Sie sind auf Seite 1von 12

Applied Energy 192 (2017) 563–574

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Technical and economic assessment of a SOFC-based energy system


for combined cooling, heating and power
Andrea L. Facci a,⇑, Viviana Cigolotti b, Elio Jannelli c, Stefano Ubertini a
a
DEIM – School of Engineering, University of Tuscia, 01100 Viterbo, Italy
b
ENEA – Portici Research Center, 80055 Portici (Na), Italy
c
Department of Engineering, University of Napoli Parthenope, 80143 Napoli, Italy

h i g h l i g h t s

 Assessment of the performances of a SOFC based CHCP plant for residential applications.
 Optimal plant design as a function of investment cost and control strategy.
 Optimized control strategies to minimize the costs and primary energy consumption.

a r t i c l e i n f o a b s t r a c t

Article history: Here we present the technical and economical performances of a small scale trigeneration power plant
Received 18 April 2016 based on solid oxide fuel cells and designed for a small residential cluster (i.e. 10 apartments). The energy
Received in revised form 17 June 2016 system features a natural gas solid oxide fuel cell, a boiler, a refrigerator, and a thermal storage system.
Accepted 18 June 2016
We compare different power plant configurations varying the size of the fuel cell and the refrigeration
Available online 5 July 2016
technology to satisfy the chilling demand (i.e. absorption or mechanical chiller). Given that the ability
to meet the power demand is crucial in this kind of applications, the plant performances are assessed
Keywords:
following an optimal control strategy, as a function of different energy demand profiles and electricity
Trigeneration
Distributed generation
prices, and of rated and part load efficiencies of each energy converter. The optimization of the energy
CHCP system operating strategy is performed through a graph theory-based methodology. Results are provided
Fuel cells in terms of electrical and thermal efficiency, operating strategy, as well as economic saving, primary
SOFC energy consumption reduction, and pay back period, considering different capital costs of the fuel cell.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction reducing the energy consumption of buildings by 50% compared


to 2010 [3]. The EU requires that all new buildings must be ‘‘near
Increasing the energy efficiency and reducing the environmen- ly-zero-energy” by 31 December 2020 [4], and all new public
tal impact of buildings is a key aspect for developed countries buildings must be ‘‘nearly-zero-energy” by 2018 [8] as a determi-
[1–3]. In fact, buildings are responsible for 40% of the energy nant measure to meet the objective of reducing the primary energy
demand and 36% of the carbon dioxide emissions in the European consumption by 20% with respect to the business as usual
Union (EU) [4]. In particular, households cause 25% of the total projection.
greenhouse gases emission related to fossil fuel combustion in Cogeneration, or combined heat and power (CHP), and trigener-
the EU [5]. Similarly, in 2010, buildings used about 41% of the pri- ation or combined heat cooling and power (CHCP) are reliable
mary energy consumption of the United States (US) [6,7], 54% of technologies that are already contributing to the global energy
which only for the residential sector [7]. In this scenario, both US demand [9]. According to the International Energy Agency (IEA),
and EU undertook actions to reduce the buildings energy waste CHP coupled with district heating and cooling could save
[3,4]. The US Department of Energy envisages the possibility of 950 Mton/year of carbon dioxide emissions by 2030 [9]. Although
the main driver for large CHP investments is economic [9],
cogeneration and distributed generation (DG) offer several other
⇑ Corresponding author.
opportunities, such as: (i) increased overall efficiency compared
E-mail addresses: andrea.facci@unitus.it (A.L. Facci), viviana.cigolotti@enea.it
(V. Cigolotti), elio.jannelli@uniparthenope.it (E. Jannelli), stefano.ubertini@unitus.it
to separate production [10,11]; (ii) reduction of pollutant and
(S. Ubertini). green house gases emissions [11,12]; (iii) deferring expensive

http://dx.doi.org/10.1016/j.apenergy.2016.06.105
0306-2619/Ó 2016 Elsevier Ltd. All rights reserved.
564 A.L. Facci et al. / Applied Energy 192 (2017) 563–574

investments on large size plants [11], and on transmission and dis- 2. Methodology
tribution system [12]; (iv) reducing losses in the distribution sys-
tem [12]; (v) providing grid support or ancillary services [12]; We dissect the performances of a SOFC-based CHCP plant in a
(vi) promoting the use of alternative technologies and renewable realistic energy management scenario. In particular, we
sources [11,13,14]. Several case studies related to the perfor- concentrate on the annual energy cost and on the primary energy
mances of CHP systems can be found in literature [15–21], also consumption. We compare different configurations, varying the
coupled with renewable sources [13,15–17,22–24]. equipment size and the refrigerator technology (further details
Solid oxide fuel cells (SOFC) are a promising technology for are given in 3.3).
high-efficiency and sustainable energy conversion [25–27], and The operating efficiency of such a plant is largely determined by
are expected to play an important role in future distributed energy its control strategy as highlighted in [10,37,38,45,46,48]. There-
generation [28]. SOFCs have many advantages over conventional after, we compare the performances of the different configurations
power plants including: (i) high efficiency [25]; (ii) fuel adaptabil- assuming two possible management policies: cost minimization
ity [29]; (iii) very low NOx ; SOx and particulate matter emissions, and PEC minimization. Both control strategies are determined
and reduced CO2 emissions [30]; (iv) vibration-free operation through the optimization methodology introduced in [45], and fur-
[31]. Theoretically, the electrical efficiency of a SOFC may be larger ther developed in [46,47]. Such a methodology minimizes a pre-
than 70% [32]. On the other hand, commercially available SOFC scribed objective function on a daily basis accounting for: (i) the
systems have a net electrical efficiency in the range 40–60% [32]. design performances of all the subsystems; (ii) the derating of
SOFC systems can produce an electrical power from few W up to the performances at part load; (iii) the effects of environmental
several hundreds of kW [25,32] and their efficiency is only slightly conditions; (iv) energy demand and costs as functions of time;
influenced by the system scale [32]. SOFC plants with an electrical (v) maintenance, and cold start costs; (vi) constraints related to
power of 1 kW and an efficiency of 60% have been demonstrated in the dynamic behavior of the equipment, such as the minimum
[32]. The high working temperature facilitates the integration of time interval between two consecutive starts or shutdowns. All
SOFCs within CHCP plants [32]. Therefore SOFCs, are becoming a the energy converters are modeled as black-boxes, through their
promising prime mover for CHCP applications [25–28,33,34]. efficiency curves as functions of the set-point and environmental
After decades of research and development, SOFCs are now conditions.
close to commercialization [32], and there are already several SOFC The prescribed objective function for cost minimization is:
high-efficiency micro-cogenerators conceived primarily for resi-
dential use or small-scale commercial applications. More than X
24
GCost ¼ C f ðh; sðhÞÞ þ C m ðh; sðhÞÞ þ C s ðh; sðhÞÞ  Rðh; sðhÞÞ; ð1Þ
1000 CHP installations of polymer electrolyte fuel cells (PEMFC)
h¼1
and SOFC have been realized starting from the late 2014, mainly
within the European project ENE.FIELD [35]. The ENE-FARM is being h the time interval, C f the cost of fuel, C m the maintenance
the most successful project on micro CHP [35]. It has been realized cost, C s the cold-start cost, and R the revenue/cost yielding from
in Japan from 2008 until the end of 2015, whith more than 150,000 the electricity exchanged with the grid. Costs are functions of the
CHP units (both PEMFC and SOFC) installed in residential areas time interval and the plant state (i.e. the set-point of the
[35]. Typical applications are those with high power and heat subsystems) sðhÞ.
demands, like office buildings, swimming pools, and small and PEC minimization requires the following objective function
medium enterprises [36]. reported in Eq. (2)
The control strategy determines the performances of DG plants
[10,37] as much as the technological level of the components X
24

