effects of Electric Vehicle Charging on the Performance of a Residential Building-integrated Micro-trigeneration System. Energy Procedia 2017, 111, 699-709. Jana, K.; De, S., Polygeneration using agricultural waste: Thermodynamic and economic feasibility study. Renewable Energy 2015, 74, 648-660.
ACCEPTED MANUSCRIPT 56 of 67
1649 1650 1651 1652 1653 1654 1655 1656 1657 1658 1659 1660 1661 1662 1663 1664 1665 1666 1667 1668 1669 1670 1671 1672 1673 1674 1675 1676 1677 1678 1679 1680 1681 1682 1683 1684 1685 1686 1687 1688 1689 1690 1691
16.
17.
18.
19.
20.
21.
22.
23.
24. 25.
26.
27. 28.
29.
30.
Calise, F.; Dentice d'Accadia, M.; Macaluso, A.; Vanoli, L.; Piacentino, A., A novel solar-geothermal trigeneration system integrating water desalination: Design, dynamic simulation and economic assessment. Energy 2016, 115, Part 3, 1533-1547. Lin, H.; Jin, H.; Gao, L.; Han, W., Techno-economic evaluation of coal-based polygeneration systems of synthetic fuel and power with CO2 recovery. Energy Conversion and Management 2011, 52, (1), 274-283. Buonomano, A.; Calise, F.; Palombo, A.; Vicidomini, M., Energy and economic analysis of geothermal–solar trigeneration systems: A case study for a hotel building in Ischia. Applied Energy 2015, 138, 224-241. Calise, F., Design of a hybrid polygeneration system with solar collectors and a Solid Oxide Fuel Cell: Dynamic simulation and economic assessment. International Journal of Hydrogen Energy 2011, 36, (10), 6128-6150. Ren, H.; Gao, W., Economic and environmental evaluation of micro CHP systems with different operating modes for residential buildings in Japan. Energy and Buildings 2010, 42, (6), 853-861. Ahmadi, P.; Rosen, M. A.; Dincer, I., Greenhouse gas emission and exergoenvironmental analyses of a trigeneration energy system. International Journal of Greenhouse Gas Control 2011, 5, (6), 1540-1549. Ahmadi, P.; Dincer, I.; Rosen, M. A., Exergo-environmental analysis of an integrated organic Rankine cycle for trigeneration. Energy Conversion and Management 2012, 64, 447-453. Duic, N.; Guzovic, Z.; Kafarov, V.; Klemes, J. J.; Mathiessen, B. v.; Yan, J., Sustainable development of energy, water and environment systems. Applied Energy 2013, 101, 3-5. Jana, K.; De, S., Sustainable polygeneration design and assessment through combined thermodynamic, economic and environmental analysis. Energy 2015, 91, 540-555. Wang, J.; Yang, Y., Energy, exergy and environmental analysis of a hybrid combined cooling heating and power system utilizing biomass and solar energy. Energy Conversion and Management 2016, 124, 566-577. Basrawi, F.; Yamada, T.; Obara, S. y., Economic and environmental based operation strategies of a hybrid photovoltaic–microgas turbine trigeneration system. Applied Energy 2014, 121, 174-183. Jana, K.; De, S., Environmental impact of biomass based polygeneration - A case study through life cycle assessment. Bioresour Technol 2017, 256-265. Kabalina, N.; Costa, M.; Yang, W.; Martin, A.; Santarelli, M., Exergy analysis of a polygeneration-enabled district heating and cooling system based on gasification of refuse derived fuel. Journal of Cleaner Production 2017, 141, 760-773. Dai, Y.; Wang, J.; Gao, L., Exergy analysis, parametric analysis and optimization for a novel combined power and ejector refrigeration cycle. Applied Thermal Engineering 2009, 29, (10), 1983-1990. Sharaf, M. A.; Nafey, A. S.; García-Rodríguez, L., Exergy and thermo-economic analyses of a combined solar organic cycle with multi effect distillation (MED) desalination process. Desalination 2011, 272, (1–3), 135-147.
ACCEPTED MANUSCRIPT 57 of 67
1692 1693 1694 1695 1696 1697 1698 1699 1700 1701 1702 1703 1704 1705 1706 1707 1708 1709 1710 1711 1712 1713 1714 1715 1716 1717 1718 1719 1720 1721 1722 1723 1724 1725 1726 1727 1728 1729 1730 1731 1732 1733
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42. 43.