[38–40]. Therefore, realistic working conditions must be considered GPEC ¼ Ef ðh; sðhÞÞPEFf þ Egrid PEFgrid ; ð2Þ
h¼1
for their economical and environmental evaluation [38,39,41–43].
Determining the set-point is critical for CHP and CHCP, as each of where Ef is the energy content of the fuel, PEFf is the primary
the energy vectors simultaneously produced represents a con- energy factor of the fuel [49], Egrid is the electricity exchanged with
straint for the plant [27,38,44]. Optimized control strategies the grid, and PEFgrid is the primary energy factor of electricity [49].
emphasize the strengths of DG plants, compared to predetermined
Eqs. (1) and (2) are discretized with respect to the plant state
strategies (i.e. thermal tracking or electrical tracking), impacting
and in time, and the problem is represented as a weighted and ori-
the return on investment and the environmental benefit [43].
ented graph. The optimal control strategy is determined seeking
In this paper we evaluate the potential of a SOFC-based CHCP
for the shortest path across the graph.
plant in terms of economic, energy, and environmental perfor-
We combine two energy demand profiles, one typical of sum-
mances. We hypothesize that such a plant satisfies the energy
mer and one representative of winter (see Section 3.1), and two
demand of a small residential cluster (i.e. 10 apartments) and we
electricity cost profiles (see Section 3.2), one for working days
compare different configurations varying the SOFC electrical power
and one for non-working days, to obtain four sample combinations.
and the refrigerator machine technology (i.e. absorption or mechan-
The control strategy is optimized for the four days, and the daily
ical chiller). Two control strategies are considered for each configu-
results are projected on the whole year to compare the perfor-
ration: an economically optimized strategy that minimizes the total
mances of the different plant configurations.
daily cost, and an energy consumption optimized strategy that min-
imizes the daily primary energy consumption (PEC). Both strategies
are determined through the methodology presented in [45–47] that 3. Case study
accounts for hourly energy demand and costs, and for the efficiency
of the energy system components as a function of their set-point. In this paper we evaluate the performances of the CHCP plant
The paper is organized as follows. In Section 2 we describe the described in Section 3.3 to comply the energy demand of a rela-
methodology utilized to determine the total energy cost. The case tively small residential unit.
study is presented in Section 3, with particular reference to the The prime mover is a cogenerative SOFC. A natural gas boiler is
plant configurations, the energy demand, and the energy tariffs. included as a back-up and to satisfy peak thermal energy demand.
Results are dissected in Section 4 and conclusions are drawn in Both absorption and mechanical chillers are considered for the
Section 5. production of chilling energy.
A.L. Facci et al. / Applied Energy 192 (2017) 563–574 565

The CHCP plant is grid connected, but the excess energy fed into Table 1
Electricity prices before VAT expressed in €/kW h for year 2015, as functions of the
the power grid is not remunerated. Thus, the term Rðh; sðhÞÞ in Eq.
consumption quota for single housing unit, the time slot, and the quarter. Slot F1
(1) may only assume negative values. We comment that such an refers to hours between 8 a.m. and 7 p.m. of working days; slot F2 refers to the rest of
assumption facilitates the generalization of the methodology and the week. Data are retrieved from [51].
of the results to different countries. In fact, the remuneration of
Q Slot Consumption quota [kW h/year]
electricity produced by DG plants is subject to stringent national
rules. Moreover, prices are strongly influenced by complex 0–1800 2801–2640 2641–4440 >4440

subsidizing mechanisms that vary on a national basis. Finally, we I F1 0.1279 0.1863 0.2541 0.3006
note that such an assumption is expected to underestimate the F2 0.1215 0.1800 0.2478 0.2942

profitability of SOFC based plants. Thereafter, the design guidelines II F1 0.1257 0.1844 0.2526 0.2990
emerging herein can be considered conservative. F2 0.1820 0.1991 0.2390 0.2816
III F1 0.1243 0.1835 0.2523 0.2988
F2 0.1824 0.1996 0.2394 0.2820
3.1. Energy demand
IV F1 0.1297 0.1895 0.2591 0.3056
F2 0.1894 0.2065 0.2463 0.2890
Fig. 1 reports the energy demand used in the present analysis,
representative of a group of 10 houses. Data relative to the single
housing unit are retrieved from [50]. Significant variations of the Table 1 are intended for singe housing unit. We comment that a
energy demand profiles are not expected at least for industrialized similar mechanism based on the total consumption per year is uti-
countries. In fact, in the residential sector, energy demand is lized on most of the European countries [53]. Hours between 8 a.m.
mainly influenced by climate and lifestyle. Thus, energy demand and 7 p.m of working days fall into the slot F1, the rest of the week
deviations among different countries are expected to compare with is considered into the slot F2. For generality, we averaged over the
regional variations. We consider two different demand profiles, quarter, while retaining the variations with respect to the con-
one representative of a winter day, the other of a summer day. sumption and to the time slot, that may impact the optimal control
Cooling is required only in summer for air conditioning. Thermal strategy. A 10% VAT must be added to the cost reported in Table 1
energy is requested for space heating, only in the winter, and to obtain the final cost, reported in Fig. 2 as a function of the total
domestic hot water, during the whole year. consumption. The increase of the electricity price varying the
During summer, heating load has an average value of 8.6 kW total consumption is significant, in particular, in the range
and a peak of 27 kW. Chilling power request is similar, with an 2–20 MW h/year, and the effective averaged cost of electricity is
average of 7.2 kW, and a peak value of 23 kW. Electricity demand influenced by the plant control strategy.
is slightly higher, being 32 kW and 21 kW the peak and average According to AEEG [52], the natural gas price varies as a func-
power demands respectively. tion of the consumption quota (per single housing unit), and the
During winter, the peak request of thermal energy is 124 kW, quarter. A similar mechanism based on the total consumption
more than 3 times larger compared to electricity demand. Simi- per year is utilized on most of the European countries [54]. The
larly, the average thermal power, whose value is 44 kW, doubles Italian territory is divided into six macro-regions characterized
the electrical one which is just 19 kW. by different prices: North West (NW), North East (NE), Center
(C), Center South West (CSW), Center South East (CSE) and South
3.2. Energy prices (S). For generality we consider an average national and annual cost,
averaging over the quarters and the regions. The dependence on
Prices of electricity and natural gas are retrieved from the the total consumption may impact the optimal control strategy
Italian Authority for Electrical Energy and Gas (AEEG) [51,52], and is retained, as evidenced in Fig. 3. Such a variation is concen-
and reported in Tables 1 and 2. trated in the range 100–500 Sm3/year, while the natural gas cost
The cost of electricity varies as a function of the quarter (Q), of could be considered constant outside this range. 10% VAT also
the time slot, and of the consumption quota. Quotas reported in applies to the natural gas cost.

Fig. 1. User energy demand time series: (a) electricity demand; (b) thermal energy demand; (c) cooling energy demand.
566 A.L. Facci et al. / Applied Energy 192 (2017) 563–574

Table 2
Natural gas prices before VAT, expressed in €/Sm3 for year 2015, as functions of the
consumption quota (for single housing unit), the term, and the region. Data from [51].

Q Reg. Total consumption [Sm3/year]


0–120 121–480 481–1560 1561–5000
NW 0.390 0.506 0.483 0.479
NE 0.385 0.483 0.462 0.458
C 0.390 0.512 0.489 0.485
I
CSW 0.386 0.566 0.538 0.534
CSE 0.388 0.540 0.514 0.510
S 0.383 0.622 0.589 0.586
NW 0.362 0.478 0.456 0.452
NE 0.357 0.456 0.435 0.430
C 0.362 0.484 0.461 0.457
II
CSW 0.359 0.538 0.510 0.506
CSE 0.360 0.512 0.486 0.482
S 0.356 0.595 0.562 0.558
NW 0.356 0.472 0.449 0.445
NE 0.350 0.449 0.428 0.424
C 0.356 0.478 0.455 0.451
III
CSW 0.352 0.532 0.504 0.500
CSE 0.354 0.505 0.480 0.476
S 0.349 0.588 0.555 0.551
Fig. 4. Schematic representation of the CHCP plant analyzed throughout the paper.
NW 0.372 0.488 0.465 0.461
NE 0.366 0.465 0.444 0.440
C 0.372 0.494 0.471 0.467
IV
CSW 0.368 0.548 0.520 0.516
CSE 0.370 0.521 0.496 0.492 Table 3
S 0.365 0.604 0.571 0.568 Configuration of the five energy systems considered for the analysis. Ref. case stands
for the reference scenario.

SOFC Absorption chiller Mechanical chiller Boiler (kW)


Ref. case – – 23 kW 125
Case A 25 kW 23 kW – 125
Case B 25 kW – 23 kW 125
Case C 40 kW 23 kW – 125
Case D 40 kW – 23 kW 125

We consider four scenarios varying the size of the SOFC and the
refrigeration technology, as reported in Table 3. The performance
of the power plant in the four configurations are compared to a
reference scenario that represents the business as usual: electricity
is only acquired from the grid, thermal energy is produced through
the boiler, and cooling energy is obtained from a mechanical refrig-
erator (see Table 3).
Fig. 2. Average electricity cost, after VAT, as a function of the total consumption per
The technical characteristics of the single fuel cell are taken
year. from the commercial unit EnGENTM  2500 module produced by
SolidPOWER [57,58]. Such a system is operated at 700–750 °C
and each module has a nominal electrical output of 2.5 kW, a ther-
mal power of 2 kW and a net electrical efficiency of 50% [59] at
maximum power output. Power can be regulated in the range
30–100%. Electrical and thermal efficiencies are displayed in
Fig. 5(a) as functions of the set-point, and are estimated according
to the data available in [60,61]. Cases A and B utilize 10 modules to
obtain 25 kW of electrical power and 20 kW of thermal power,
while cases C and D consider 16 modules, yielding 40 kW and
32 kW of electricity and heat, respectively.
The nominal power of the boiler is 125 kW for all the configura-
tions (see Table 3), to entirely satisfy the peak thermal demand.
The efficiency curve is represented in Fig. 5(b) and is estimated
according to the model presented in [37].
A 23 kW double effect absorption chiller [62] is utilized to pro-
Fig. 3. Average natural gas cost, after VAT, as a function of the total consumption
duce cooling energy in cases A and C. Given the high temperature
per year.
operation of a SOFC, without loosing generality, we assume here
3.3. Power plant that the fuel cell is able to provide heat at the temperature required
by the chiller [63–65]. We note that this assumption would require
Fig. 4 schematically represents the CHCP plant studied in this an optimization of the thermal flows within the fuel cell energy
paper. It features a cogenerative SOFC, a natural gas fuel boiler, a system with respect to the state of the art [62]. Fig. 5 reports the
refrigeration machine, and a heat storage [55,56]. COP as a function of the thermal load [62].
A.L. Facci et al. / Applied Energy 192 (2017) 563–574 567

Fig. 5. Efficiency of each of the energy converter in the CHCP plant as functions of the set-point: (a) SOFC; (b) boiler; (c) absorption chiller; (d) mechanical chiller.