Akkaya, A. V.; Sahin, B.; Huseyin Erdem, H., An analysis of SOFC/GT CHP system based on exergetic performance criteria. International Journal of Hydrogen Energy 2008, 33, (10), 2566-2577. Hegner, R.; Atakan, B., A polygeneration process concept for HCCI-engines – Modeling product gas purification and exergy losses. International Journal of Hydrogen Energy 2017, 42, (2), 1287-1297. Calise, F.; Dentice d'Accadia, M.; Piacentino, A., Exergetic and exergoeconomic analysis of a renewable polygeneration system and viability study for small isolated communities. Energy 2015, 92, Part 3, 290-307. Li, Z.; Liu, P.; He, F.; Wang, M.; Pistikopoulos, E. N., Simulation and exergoeconomic analysis of a dual-gas sourced polygeneration process with integrated methanol/DME/DMC catalytic synthesis. Computers & Chemical Engineering 2011, 35, (9), 1857-1862. Yilmazoglu, M. Z., Effects of the selection of heat transfer fluid and condenser type on the performance of a solar thermal power plant with technoeconomic approach. Energy Conversion and Management 2016, 111, 271-278. Elsafi, A. M., Exergy and exergoeconomic analysis of sustainable direct steam generation solar power plants. Energy Conversion and Management 2015, 103, 338347. Calise, F.; d’Accadia, M. D.; Macaluso, A.; Piacentino, A.; Vanoli, L., Exergetic and exergoeconomic analysis of a novel hybrid solar–geothermal polygeneration system producing energy and water. Energy Conversion and Management 2016, 115, 200220. Ghaebi, H.; Saidi, M. H.; Ahmadi, P., Exergoeconomic optimization of a trigeneration system for heating, cooling and power production purpose based on TRR method and using evolutionary algorithm. Applied Thermal Engineering 2012, 36, 113-125. Calise, F.; Macaluso, A.; Piacentino, A.; Vanoli, L., A novel hybrid polygeneration system supplying energy and desalinated water by renewable sources in Pantelleria Island. Energy 2017. Arsalis, A.; Alexandrou, A. N.; Georghiou, G. E., Thermoeconomic Modeling and Parametric Study of a Photovoltaic-Assisted 1MWe Combined Cooling, Heating, and Power System. Energies 2016, 9, 663. Calise, F.; Dentice d'Accadia, M.; Figaj, R. D.; Vanoli, L., A novel solar-assisted heat pump driven by photovoltaic/thermal collectors: Dynamic simulation and thermoeconomic optimization. Energy 2016, 95, 346-366. Marrasso, E.; Roselli, C.; Sasso, M.; Tariello, F., Analysis of a Hybrid Solar-Assisted Trigeneration System. Energies 2016, 9, 705. 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 Conversion and Management 2015, 90, 347-363.
ACCEPTED MANUSCRIPT 58 of 67
1734 1735 1736 1737 1738 1739 1740 1741 1742 1743 1744 1745 1746 1747 1748 1749 1750 1751 1752 1753 1754 1755 1756 1757 1758 1759 1760 1761 1762 1763 1764 1765 1766 1767 1768 1769 1770 1771 1772 1773 1774 1775
44.
45.
46. 47.
48.
49.
50.
51. 52. 53.
54.
55.
56.
57.
58.
Calise, F.; Capuozzo, C.; Carotenuto, A.; Vanoli, L., Thermoeconomic analysis and off-design performance of an organic Rankine cycle powered by mediumtemperature heat sources. Solar Energy 2014, 103, 595-609. Patel, B.; Desai, N. B.; Kachhwaha, S. S.; Jain, V.; Hadia, N., Thermo-economic analysis of a novel organic Rankine cycle integrated cascaded vapor compression– absorption system. Journal of Cleaner Production 2017, 154, 26-40. Menon, R. P.; Paolone, M.; Maréchal, F., Study of optimal design of polygeneration systems in optimal control strategies. Energy 2013, 55, 134-141. Ferrari, M. L.; Pascenti, M.; Sorce, A.; Traverso, A.; Massardo, A. F., Real-time tool for management of smart polygeneration grids including thermal energy storage. Applied Energy 2014, 130, 670-678. Calise, F.; Dentice d'Accadia, M.; Libertini, L.; Quiriti, E.; Vicidomini, M., A novel tool for thermoeconomic analysis and optimization of trigeneration systems: A case study for a hospital building in Italy. Energy 2017, 126, 64-87. Calise, F.; Libertini, L.; Vicidomini, M., Design and optimization of a novel solar cooling system for combined cycle power plants. Journal of Cleaner Production 2017, 161, 1385-1403. Piacentino, A.; Barbaro, C.; Cardona, F., Promotion of polygeneration for buildings applications through sector- and user-oriented “high efficiency CHP” eligibility criteria. Applied Thermal Engineering 2014, 71, (2), 882-894. Conte, B.; Bruno, J. C.; Coronas, A., Optimal Cooling Load Sharing Strategies for Different Types of Absorption Chillers in Trigeneration Plants. Energies 2016, 9. Buoro, D.; Pinamonti, P.; Reini, M., Optimization of a Distributed Cogeneration System with solar district heating. Applied Energy 2014, 124, 298-308. Peng, X.; Lin, L.; Zheng, W.; Liu, Y., Crisscross Optimization Algorithm and Monte Carlo Simulation for Solving Optimal Distributed Generation Allocation Problem. Energies 2015, 8, 13641–13659. Rivarolo, M.; Cuneo, A.; Traverso, A.; Massardo, A. F., Design optimisation of smart poly-generation energy districts through a model based approach. Applied Thermal Engineering 2016, 99, 291-301. Kyriakarakos, G.; Dounis, A. I.; Rozakis, S.; Arvanitis, K. G.; Papadakis, G., Polygeneration microgrids: A viable solution in remote areas for supplying power, potable water and hydrogen as transportation fuel. Applied Energy 2011, 88, (12), 4517-4526. Calise, F.; Dentice d’Accadia, M.; Piacentino, A.; Vicidomini, M., Thermoeconomic Optimization of a Renewable Polygeneration System Serving a Small Isolated Community. Energies 2015, 8, (2), 995-1024. Calise, F.; d’Accadia, M. D.; Vanoli, L., Design and dynamic simulation of a novel solar trigeneration system based on hybrid photovoltaic/thermal collectors (PVT). Energy Conversion and Management 2012, 60, 214-225. Khojasteh Salkuyeh, Y.; Adams, T. A., Integrated petroleum coke and natural gas polygeneration process with zero carbon emissions. Energy 2015, 91, 479-490.
ACCEPTED MANUSCRIPT 59 of 67
1776 1777 1778 1779 1780 1781 1782 1783 1784 1785 1786 1787 1788 1789 1790 1791 1792 1793 1794 1795 1796 1797 1798 1799 1800 1801 1802 1803 1804 1805 1806 1807 1808 1809 1810 1811 1812 1813 1814 1815 1816 1817 1818
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
Bai, Z.; Liu, Q.; Lei, J.; Li, H.; Jin, H., A polygeneration system for the methanol production and the power generation with the solar–biomass thermal gasification. Energy Conversion and Management 2015, 102, 190-201. Wang, Z.; Han, W.; Zhang, N.; Liu, M.; Jin, H., Proposal and assessment of a new CCHP system integrating gas turbine and heat-driven cooling/power cogeneration. Energy Conversion and Management 2017, 144, 1-9. De Kam, M. J.; Vance Morey, R.; Tiffany, D. G., Biomass Integrated Gasification Combined Cycle for heat and power at ethanol plants. Energy Conversion and Management 2009, 50, (7), 1682-1690. Maraver, D.; Uche, J.; Royo, J., Assessment of high temperature organic Rankine cycle engine for polygeneration with MED desalination: A preliminary approach. Energy Conversion and Management 2012, 53, (1), 108-117. Chen, Y.; Han, W.; Jin, H., Investigation of an ammonia-water combined power and cooling system driven by the jacket water and exhaust gas heat of an internal combustion engine. International Journal of Refrigeration 2017, 82, 174-188. Klein, S. A.; Beckman, W. A.; Mitchell, J. W.; Duffie, J. A.; Duffie, N. A.; Freeman, T. L.; et al, Solar Energy Laboratory, TRNSYS. A transient system simulation program. University of Wisconsin, Madison 2006. Calise, F.; Palombo, A.; Vanoli, L., Design and dynamic simulation of a novel polygeneration system fed by vegetable oil and by solar energy. Energy Conversion and Management 2012, 60, 204-213. Calise, F., High temperature solar heating and cooling systems for different Mediterranean climates: Dynamic simulation and economic assessment. Applied Thermal Engineering 2012, 32, 108-124. Buonomano, A.; Calise, F.; Ferruzzi, G.; Vanoli, L., A novel renewable polygeneration system for hospital buildings: Design, simulation and thermoeconomic optimization. Applied Thermal Engineering 2014, 67, (1–2), 43-60. Calise, F., Thermoeconomic analysis and optimization of high efficiency solar heating and cooling systems for different Italian school buildings and climates. Energy and Buildings 2010, 42, (7), 992-1003. Calise, F.; Figaj, R. D.; Massarotti, N.; Mauro, A.; Vanoli, L., Polygeneration system based on PEMFC, CPVT and electrolyzer: Dynamic simulation and energetic and economic analysis. Applied Energy 2017, 192, 530-542. Buonomano, A.; Calise, F.; Palombo, A.; Vicidomini, M., Adsorption chiller operation by recovering low-temperature heat from building integrated photovoltaic thermal collectors: Modelling and simulation. Energy Conversion and Management 2017, 149, 1019-1036. Buonomano, A.; Calise, F.; Palombo, A.; Vicidomini, M., BIPVT systems for residential applications: An energy and economic analysis for European climates. Applied Energy 2016, 184, 1411-1431. Ramos, A.; Chatzopoulou, M. A.; Guarracino, I.; Freeman, J.; Markides, C. N., Hybrid photovoltaic-thermal solar systems for combined heating, cooling and power provision in the urban environment. Energy Conversion and Management 2017.
ACCEPTED MANUSCRIPT 60 of 67
1819 1820 1821 1822 1823 1824 1825 1826 1827 1828 1829 1830 1831 1832 1833 1834 1835 1836 1837 1838 1839 1840 1841 1842 1843 1844 1845 1846 1847 1848 1849 1850 1851 1852 1853 1854 1855 1856 1857 1858 1859 1860 1861
73.
74.
75.
76.
77.
78. 79. 80.
81. 82. 83.
84.
85.
86. 87.