A 23 kW mechanical chiller is utilized to satisfy the cooling All the considered configurations significantly reduce the yearly
demand for cases B and D. Its design coefficient of performance energy cost with respect to the reference scenario, as evidenced in
is 4.3 (Daikin models RYYQ/RXYQ) [66]. The COP varies as a func- Fig. 6(a). In particular, from 58% to 63% of the energy expenditure
tion of the set-point according to the model in [37], as reported is saved with respect to the reference case. The lowest cost,
in Fig. 5(d). 26 k€/year, pertains to the SOFC-based power plant with electrical
Following the results in [47] we sized the heat storage to accu- chiller (Case D). The SOFC size has a limited impact on the yearly
mulate thermal energy to satisfy the peak demand for three hours. energy cost: cost is reduced by about 3.5–3.7% by increasing the
We here assume a 90% round trip efficiency for the heat storage nominal power from 25 kW to 40 kW. Similar results are obtained
system [67]. switching from the absorption to the mechanical refrigerator.
The plant is grid-connected such that electricity can be acquired Trigeneration also reduces the PEC (see Fig. 6(b)), despite the
from the grid when the energy produced by the SOFC is lower than CHCP plants are controlled with the only objective of minimizing
the one requested by the user. Electricity can also be fed into the energy costs. PEC reduction is in the range 21–27% with respect
grid in case of over-production. Nevertheless, the excess energy to the reference case. Case D yields the minimum PEC. From the
fed into the power grid does not produce any income. comparisons between case A and case C and between case B and
We comment that the proposed configuration maximizes the case D shown in Fig. 6(b), it is evidenced that PEC is reduced by
flexibility of the plant. In fact, the thermal storage and the grid con- about 5% by increasing the nominal power of the SOFC from
nection relax the constraints imposed by the energy demand on 25 kW to 40 kW. Also the utilization of the mechanical chiller
the control strategy. Thermal energy can be produced and con- reduces the PEC by about 5% with respect to the absorption chiller
sumed at different times, mitigating the non-simultaneity of ther- (i.e. Case D vs Case C).
mal and electrical demand profiles. Electricity can be acquired or Electricity exchanged with the grid is drastically reduced to
fed into the power grid, thus releasing the SOFC from electrical below 10% of the amount required by the reference case, as shown
tracking. in Fig. 6(d). The electricity price is significantly reduced, as a con-
The PEF of the electricity exchanged with the grid is set to 2.174 sequence of its dependence on the total consumption exemplified
[46], and the PEF of the natural gas is set to 1.1 [46,49]. in Fig. 2. Electricity unit price is 0:293 €/kWh for slot F2, and
0:297 €/kW h for slot F1, in the reference case. The electricity
acquired from the grid is below 1800 kW h/year per single housing
4. Results and discussion
unit for all the other plant configurations. Thereafter the unit price,
is 0.140 €/kW h for slot F1 and 0.133 €/kW h for slot F2.
4.1. Minimum cost control strategy
Trigeneration increases the natural gas consumption by 54–66%
with respect to separate production (see Fig. 6(c)). However, the
Figs. 6 and 7 compare the global performances obtained
effect of such a variation on the natural gas unit price is negligible.
through the minimum cost control strategy.

Fig. 6. Global performances of the power plant with the minimum cost control strategy: (a) total cost for the energy supply; (b) primary energy consumption; (c) total fuel
consumption; (d) electricity acquired from the grid.
568 A.L. Facci et al. / Applied Energy 192 (2017) 563–574

Fig. 7. Comparison between the different plant configurations in term of economic saving, PEC reduction, and pay back period obtained thorough the minimum cost control
strategy: the area of the circles is proportional to the pay back period and the center of the circles defines relative cost and PEC reduction. The three plots correspond to
different SOFC costs: (a) SOFC cost 2000 €/kW; (b) SOFC cost 10,000 €/kW; (c) SOFC cost 20,000 €/kW.

In fact, the gas acquired by the single housing unit, being in the None of the proposed scenarios appear to be competitive, from
range 3000–5000 Sm3/year, is located in the plateau depicted in an economical perspective, when the SOFC energy system cost is
Fig. 3. equal or above 20,000 €/kW, given that more than 10 years are
The size of the SOFC is expected to significantly impact the required to pay the investment for all the analyzed cases.
system cost. SOFCs are at a pre-commercial stage, thus a precise Notably, configurations utilizing the mechanical refrigerator
estimation of the SOFC capital cost per installed kW is still difficult. (B and D) perform better with respect to those featuring the
According to [68], the cost of a 1 kW SOFC system is between absorption chiller (A and C). This is related to the fact that, if we
12,000 €/kW and 23,000 €/kW, and between 3000 €/kW and compare the energy demand, reported in Fig. 1, and the SOFC per-
5500 €/kW for a 5 kW system. The IEA [69] reports a cost of formance curves, displayed in Fig. 5(a), we could notice that the
12,000–15,000 €/kW for stationary systems with a power in the thermal efficiency of the SOFC is always lower than the electrical
range 200–300 kW, and forecasts a reduction to 1500–1600 €/kW one. On average, the producible thermal energy is about 1/2 of
with increasing production volumes. According to the model in the electrical energy. The energy demand in the summer is equal
[70], the cost of a 50 kW SOFC energy system is in the range to 454 kW h/day of electricity and 205 kW h/day of heat. Thus,
3000–5500 €/kW. In [71] the installed cost of a SOFC plant is for configuration B and D, heat can be completely produced
expressed as a function of the installed power and of the number through cogeneration. On the other hand, utilizing the absorption
of units produced per year and varies between 12,000 €/kW, for a chiller, 172 kW h/day of chilling demand are converted into a fur-
production of 100 systems per year with an average power of ther thermal demand. As a consequence, in configurations A and C
1 kW, and 400 €/kW for a production 50,000 units per year and heat cannot be completely cogenerated.
an average power of 100 kW. According to [72] the cost of a Fig. 8 displays the set-point time series that minimizes the
270 kW system decreases from 3500 €/kW to 670 €/kW by increas- energy cost for the SOFC. In all the configurations and energy
ing the production volume from 50 units/year to 10,000 units/year. demand profiles, the fuel cell is never turned off. For cases A and
Fig. 7, summarizes the performances of the analyzed scenarios B, that utilize the 25 kW SOFC, the prime mover is operated at full
in terms of relative cost and PEC reduction with respect to the load during evening hours (from 6 p.m. to 11 p.m), that are charac-
reference scenario, and of the pay back period (PBP). We hypothe- terized by large electricity and heat demand. During the rest of day
size to update the reference scenario to a CHCP plant. Thereafter, the SOFC power is modulated to match the energy demand profile
the PBP is evaluated considering only the costs of the equipment and to increase the electrical efficiency of the fuel cell. The 40 kW
that are not present in the reference scenario (i.e. the absorption SOFC utilized for cases C and D always operates at part load,
chiller and the SOFC). For the absorption chiller we assumed a cap- because the electricity sold to the grid is not remunerated and
ital cost of 35 k€ in line with the prices of double effect systems the electrical efficiency of the fuel cell increases by reducing the
available in the market. Given the large variability of the capital set-point.
cost for the SOFC, we considered three different scenarios: (i) SOFC The control strategy of the boiler, shown in Fig. 9, is not directly
cost = 2000 €/kW; (ii) SOFC cost = 10,000 €/kW; (iii) SOFC related to the thermal demand. During winter, the boiler is always
cost = 20,000 €/kW. If we assume that the cost of the SOFC energy utilized to complement the cogenerative thermal production, but
system is 2000 €/kW (see Fig. 7(a)), for case C the investment heat demand and production need not to be strictly simultaneous.
is paid during year 2 by utilizing the minimum cost control The thermal storage, whose state of charge (SOC) is reported in
strategy, while PBP ¼ 3 for the other cases. Thereafter, in this Fig. 10, allows deferred thermal production. In particular, in win-
scenario, Case D is characterized by the largest primary energy ter, the SOC is gradually reduced during central hours, and is then
saving and cash flow and thus it should be considered the suddenly recharged during evening, when the SOFC operates at full
optimal solution. load, and the boiler is turned on at high load. We comment that,
When the cost of the SOFC is increased up to 10,000 €/kW, only having hypothesized a cyclic behavior of the CHCP plant, we con-
cases A and B have a PBP lower than 8 years and could still be strained the SOC to be 50% both at the first and last time intervals.
considered feasible investments for the residential sector. On the Only cases A and C utilize the boiler to satisfy the summer thermal
contrary, cases C and D are characterized by PBP ¼ 11 years that demand, confirming that the configurations that utilize the
may be too long for a profitable investment into the housing sector. mechanical refrigerator are better tailored for the energy demand
Case B has a lower PBP (7 years) compared to case A (8 years) and in study. Finally, the thermal storage allows the boiler to operate
higher cash flow and energy saving, and can be considered the almost always at high efficiency (set-point larger than 25%)
optimal solution in this scenario. irrespective of the heat demand.
A.L. Facci et al. / Applied Energy 192 (2017) 563–574 569