Calise, F.; Dentice d'Accadia, M.; Libertini, L.; Quiriti, E.; Vanoli, R.; Vicidomini, M., Optimal operating strategies of combined cooling, heating and power systems: A case study for an engine manufacturing facility. Energy Conversion and Management 2017, 149, 1066-1084. Calise, F.; Ferruzzi, G.; Vanoli, L., Transient simulation of polygeneration systems based on PEM fuel cells and solar heating and cooling technologies. Energy 2012, 41, (1), 18-30. Ge, Y. T.; Tassou, S. A.; Chaer, I.; Suguartha, N., Performance evaluation of a trigeneration system with simulation and experiment. Applied Energy 2009, 86, (11), 2317-2326. Mohan, G.; Kumar, U.; Pokhrel, M. K.; Martin, A., A novel solar thermal polygeneration system for sustainable production of cooling, clean water and domestic hot water in United Arab Emirates: Dynamic simulation and economic evaluation. Applied Energy 2016, 167, 173-188. Sahoo, U.; Kumar, R.; Pant, P. C.; Chaudhary, R., Development of an innovative polygeneration process in hybrid solar-biomass system for combined power, cooling and desalination. Applied Thermal Engineering 2017, 120, 560-567. Khalid, F.; Dincer, I.; Rosen, M. A., Energy and exergy analyses of a solar-biomass integrated cycle for multigeneration. Solar Energy 2015, 112, 290-299. Bicer, Y.; Dincer, I., Analysis and performance evaluation of a renewable energy based multigeneration system. Energy 2016, 94, 623-632. Tempesti, D.; Manfrida, G.; Fiaschi, D., Thermodynamic analysis of two micro CHP systems operating with geothermal and solar energy. Applied Energy 2012, 97, 609617. Suleman, F.; Dincer, I.; Agelin-Chaab, M., Development of an integrated renewable energy system for multigeneration. Energy 2014, 78, (Supplement C), 196-204. Al-Ali, M.; Dincer, I., Energetic and exergetic studies of a multigenerational solar– geothermal system. Applied Thermal Engineering 2014, 71, (1), 16-23. Islam, S.; Dincer, I., Development, analysis and performance assessment of a combined solar and geothermal energy-based integrated system for multigeneration. Solar Energy 2017, 147, (Supplement C), 328-343. Ozturk, M.; Dincer, I., Thermodynamic assessment of an integrated solar power tower and coal gasification system for multi-generation purposes. Energy Conversion and Management 2013, 76, (Supplement C), 1061-1072. Carles Bruno, J.; Valero, A.; Coronas, A., Performance analysis of combined microgas turbines and gas fired water/LiBr absorption chillers with post-combustion. Applied Thermal Engineering 2005, 25, (1), 87-99. Al-Sulaiman, F. A.; Dincer, I.; Hamdullahpur, F., Exergy modeling of a new solar driven trigeneration system. Solar Energy 2011, 85, (9), 2228-2243. Medrano, M.; Mauzey, J.; McDonell, V.; Samuelsen, S.; Boer, D., Theoretical analysis of a novel integrated energy system formed by a microturbine and an exhaust fired single-double effect absorption chiller. International Journal of Thermodynamics 2006, 9, 29-36.
ACCEPTED MANUSCRIPT 61 of 67
1862 1863 1864 1865 1866 1867 1868 1869 1870 1871 1872 1873 1874 1875 1876 1877 1878 1879 1880 1881 1882 1883 1884 1885 1886 1887 1888 1889 1890 1891 1892 1893 1894 1895 1896 1897 1898 1899 1900 1901 1902 1903
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
Uris, M.; Linares, J. I.; Arenas, E., Feasibility assessment of an Organic Rankine Cycle (ORC) cogeneration plant (CHP/CCHP) fueled by biomass for a district network in mainland Spain. Energy 2017, 133, 969-985. Li, Y.; Zhang, G.; Yang, Y.; Zhai, D.; Zhang, K.; Xu, G., Thermodynamic analysis of a coal-based polygeneration system with partial gasification. Energy 2014, 72, 201-214. Bose, A.; Jana, K.; Mitra, D.; De, S., Co-production of power and urea from coal with CO2 capture: performance assessment. Clean Techn Environ Policy 2015, 12711280. Fan, J.; Hong, H.; Zhu, L.; Wang, Z.; Jin, H., Thermodynamic evaluation of chemical looping combustion for combined cooling heating and power production driven by coal. Energy Conversion and Management 2017, 135, 200-211. Gao, L.; Li, H.; Chen, B.; Jin, H.; Lin, R.; Hong, H., Proposal of a natural gas-based polygeneration system for power and methanol production. Energy 2008, 33, (2), 206-212. Yi, Q.; Feng, J.; Li, W. Y., Optimization and efficiency analysis of polygeneration system with coke-oven gas and coal gasified gas by Aspen Plus. Fuel 2012, 96, (Supplement C), 131-140. Jana, K.; De, S., Techno-economic evaluation of a polygeneration using agricultural residue – A case study for an Indian district. Bioresource Technology 2015, 181, 163173. Becker, W. L.; Braun, R. J.; Penev, M.; Melaina, M., Design and technoeconomic performance analysis of a 1MW solid oxide fuel cell polygeneration system for combined production of heat, hydrogen, and power. Journal of Power Sources 2012, 200, 34-44. Meerman, J. C.; Ramírez, A.; Turkenburg, W. C.; Faaij, A. P. C., Performance of simulated flexible integrated gasification polygeneration facilities. Part A: A technical-energetic assessment. Renewable and Sustainable Energy Reviews 2011, 15, (6), 2563-2587. Ng, K. S.; Zhang, N.; Sadhukhan, J., Techno-economic analysis of polygeneration systems with carbon capture and storage and CO2 reuse. Chemical Engineering Journal 2013, 219, (Supplement C), 96-108. Khojasteh Salkuyeh, Y.; Adams, T. A., A novel polygeneration process to co-produce ethylene and electricity from shale gas with zero CO2 emissions via methane oxidative coupling. Energy Conversion and Management 2015, 92, (Supplement C), 406-420. Eveloy, V.; Rodgers, P.; Qiu, L., Performance investigation of a power, heating and seawater desalination poly-generation scheme in an off-shore oil field. Energy 2016, 98, 26-39. Pan, M.; Aziz, F.; Li, B.; Perry, S.; Zhang, N.; Bulatov, I.; Smith, R., Application of optimal design methodologies in retrofitting natural gas combined cycle power plants with CO2 capture. Applied Energy 2016, 161, 695-706.
ACCEPTED MANUSCRIPT 62 of 67
1904 1905 1906 1907 1908 1909 1910 1911 1912 1913 1914 1915 1916 1917 1918 1919 1920 1921 1922 1923 1924 1925 1926 1927 1928 1929 1930 1931 1932 1933 1934 1935 1936 1937 1938 1939 1940 1941 1942 1943 1944 1945 1946
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111. 112.
113.
114.
Ray, A.; Jana, K.; De, S., Polygeneration for an off-grid Indian village: Optimization by economic and reliability analysis. Applied Thermal Engineering 2017, 116, 182196. Taljan, G.; Verbič, G.; Pantoš, M.; Sakulin, M.; Fickert, L., Optimal sizing of biomass-fired Organic Rankine Cycle CHP system with heat storage. Renewable Energy 2012, 41, 29-38. Huicochea, A.; Rivera, W.; Gutiérrez-Urueta, G.; Bruno, J. C.; Coronas, A., Thermodynamic analysis of a trigeneration system consisting of a micro gas turbine and a double effect absorption chiller. Applied Thermal Engineering 2011, 31, (16), 3347-3353. Rahman, M.; Malmquist, A., Modeling and Simulation of an Externally Fired MicroGas Turbine for Standalone Polygeneration Application. Journal of Engineering for Gas Turbines and Power 2016, 138, (11). Sarr, J.-A. R.; Mathieu-Potvin, F., Increasing thermal efficiency of Rankine cycles by using refrigeration cycles: A theoretical analysis. Energy Conversion and Management 2016, 121, 358-379. Mata-Torres, C.; Escobar, R. A.; Cardemil, J. M.; Simsek, Y.; Matute, J. A., Solar polygeneration for electricity production and desalination: Case studies in Venezuela and northern Chile. Renewable Energy 2017, 101, 387-398. Huang, Y.; Wang, Y. D.; Chen, H.; Zhang, X.; Mondol, J.; Shah, N.; Hewitt, N. J., Performance analysis of biofuel fired trigeneration systems with energy storage for remote households. Applied Energy 2017, 186, Part 3, 530-538. Cormos, C.-C., Biomass direct chemical looping for hydrogen and power coproduction: Process configuration, simulation, thermal integration and technoeconomic assessment. Fuel Processing Technology 2015, 137, (Supplement C), 1623. Cormos, C.-C., Renewable hydrogen production concepts from bioethanol reforming with carbon capture. International Journal of Hydrogen Energy 2014, 39, (11), 55975606. Hernández, B.; León, E.; Martín, M., Bio-waste selection and blending for the optimal production of power and fuels via anaerobic digestion. Chemical Engineering Research and Design 2017, 121, (Supplement C), 163-172. Ghaith, F. A.; Abusitta, R., Energy analyses of an integrated solar powered heating and cooling systems in UAE. Energy and Buildings 2014, 70, 117-126. Starfelt, F.; Thorin, E.; Dotzauer, E.; Yan, J., Performance evaluation of adding ethanol production into an existing combined heat and power plant. Bioresource Technology 2010, 101, (2), 613-618. Lythcke-Jørgensen, C.; Clausen, L. R.; Algren, L.; Hansen, A. B.; Münster, M.; Gadsbøll, R. Ø.; Haglind, F., Optimization of a flexible multi-generation system based on wood chip gasification and methanol production. Applied Energy 2017, 192, 337-359. Djuric Ilic, D.; Dotzauer, E.; Trygg, L., District heating and ethanol production through polygeneration in Stockholm. Applied Energy 2012, 91, (1), 214-221.
ACCEPTED MANUSCRIPT 63 of 67
1947 1948 1949 1950 1951 1952 1953 1954 1955 1956 1957 1958 1959 1960 1961 1962 1963 1964 1965 1966 1967 1968 1969 1970 1971 1972 1973 1974 1975 1976 1977 1978 1979 1980 1981 1982 1983 1984 1985 1986 1987 1988 1989
115.
116.
117.
118. 119.
120.
121.
122.
123.
124.
125. 126.
127.
128.
129.