Fig. 8. Set-point of the SOFC as a function of time according to the minimum cost control strategy, for all the configurations. Each plot represents a different daily demand
profile: (a) winter working day; (b) winter holiday; (c) summer working day; (c) summer holiday.

Fig. 9. Set-point of the boiler as a function of time according to the minimum cost control strategy, for all the configurations. Each plot represents a different daily demand
profile: (a) winter working day; (b) winter holiday; (c) summer working day; (c) summer holiday.
570 A.L. Facci et al. / Applied Energy 192 (2017) 563–574

Fig. 10. Thermal storage SOC as a function of time according to the minimum cost control strategy, for all the configurations. Each plot represents a different daily demand
profile: (a) winter working day; (b) winter holiday; (c) summer working day; (c) summer holiday.

The variation of the electricity price has a negligible impact on exchanged with the power grid. As a consequence, PEC needs to
the power plant control strategy, as both the SOFC and the boiler be minimized considering the relevant interactions with the
are controlled in a similar way for working and non-working days. encompassing energy system. When electricity is acquired from
In fact, the cost difference between the time slots F1 and F2 is very the grid, the relative PEC is added to the contribution of the fuel
limited. locally consumed. Similarly, when electricity is fed to the power
grid, the corresponding primary energy is subtracted.
4.2. Minimum PEC control strategy Fig. 11 summarizes the results obtained by the different energy
system configurations when operated through the minimum PEC
It is worth to note that the minimization of the daily energy cost control strategy. With this management policy, trigeneration
does not guarantee the maximum energy efficiency [46,47]. reduces the PEC of an amount comprised between 26% and 39%
Therefore, in this section, we study the behavior of the proposed with respect to the reference case. Compared to the minimum cost
CHCP solutions when the energy system is controlled following control strategy, the minimum PEC policy yields a further PEC
the management policy that minimizes the PEC. As evidenced in reduction in a range between 13 and 15 percentage points. PEC
Eq. (2), the PEC is calculated accounting for the electricity is minimized by configuration C, that utilizes the 40 kW SOFC

Fig. 11. Global performances of the power plant with the minimum PEC control strategy: (a) total cost for the energy supply; (b) primary energy consumption; (c) total fuel
consumption; (d) electricity acquired from the grid.
A.L. Facci et al. / Applied Energy 192 (2017) 563–574 571

and the absorption refrigerator. This is largely related to the fact consequence, the most effective configuration utilizes the large
that a larger fuel cell allows to cogenerate substantially all the SOFC, to maximize electricity production, and the absorption chil-
electrical and thermal power, and the absorption chiller utilizes ler, to increase heat recovery during summer.
the waste energy of the prime mover. Fig. 12 compares different plant configurations, accounting for
The energy cost is reduced in all the configurations, also both economical and energy efficiency related aspects. With a cap-
through the minimum PEC control strategy (see Fig. 11(a)). ital cost of the SOFC of 2000 €/kW the PBP is below 5 years for all
Specifically, savings are comprised between 41% and 56% com- the selected cases. In general, the most effective solution could
pared to reference scenario. Nevertheless, PEC minimization dete- be configuration C, that has a reasonable PBP (4 years) and the lar-
riorates the economic performances with respect to cost gest relative PEC reduction (40% with respect to the reference
minimization, as evidenced comparing Fig. 6 with Fig. 11. For con- case). In fact, even if configuration A yields a larger cost reduction,
figuration A such a variation is relatively limited (about 4%). The which is not the primary objective in this case, it produces a signif-
daily cost is moderately incremented by 8% for cases B and C, icantly lower PEC (26% with respect to the reference case). If the
switching from economical to PEC based optimization. Configura- capital cost of the fuel cell is increased to 10,000 €/kW, the PBP
tion D experiences the most important variation of the operating becomes too long for cases C and D, assuming that 10 years is a
cost, with a 36% increment compared to cost minimization. reasonable time horizon for an investment into the housing field.
The fuel consumption of each single housing unit (Fig. 11(c)) Cases A and B have both PBP ¼ 8 years, and are characterized by
varies between 3000 Sm3/year and 6100 Sm3/year. Thereafter, the a similar primary energy consumption. Case A takes advantage
price of natural gas can be considered constant for all the from a larger cost reduction, 56% with respect to the reference
configurations (Fig. 3), including the reference case, and equal to case, and can be considered the optimal control strategy with a
0.533 €/Sm3. We comment that PEC minimization does not SOFC cost of 10,000 €/kW. In the last scenario, the large capital cost
necessarily reduce the local fuel consumption with respect to cost of the SOFC (20,000 €/kW) hinders the investment into the CHCP
optimization, as noted by comparing Fig. 6(c) with Fig. 11(c). plant. In fact, the PBP becomes higher than 10 years for all the
Rather, it reduces the global PEC, considering also the real or configurations.
avoided consumptions of the surrounding electrical system. The SOFC follows the control strategy reported in Fig. 13 to min-
The electricity bought from the grid by each housing unit varies imize the PEC. In case of PEC minimization the optimal control
from 300 kW h/year for case C, to 4800 kW h/year for case B, strategy is not a function of the energy prices. Thereafter, only
significantly impacting the average price of the electrical energy two scenarios are simulated: the winter and the summer energy
(see Fig. 2). The resulting cost of electricity is reported in Table 4. demand. Comparing Fig. 8 to Fig. 13 we note that, the set-point
Notably, the electrical energy acquired from the grid is minimized of the SOFC is generally larger for PEC minimization, compared
by the plant configuration C that also minimizes the PEC. Then, to economical optimization, except for the case D in summer. In
such a configuration takes also advantage from a reduced cost of the latter case, the electrical energy produced by the fuel cell is
electricity, with respect the other cases, as evidenced in Table 4. reduced as a compromise between electricity production and elec-
Configuration C has also the largest local fuel consumption. The trical efficiency, that, for the SOFC, is maximized at minimum load
high electrical efficiency of the SOFC, compared to conventional (see Fig. 5). We comment that, despite the electrical efficiency of
power plants utilized for centralized electricity production (i.e. to the SOFC is higher than that of the grid, the prime mover does
the efficiency of the grid) explains this behavior. PEC optimization not run at constant full load, because the efficiency of the SOFC
maximizes the distributed energy production also when the energy increases by decreasing the power output (see Fig. 5).
demand is low. Excess electricity is fed to the grid thus reducing The control strategy of the boiler is described in Fig. 14. During
the demand for relatively inefficient conventional plants. As a winter, the thermal demand is larger than the heat produced
through cogeneration, and the boiler is utilized to complement
the thermal production of the SOFC. As already evidenced in
Table 4 Section 4.1, the boiler control strategy is not directly related to
Average electricity cost expressed in €/kW h, for the minimum PEC control strategy. the heat demand. In fact, thanks to the thermal storage, thermal
Case A Case B Case C Case D demand and production do not have to be simultaneous. As a con-
sequence, the boiler operates only in the set-point range of higher
F1 0.1558 0.2183 0.1396 0.1628
F2 0.1493 0.2118 0.1331 0.1563 efficiency. During summer, thermal energy is entirely produced by
cogeneration, except for case A, which is characterized by the