Kumar, S.; Kwon, H.-T.; Choi, K.-H.; Hyun Cho, J.; Lim, W.; Moon, I., Current status and future projections of LNG demand and supplies: A global prospective. Energy Policy 2011, 39, (7), 4097-4104. Gao, P.; Li, W.; Cheng, Y.; Tong, Y.; Dai, Y.; Wang, R., Thermodynamic performance assessment of CCHP system driven by different composition gas. Applied Energy 2014, 136, 599-610. Li, M.; Mu, H.; Li, N.; Ma, B., Optimal design and operation strategy for integrated evaluation of CCHP (combined cooling heating and power) system. Energy 2016, 99, 202-220. Adams, T. A.; Barton, P. I., Combining coal gasification and natural gas reforming for efficient polygeneration. Fuel Processing Technology 2011, 92, (3), 639-655. Zhu, L.; Zhang, Z.; Fan, J.; Jiang, P., Polygeneration of hydrogen and power based on coal gasification integrated with a dual chemical looping process: Thermodynamic investigation. Computers & Chemical Engineering 2016, 84, 302-312. Chiesa, P.; Consonni, S.; Kreutz, T.; Williams, R., Co-production of hydrogen, electricity, and CO2 from coal with commercially ready technology. Part A: Performance and emissions. International Journal of Hydrogen Energy 2005, 30, 747-767. Kreutz, T.; Williams, R.; Consonni, S.; Chiesa, P., Co-production of hydrogen, electricity and CO2 from coal with commercially ready technology. Part B: Economic analysis. International Journal of Hydrogen Energy 2005, 30, (7), 769-784. Tock, L.; Maréchal, F., Thermo-environomic optimisation strategy for fuel decarbonisation process design and analysis. Computers & Chemical Engineering 2015, 83, 110-120. Rosato, A.; Sibilio, S.; Ciampi, G., Dynamic performance assessment of a buildingintegrated cogeneration system for an Italian residential application. Energy and Buildings 2013, 64, 343-358. Rubio-Maya, C.; Uche-Marcuello, J.; Martínez-Gracia, A.; Bayod-Rújula, A. A., Design optimization of a polygeneration plant fuelled by natural gas and renewable energy sources. Applied Energy 2011, 88, (2), 449-457. Rubio-Maya, C.; Uche, J.; Martínez, A., Sequential optimization of a polygeneration plant. Energy Conversion and Management 2011, 52, (8–9), 2861-2869. Wu, J. Y.; Wang, J. L.; Li, S.; Wang, R. Z., Experimental and simulative investigation of a micro-CCHP (micro combined cooling, heating and power) system with thermal management controller. Energy 2014, 68, 444-453. Maidment, G. G.; Zhao, X.; Riffat, S. B.; Prosser, G., Application of combined heatand-power and absorption cooling in a supermarket. Applied Energy 1999, 63, (3), 169-190. Zhang, Y.; Wang, X.; Zhuo, S.; Zhang, Y., Pre-feasibility of building cooling heating and power system with thermal energy storage considering energy supply–demand mismatch. Applied Energy 2016, 167, 125-134. Zabihian, F.; Fung, A., A Review on Modeling of Hybrid Solid Oxide Fuel Cell Systems. International Journal of Engineering 2009, 3, (2), 85--119.
ACCEPTED MANUSCRIPT 64 of 67
1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017 2018 2019 2020 2021 2022 2023 2024 2025 2026 2027 2028 2029 2030 2031 2032
130.
131.
132.
133.
134.
135.
136. 137. 138.
139. 140. 141.
142.
143.
144.
Bentsen, N. S.; Felby, C.; Thorsen, B. J., Agricultural residue production and potentials for energy and materials services. Progress in Energy and Combustion Science 2014, 40, 59-73. Sahoo, U.; Kumar, R.; Pant, P. C.; Chaudhury, R., Scope and sustainability of hybrid solar–biomass power plant with cooling, desalination in polygeneration process in India. Renewable and Sustainable Energy Reviews 2015, 51, 304-316. Angrisani, G.; Bizon, K.; Chirone, R.; Continillo, G.; Fusco, G.; Lombardi, S.; Marra, F. S.; Miccio, F.; Roselli, C.; Sasso, M.; Solimene, R.; Tariello, F.; Urciuolo, M., Development of a new concept solar-biomass cogeneration system. Energy Conversion and Management 2013, 75, 552-560. Pantaleo, A. M.; Camporeale, S. M.; Miliozzi, A.; Russo, V.; Shah, N.; Markides, C. N., Novel hybrid CSP-biomass CHP for flexible generation: Thermo-economic analysis and profitability assessment. Applied Energy 2017. Vidal, M.; Martín, M., Optimal coupling of a biomass based polygeneration system with a concentrated solar power facility for the constant production of electricity over a year. Computers & Chemical Engineering 2015, 72, 273-283. Ondeck, A. D.; Edgar, T. F.; Baldea, M., Optimal operation of a residential districtlevel combined photovoltaic/natural gas power and cooling system. Applied Energy 2015, 156, 593-606. Chicco, G.; Mancarella, P., Distributed multi-generation: A comprehensive view. Renewable and Sustainable Energy Reviews 2009, 13, (3), 535-551. Bizzarri, G.; Morini, G. L., New technologies for an effective energy retrofit of hospitals. Applied Thermal Engineering 2006, 26, (2), 161-169. Kieffer, M.; Brown, T.; Brown, R. C., Flex fuel polygeneration: Integrating renewable natural gas into Fischer–Tropsch synthesis. Applied Energy 2016, 170, 208-218. Angrisani, G.; Roselli, C.; Sasso, M., Distributed microtrigeneration systems. Progress in Energy and Combustion Science 2012, 38, (4), 502-521. Ziher, D.; Poredos, A., Economics of a trigeneration system in a hospital. Applied Thermal Engineering 2006, 26, (7), 680-687. Ghaebi, H.; Amidpour, M.; Karimkashi, S.; Rezayan, O., Energy, exergy and thermoeconomic analysis of a combined cooling, heating and power (CCHP) system with gas turbine prime mover. International Journal of Energy Research 2010, 35, (8), 697-709. Khaliq, A.; Kumar, R., Thermodynamic performance assessment of gas turbine trigeneration system for combined heat cold and power production. Journal of Engineering for Gas Turbines and Power 2008, 130, (2). Bruno, J. C.; Massagués, L.; Coronas, A. In Power quality and air emission tests in a microgas turbine cogeneration plant, Proceedings of the International Conference on Renewable Energy and Power Quality (ICREPQ'03), Vigo, Spain, 2003; Vigo, Spain, 2003. Vélez, F.; Segovia, J. J.; Martín, M. C.; Antolín, G.; Chejne, F.; Quijano, A., A technical, economical and market review of organic Rankine cycles for the
ACCEPTED MANUSCRIPT 65 of 67
2033 2034 2035 2036 2037 2038 2039 2040 2041 2042 2043 2044 2045 2046 2047 2048 2049 2050 2051 2052 2053 2054 2055 2056 2057 2058 2059 2060 2061 2062 2063 2064 2065 2066 2067 2068 2069 2070 2071 2072 2073 2074 2075
145.
146. 147. 148.
149. 150.
151.
152.
153. 154.
155.
156.
157.
158. 159.
conversion of low-grade heat for power generation. Renewable and Sustainable Energy Reviews 2012, 16, (6), 4175-4189. Ji, J.; He, C.; Xiao, Z.; Miao, C.; Luo, L.; Chen, H.; Zhang, X.; Guo, H.; Wang, Y.; Roskilly, T., Simulation Study of an ORC System Driven by the Waste Heat Recovered from a Trigeneration System. Energy Procedia 2017, 105, 5040-5047. Bae, C.; Kim, J., Alternative fuels for internal combustion engines. Proceedings of the Combustion Institute 2017, 36, (3), 3389-3413. Wu, D. W.; Wang, R. Z., Combined cooling, heating and power: A review. Progress in Energy and Combustion Science 2006, 32, (5), 459-495. Hossain, A. K.; Thorpe, R.; Vasudevan, P.; Sen, P. K.; Critoph, R. E.; Davies, P. A., Omnigen: Providing electricity, food preparation, cold storage and pure water using a variety of local fuels. Renewable Energy 2013, 49, 197-202. Taymaz, I., An experimental study of energy balance in low heat rejection diesel engine. Energy 2006, 31, (2), 364-371. Mı́guez, J. L.; Murillo, S.; Porteiro, J.; López, L. M., Feasibility of a new domestic CHP trigeneration with heat pump: I. Design and development. Applied Thermal Engineering 2004, 24, (10), 1409-1419. Porteiro, J.; Mı́guez, J. L.; Murillo, S.; López, L. M., Feasibility of a new domestic CHP trigeneration with heat pump: II. Availability analysis. Applied Thermal Engineering 2004, 24, (10), 1421-1429. Rey, G.; Ulloa, C.; Míguez, J. L.; Arce, E., Development of an ICE-Based MicroCHP System Based on a Stirling Engine; Methodology for a Comparative Study of its Performance and Sensitivity Analysis in Recreational Sailing Boats in Different European Climates. Energies 2016, 9, 239. Sonar, D.; Soni, S. L.; Sharma, D., Micro-trigeneration for energy sustainability: Technologies, tools and trends. Applied Thermal Engineering 2014, 71, (2), 790-796. Onovwiona, H. I.; Ugursal, V. I., Residential cogeneration systems: review of the current technology. Renewable and Sustainable Energy Reviews 2006, 10, (5), 389431. Samavati, M.; Raza, R.; Zhu, B., Design of a 5-kW advanced fuel cell polygeneration system. Wiley Interdisciplinary Reviews: Energy and Environment 2012, 1, (2), 173180. Ranjbar, F.; Chitsaz, A.; Mahmoudi, S. M. S.; Khalilarya, S.; Rosen, M. A., Energy and exergy assessments of a novel trigeneration system based on a solid oxide fuel cell. Energy Conversion and Management 2014, 87, 318-327. Chitsaz, A.; Mehr, A. S.; Mahmoudi, S. M. S., Exergoeconomic analysis of a trigeneration system driven by a solid oxide fuel cell. Energy Conversion and Management 2015, 106, 921-931. Kronenberg, G., Cogeneration with the LT-MED desalination process. Desalination 1997, 208, 287-294. Tzimas, E.; Cormos, C.-C.; Starr, F.; Garcia-Cortes, C., The design of carbon capture IGCC-based plants with hydrogen co-production. Energy Procedia 2009, 1, (1), 591598.