Fig. 12. Comparison between the different plant configurations in term of economic saving, PEC reduction, and pay back period obtained thorough the minimum PEC
strategy: the area of the circles is proportional to the pay back period and the center of the circles defines relative cost and PEC reduction. The three plots correspond to
different SOFC costs: (a) SOFC cost 2000 €/kW; (b) SOFC cost 10,000 €/kW; (c) SOFC cost 20,000 €/kW.
572 A.L. Facci et al. / Applied Energy 192 (2017) 563–574

Fig. 13. Set-point of the SOFC as a function of time according to the minimum PEC control strategy, for all the configurations. Each plot represent a different daily demand
profile: (a) winter working day; (b) winter holiday; (c) summer working day; (c) summer holiday.

Fig. 14. Set-point of the boiler as a function of time according to the minimum cost control strategy, for all the configurations. Each plot represent a different daily demand
profile: (a) winter working day; (b) winter holiday; (c) summer working day; (c) summer holiday.

smaller fuel cell (25 kW) and utilizes the heat as an input to the scenario. PEC minimization yields a cost reduction in the range
absorption refrigerator to satisfy the chilling demand. In this case, 41–56% and a decrease of the primary energy consumption
as already shown in Section 4.1, the thermal power of the SOFC is comprised between 26% and 39% with respect to the reference
not sufficient to cover the whole heat demand, that includes both scenario.
domestic hot water as well as the input energy for the refrigerator. The capital cost of the SOFC is still an open issue, and signifi-
cantly impacts the optimal configuration of the CHCP plant. The
5. Conclusion sensitivity analysis performed herein suggests that investments
on SOFC based DG plants are still feasible for capital costs of the
This paper presents the technical and economical evaluation of SOFC as high as 10 k€/kW. At least one configuration with PBP
a small scale trigeneration power plant based on SOFCs for a resi- lower than 10 years is present in such a scenario, for both the con-
dential application. We compare different configurations varying trol strategies. A further increase of the initial cost hampers the
the SOFC power and the refrigerator machine technology. Two investments. PBP is always longer than 10 years for a capital cost
control strategies are considered for each configuration: an of 20 k€/kW for the SOFC. Conversely, when such a cost is reduced
economically optimized strategy that minimizes the total daily to 2000 €/kW or below, as envisaged in [69,72], all the combina-
cost, and an energy consumption optimized strategy that mini- tions of plant design and control strategy yield PBP below 5 year.
mizes the daily primary energy consumption. In this scenario the economical potential of SOFC based DG plants
Results presented in Section 4, highlight that the chosen objec- is huge.
tive function determines both the control strategy and the optimal Finally, we comment that PEC minimization generates positive
system design. Thereafter, the principal purpose of the investment economical results, and guarantees further 15 percentage points
must be established a priori to correctly configure the CHCP. reduction of the PEC with respect to economical minimization.
Notably, cost minimization fosters the production of chilling Thereafter, designing and controlling the CHCP plant following an
energy through the mechanical refrigerator. Such configurations energy efficiency perspective rather than a pure economical objec-
are not, in general, the most efficient. Rather, these are the better tive, may be reasonable.
tailored on the analyzed energy demand and cost profiles, from The present analysis is performed with reference to the Italian
an economical perspective. On the other hand, PEC is minimized energy market. The general framework, in terms of regulatory
utilizing the absorption chiller, in order to maximize the total effi- aspects and of energy prices, is similar also for other European
ciency of the power plant. countries. Thereafter, the general results of this paper can be
All the considered configurations reduce both the PEC and the extended also to different energy markets. On the other hand, dif-
energy cost with respect to separate production. Specifically, fol- ferences on the energy prices are expected to impact the specific
lowing the minimum cost control strategy, 58–63% of the energy numerical results presented herein. Such a sensitivity analysis will
cost, and 21–27% of the PEC are saved with respect to the reference be performed in future works.
A.L. Facci et al. / Applied Energy 192 (2017) 563–574 573

Acknowledgments [28] Kazempoor P, Dorer V, Ommi F. Evaluation of hydrogen and methane-fuelled


solid oxide fuel cell systems for residential applications: system design
alternative and parameter study. Int J Hydrogen Energy 2009;34(20):8630–44.
This research is supported by the Ministry of Education, Univer- [29] Singhal SC, Kendall K. High-temperature solid oxide fuel cells: fundamentals,
sities and Research Grant PON_03_PE_00109_1 ‘‘Fuel Cell Lab”. The design and applications. Elsevier; 2003.
[30] Baratto F, Diwekar UM, Manca D. Impacts assessment and tradeoffs of fuel cell
authors would also like to express their gratitude to Eng. Lorenzo
based auxiliary power units: Part ii. Environmental and health impacts, LCA,
Lipardi for his profitable support. and multi-objective optimization. J Power Sources 2005;139(1):214–22.
[31] Singhal S. Advances in solid oxide fuel cell technology. Solid State Ion
2000;135(1):305–13.
[32] McPhail SJ, Leto L, Boigues-Muñoz C. The yellow pages of SOFC technology
References international status of SOFC deployment 2012–2013. ENEA; 2013.
[33] Park SK, Kim TS, Sohn JL, Lee YD. An integrated power generation system
[1] The European Parliament and the Council of the European Union. Directive combining solid oxide fuel cell and oxy-fuel combustion for high performance
2012/27/eu of the european parliament and of the council of 25 october 2012 and CO2 capture. Appl Energy 2011;88(4):1187–96.
on energy efficiency, amending directives 2009/125/ec and 2010/30/eu and [34] Papurello D, Borchiellini R, Bareschino P, Chiodo V, Freni S, Lanzini A, et al.
repealing directives 2004/8/ec and 2006/32/ec. Official Journal of the European Performance of a solid oxide fuel cell short-stack with biogas feeding. Appl
Union. Energy 2014;125:254–63.
[2] Concerned Action Energy Performance of Buildings. Homepage; 2014. <http:// [35] Annex 33. Fuel cells for stationary applications. Tech rep. International Energy
www.epbd-ca.eu/> [accessed March 11, 2016]. Agency; 2014. <http://www.ieafuelcell.com/annexdescriptions.php>.
[3] Building Technologies Office. Buildings; 2016. <http://energy.gov/ [36] Enefield. Fuel cells x combined heat and power; 2016. <http://enefield.eu/>
eere/efficiency/buildings> [accessed March 11, 2016]. [accessed March 11, 2016].
[4] European Commission. Buildings; 2016. <https://ec.europa.eu/energy/ [37] Fabrizio E, Filippi M, Virgone J. An hourly modelling framework for the
en/topics/energy-efficiency/buildings> [accessed March 11, 2016]. assessment of energy sources exploitation and energy converters selection and
[5] European Environmental Agency. Households and industry responsible for sizing in buildings. Energy Build 2009;41(10):1037–50.
half of eu greenhouse gas emissions from fossil fuels; 2012. <http://www.eea. [38] Andreassi L, Ciminelli M, Feola M, Ubertini S. Innovative method for energy
europa.eu/highlights/households-and-industry-responsible-for?&utm_campaign= management: modelling and optimal operation of energy systems. Energy
households-and-industry-responsible-for&utm_medium=email&utm_source= Build 2009;41(4):436–44.
EEASubscriptions> [accessed March 11, 2016]. [39] Andreassi L, Ubertini S. Optimal management of power systems. In: Energy
[6] Afram A, Janabi-Sharifi F. Theory and applications of HVAC control systems – a management. Francisco Maria Perez; 2010. intech ISBN: 978-953-307-065-0..
review of model predictive control (MPC). Build Environ 2014;72:343–55. [40] Wakui T, Kawayoshi H, Yokoyama R. Optimal structural design of residential
[7] US Department of Energy. Buildings energy data book; 2012. <http:// power and heat supply devices in consideration of operational and capital
buildingsdatabook.eren.doe.gov/ChapterIntro2.aspx> [accessed March 11, recovery constraints. Appl Energy 2016;163:118–33.
2016]. [41] Doering R, Lin B. Optimum operation of a total energy plant. Comput Oper Res
[8] Europen-Commission. Nearly zero-energy buildings; 2016. <https://ec.europa. 1979;6(1):33–8.
eu/energy/en/topics/energy-efficiency/buildings/nearly-zero-energy- [42] Cardona E, Piacentino A. Optimal design of CHCP plants in the civil sector by
buildings> [accessed March 11, 2016]. thermoeconomics. Appl Energy 2007;84(7–8):729–48.
[9] Agency IE. Combined heat and power evaluating the benefits of greater global [43] Cappa F, Facci AL, Ubertini S. Proton exchange membrane fuel cell for
investment; 2008. cooperating households: a convenient combined heat and power solution for
[10] Onovwiona H, Ugursal V. Residential cogeneration systems: review of the residential applications. Energy 2015;90:1229–38.
current technology. Renew Sust Energy Rev 2006;10(5):389–431. [44] Kong X, Wang R, Li Y, Huang X. Optimal operation of a micro-combined
[11] Chicco G, Mancarella P. Distributed multi-generation: a comprehensive view. cooling, heating and power system driven by a gas engine. Energy Convers
Renew Sust Energy Rev 2009;13(3):535–51. Manage 2009;50(3):530–8.
[12] International Energy Agengy. Distributed generation in liberalised electricity [45] Facci AL, Andreassi L, Ubertini S. Optimization of CHCP (combined heat power
markets; 2002. and cooling) systems operation strategy using dynamic programming. Energy
[13] Lau K, Yousof M, Arshad S, Anwari M, Yatim A. Performance analysis of hybrid 2014;66:387–400.
photovoltaic/diesel energy system under malaysian conditions. Energy [46] Facci AL, Andreassi L, Martini F, Ubertini S. Comparing energy and cost
2010;35(8):3245–55. optimization in distributed energy systems management. J Energy Resour
[14] Franco A, Salza P. Strategies for optimal penetration of intermittent Technol 2014;136(3):032001.
renewables in complex energy systems based on techno-operational [47] Facci AL, Andreassi L, Ubertini S, Sciubba E. Analysis of the influence of thermal
objectives. Renew Energy 2011;36(2):743–53. energy storage on the optimal management of a trigeneration plant. Energy
[15] Cardona E, Piacentino A. A measurement methodology for monitoring a CHCP Proc 2014;45:1295–304.
pilot plant for an office building. Energy Build 2003;35(9):919–25. [48] Hawkes A, Leach M. Cost-effective operating strategy for residential micro-
[16] Bizzarri G, Morini GL. Greenhouse gas reduction and primary energy savings combined heat and power. Energy 2007;32(5):711–23.
via adoption of a fuel cell hybrid plant in a hospital. Appl Therm Eng 2004;24 [49] Schicktanz M, Wapler J, Henning H-M. Primary energy and economic
(23):383–400. analysis of combined heating, cooling and power systems. Energy 2011;36
[17] Merše S, Petelin Visočnik B, Riddoch F, Craenen S, Gardiner P, Rotheray T, et al. (1):575–85.
Cogeneration case studies handbook. Jožef Stefan Institute Energy Efficiency [50] Ren H, Gao W. Economic and environmental evaluation of micro CHP systems
Centre; 2011. with different operating modes for residential buildings in japan. Energy Build
[18] Jannelli E, Minutillo M, Cozzolino R, Falcucci G. Thermodynamic performance 2010;42(6):853–61.
assessment of a small size CCHP (combined cooling heating and power) [51] AEEG. Condizioni economiche per i clienti del mercato tutelato, energia
system with numerical models. Energy 2014;65:240–9. elettrica – servizio di maggior tutela, anno 2015, clienti domestici e non
[19] Hawkes AD, Brett DJL, Brandon N. Fuel cell micro-CHP techno-economics: Part domestici; 2015. <http://www.autorita.energia.it/it/dati/condec.htm]>
1 – Model concept and formulation. Int J Hydrogen Energy 2009;34 [accessed March 11, 2016].
(23):9545–57. [52] AEEG. Condizioni economiche per i clienti del mercato tutelato, gas – servizio
[20] Hawkes AD, Brett DJL, Brandon NP. Fuel cell micro-CHP techno-economics: di tutela, anno 2015, clienti domestici e condomini con uso domestico; 2015.
Part 2 – Model application to consider the economic and environmental <http://www.autorita.energia.it/it/dati/condec.htm]> [accessed March 11,
impact of stack degradation. Int J Hydrogen Energy 2009;34(23):9558–69. 2016].
[21] Fubara TC, Cecelja F, Yang A. Modelling and selection of micro-CHP systems for [53] Eurostat. Electricity prices – price systems 2014; 2015.
domestic energy supply: the dimension of network-wide primary energy [54] Eurostat. Gas prices – price systems 2014; 2015.
consumption. Appl Energy 2014;114:327–34. [55] Haeseldonckx D, Peeters L, Helsen L, Dâhaeseleer W. The impact of thermal
[22] Valente LCG, de Almeida SCA. Economic analysis of a diesel/photovoltaic storage on the operational behaviour of residential CHP facilities and the
hybrid system for decentralized power generation in northern brazil. Energy overall CO2 emissions. Renew Sust Energy Rev 2007;11(6):1227–43.
1998;23(4):317–23. [56] Barbieri ES, Melino F, Morini M. Influence of the thermal energy storage on the
[23] McGowan J, Manwell J. Hybrid wind/PV/diesel system experiences. Renew profitability of micro-CHP systems for residential building applications. Appl
Energy 1999;16(14):928–33. Renewable energy energy efficiency, policy and Energy 2012;97:714–22.
the environment. [57] SolidPOWER. EngenTM2500; 2016. <http://www.solidpower.com/en/engen-
[24] Cozzolino R, Tribioli L, Bella G. Power management of a hybrid renewable 2500/]engen2500> [accessed March 11, 2016].
system for artificial islands: A case study. Energy 2016;106:774–89. [58] Bucheli O, Bertoldi M, Modena S, Ravagni A. Development and manufacturing
[25] Sammes N, Boersma R. Small-scale fuel cells for residential applications. J of SOFC-based products at SOFCpower SpA. ECS Trans 2013;57(1):81–8.
Power Sources 2000;86(12):98–110. [59] SolidPOWER. EnGENTM2500 technical specifications; 2016. <http://
[26] Farhad S, Hamdullahpur F, Yoo Y. Performance evaluation of different www.solidpower.com/wp-content/uploads/2014/03/Data_Sheet_Engen2500_
configurations of biogas-fuelled {SOFC} micro-CHP systems for residential eng.pdf>.
applications. Int J Hydrogen Energy 2010;35(8):3758–68. [60] Wuillemin Z, Ceschini S, Antonetti Y, Beetschen C, Modena S, Montinaro D,
[27] Chiappini D, Facci AL, Tribioli L, Ubertini S. SOFC management in distributed et al. High-performance SOFC stacks tested under different reformate
energy systems. J Fuel Cell Sci Technol 2011;8(3). compositions. In: Proc 11th Eur SOFC SOE Forum.
574 A.L. Facci et al. / Applied Energy 192 (2017) 563–574