ACCEPTED MANUSCRIPT 66 of 67
2076 2077 2078 2079 2080 2081 2082 2083 2084 2085 2086 2087 2088 2089 2090 2091 2092 2093 2094 2095 2096 2097 2098 2099 2100 2101 2102 2103 2104 2105 2106 2107 2108 2109 2110 2111 2112 2113 2114 2115 2116 2117 2118
160. 161. 162. 163. 164.
165.
166. 167. 168.
169.
170.
171.
172. 173.
174. 175.
176.
Farhat, K.; Reichelstein, S., Economic value of flexible hydrogen-based polygeneration energy systems. Applied Energy 2016, 164, 857-870. Consonni, S.; Viganò, F., Decarbonized hydrogen and electricity from natural gas. International Journal of Hydrogen Energy 2005, 30, (7), 701-718. Adams, T. A.; Barton, P. I., High-efficiency power production from natural gas with carbon capture. Journal of Power Sources 2010, 195, (7), 1971-1983. Adams, T. A.; Barton, P. I., High-efficiency power production from coal with carbon capture. AIChe 2010. Costa, A.; Keane, M. M.; Torrens, J. I.; Corry, E., Building operation and energy performance: Monitoring, analysis and optimisation toolkit. Applied Energy 2013, 101, (Supplement C), 310-316. Marín-Sáez, J.; Chemisana, D.; Moreno, Á.; Riverola, A.; Atencia, J.; Collados, M.V., Energy Simulation of a Holographic PVT Concentrating System for Building Integration Applications. Energies 2016, 9, 577. Syms, R. R. A., Vector Effects in Holographic Optical Elements. Optica Acta: International Journal of Optics 1985, 32, (11), 1413-1425. Shaneb, O. A.; Taylor, P. C.; Coates, G., Optimal online operation of residential μCHP systems using linear programming. Energy and Buildings 2012, 44, 17-25. Kong, X. Q.; Wang, R. Z.; Huang, X. H., Energy optimization model for a CCHP system with available gas turbines. Applied Thermal Engineering 2005, 25, (2), 377391. Espirito Santo, D. B., Energy and exergy efficiency of a building internal combustion engine trigeneration system under two different operational strategies. Energy and Buildings 2012, 53, (Supplement C), 28-38. Ahmadi, P.; Dincer, I.; Rosen, M. A., Thermodynamic modeling and multi-objective evolutionary-based optimization of a new multigeneration energy system. Energy Conversion and Management 2013, 76, 282-300. Kavvadias, K. C.; Tosios, A. P.; Maroulis, Z. B., Design of a combined heating, cooling and power system: Sizing, operation strategy selection and parametric analysis. Energy Conversion and Management 2010, 51, (4), 833-845. Willis, H. L.; Scott, W. G., Distributed Power Generation: Planning and Evaluation. 2000. Barsali, S.; De Marco, A.; Giglioli, R.; Ludovici, G.; Possenti, A., Dynamic modelling of biomass power plant using micro gas turbine. Renewable Energy 2015, 80, 806-818. Alanne, K.; Saari, A., Distributed energy generation and sustainable development. Renewable and Sustainable Energy Reviews 2006, 10, (6), 539-558. Stadler, P.; Ashouri, A.; Maréchal, F., Model-based optimization of distributed and renewable energy systems in buildings. Energy and Buildings 2016, 120, (Supplement C), 103-113. Manfren, M.; Caputo, P.; Costa, G., Paradigm shift in urban energy systems through distributed generation: Methods and models. Applied Energy 2011, 88, (4), 10321048.
ACCEPTED MANUSCRIPT 67 of 67
2119 2120 2121 2122 2123 2124 2125 2126 2127 2128 2129 2130 2131 2132 2133
177.
178. 179.
180.
181.
Sučić, S.; Capuder, T., Automation of flexible distributed multi-generation systems by utilizing optimized middleware platform. Applied Energy 2016, 169, (Supplement C), 542-554. Mancarella, P., MES (multi-energy systems): An overview of concepts and evaluation models. Energy 2014, 65, (Supplement C), 1-17. Akbari, K.; Nasiri, M. M.; Jolai, F.; Ghaderi, S. F., Optimal investment and unit sizing of distributed energy systems under uncertainty: A robust optimization approach. Energy and Buildings 2014, 85, (Supplement C), 275-286. Mehleri, E. D.; Sarimveis, H.; Markatos, N. C.; Papageorgiou, L. G., A mathematical programming approach for optimal design of distributed energy systems at the neighbourhood level. Energy 2012, 44, (1), 96-104. Tian, Z.; Niu, J.; Lu, Y.; He, S.; Tian, X., The improvement of a simulation model for a distributed CCHP system and its influence on optimal operation cost and strategy. Applied Energy 2016, 165, 430-444.
ACCEPTED MANUSCRIPT
Current literature studies on polygeneration systems simulation The review includes conversion technologies fuelled by fossil and/or renewable energy A classification of methodology, outputs, control and optimization Contribution of polygeneration to the development of distribute generation systems