[61] Modena S, Ceschini S, Contino AR, Bini R, Tognana L, Bertoldi M, et al. Evolution [67] IEA-IRENA. Thermal energy storage technology brief; 2013. <www.irena.org/
of sofcpowerâstack performances. ECS Trans 2013;57(1):359–66. Pubblications>.
[62] Systema s.p.a. SYBCT assorbitori monoblocco per esterno e interno; 2016. [68] Manufacturing cost analysis of 1 kw and 5 kw solid oxide fuel cell (SOFC) for
<http://www.systema.it/scheda-sybct> [accessed March 11, 2016]. auxilliary power applications. Tech rep. DE-EE0005250, Battelle; 2014.
[63] Burer M, Tanaka K, Favrat D, Yamada K. Multi-criteria optimization of a district [69] International Energy Agency. Fuel cells. In: IEA energy technology essentials;
cogeneration plant integrating a solid oxide fuel cell–gas turbine combined 2007.
cycle, heat pumps and chillers. Energy 2003;28(6):497–518. [70] Palazzi F, Autissier N, Marechal FM, Favrat D. A methodology for thermo-
[64] Yu Z, Han J, Cao X, Chen W, Zhang B. Analysis of total energy system based on economic modeling and optimization of solid oxide fuel cell systems. Appl
solid oxide fuel cell for combined cooling and power applications. Int J Therm Eng 2007;27(16):2703–12.
Hydrogen Energy 2010;35(7):2703–7. [71] James BD, Spisak AB, Colella WG. Manufacturing cost analysis of stationary
[65] Al-Sulaiman FA, Dincer I, Hamdullahpur F. Energy analysis of a trigeneration fuel cell systems. Tech rep. Strategic Analysis; 2012.
plant based on solid oxide fuel cell and organic rankine cycle. Int J Hydrogen [72] Weimar MR, Chick LA, G DW, Whyatt GA. Cost study for manufacturing of solid
Energy 2010;35(10):5104–13. oxide fuel cell power systems. Tech rep. PNNL-22732. US Department of
[66] Daikin. VRV Catalogo Prodotti; 2015. <http://www.daikin.it/docs/Catalogo% Energy.
20VRV%202015-tcm478-349195.pdf> [accessed March 11, 2016].

Das könnte Ihnen auch gefallen