You are on page 1of 10

Proceedings of ASME Turbo Expo 2012

GT2012
June 11-15, 2012, Copenhagen, Denmark

DRAFT GT2012-69579

TECHNO-ECONOMIC ANALYSIS OF INTEGRATED GASIFICATION FUEL


CELL POWER PLANTS CAPTURING CO2

Andrea Lanzini Thomas G. Kreutz


Energetics Department, Politecnico di Torino Princeton Environmental Institute, Princeton University,
Torino, 10129 ITALY Princeton, NJ, 08544 USA

Emanuele Martelli
Energy Department, Politecnico di Milano,
Milano, 20156, ITALY

ABSTRACT For NG-based systems without CO2 capture, as far back as a


This work analyzes the efficiency and economic decade ago both Singhal [1] and Hassmann [2] mentioned ideal
performance of different configurations of a coal-fed Integrated conversion LHV efficiencies as high as 70% using a Siemens-
Gasification Fuel Cell (IGFC) plant with CO2 capture. Our Westinghouse hybrid cycle for plants larger than 2-3 MW;
analysis evaluates novel configurations, providing a detailed Haussmann refers to a pressurized system with a more
economic assessment for each case. ‘sophisticated’ turbine cycle, replacing the micro-turbine, to
The plants studied here are based on a pressurized Solid achieve such high efficiency. More recently, a 2011 DOE-
Oxide Fuel Cell (SOFC) based power cycle integrated with a NETL report [4] on NG SOFC plants capturing CO2 calculated
Shell coal gasifier. The design variations focus on syngas a LHV efficiency of 70.3% for a pressurized plant with external
cleaning and pre-processing upstream of the SOFC power reforming of methane. According to a NETL review of
island. In particular, we have designed, simulated and conventional fossil energy power plants [3], the penalty due to
optimized three main system configurations; two with a partial CCS is almost 8% (LHV basis) in NG-fed combined cycles and
methanation process upstream of the SOFC (‘TREMP’ and ~8-11% for IGCC plants (depending on the specific gasifier
‘HICOM’ cases, respectively) and one without (‘DIRECT’ employed). For NG pressurized hybrid SOFC cycles, Franzoni
case). Depending on the specific plant layout, carbon capture is et al. [4] calculated an efficiency loss less than 3% using post-
accomplished either before or after the SOFC power island, or SOFC CO2 capture. Similarly, Park et al. [5] calculated a LHV
both. efficiency of 67.4% for a NG hybrid SOFC plant venting CO2,
The best performance, both thermodynamic and economic, which reduced to a value of 65.0% when CO2 was captured
was achieved by the ‘HICOM’ case, whose coal-to-electricity after the SOFC reactors. With post-SOFC capture, the anode
conversion efficiency is 50.1% (lower heating value basis). In exhaust is oxy-combusted yielding a stream containing only
addition to outperforming the other IGFC configurations steam and CO2, from which the latter is separated after steam
analysed, compared to a conventional IGCC-CCS plant, the condensation and water knock-out. Concerning coal IGFC
‘HICOM’ case produced almost 30% reduction in the levelized- plants, Grol et al. [6] calculated the efficiency of a pressurized
cost-of-electricity (LCOE) delivered by the power plant. SOFC integrated with a low temperature catalytic gasifier
(producing a syngas with a relatively high methane content, ~
INTRODUCTION 18%) with/without post-SOFC CCS, obtaining LHV
SOFC-based power cycles have the potential to efficiencies of 62.4% and 59.2% in the vented and CCS cases,
significantly increase the fuel-to-electricity conversion rates of respectively. In another recent report, DOE-NETL [7]
new generation power plants; furthermore, when CCS is calculated a LHV efficiency of 51.6% for a pressurized IGFC-
employed, the penalty efficiency paid is significantly reduced CCS plant that employs an enhanced gasifier able to produce a
compared to conventional power cycles. syngas with up to 11% vol. of methane.

1 Copyright © 2012 by ASME


Spallina et al. [8] designed and analyzed several IGFC plant configuration, an LHV efficiency of 51.7% was achieved.
configurations (all based on a Shell gasifier and a pressurized Rao [11] was the first to propose using an (exothermic)
SOFC), essentially leading to different gas turbine inlet methanator to convert part of the chemical energy of a synthesis
temperatures (TITs); LHV efficiencies in the range of 52-54% gas in heat and then to recover power from it efficiently
were calculated. In [9], the authors added post-SOFC CO2 through an expander. Recently, Rao et al. [12] investigated the
capture via oxy-combustion of the anode exhaust and role of a methanator and syngas expander topping cycle placed
calculated a 6% drop in efficiency. Romano et al. [10] studied a upstream an IGFC plant featuring a low-temperature catalytic
novel IGFC plant configuration that captures CO2 before the gasifier (as also employed in [6]); however, no significant
SOFC through physical absorption in a MDEA-based AGR improvement in efficiency was observed because the methane
unit; the plant design also features a methanation process content of the syngas coming from the catalytic hydro-gasifier
upstream of the SOFC (to increase the methane content of the was already so high (36% vol.) to obscure the potential benefit
syngas to ~26% vol.) and a complete recycle of the shifted of methanation upstream of the SOFC.
anode exhaust to the AGR unit for CO2 capture. With this

Fig. 1. Schematic of the three IGFC plants.

2 Copyright © 2012 by ASME


MOTIVATION AND SCOPE bar is included in the performance calculations for all cases.)
This work evaluates the feasibility of and impact of An uncooled gas turbine is used to both compress and heat air
employing syngas methanation upstream of the SOFC in an to feed the SOFC cathode, and to expand the hot cathode
IGFC based on commercial, high temperature, entrained flow exhaust. Key modeling assumptions are summarized in Table
(Shell) gasification. Three IGFC plant configurations, all with A1. Figure 1 illustrates the three alternative gas cleaning/pre-
CO2 capture, were designed and optimized in this study. In the processing routes and the gasifier island that is common to all
reference case (DIRECT), sweet syngas is fed directly to the three cases. A detailed description of the Shell gasifier model is
SOCF, while in the other two cases, the syngas is first partially found in [20].
methanated using the commercially available processes:
TREMP [13] and HICOM [14].
In SOFC stacks with direct internal reforming, which has
become an established option for planar Ni-based anode-
supported stacks [15,16], the presence of methane in the fuel
gas provides a notable boost in conversion efficiency. Not only
does the endothermic reforming process effectively convert
SOFC waste heat directly into chemical enthalpy, but the
reduced cooling duty lowers the (substantial) auxiliary power
required to compress cathode air for SOCF cooling. (Note that Fig. 2. Methanation unit for the ‘TREMP’ and ‘HICOM’ cases.
cathode air flow is the primary mean of removing heat
generated within the stack from the heat of reaction and Syngas methanation upstream the SOFC
overvoltage irreversibilities.) These considerations motivate In the ‘TREMP’ and ‘HICOM’ cases, the syngas feeding the
adding a methanation unit upstream of the SOFC when the fuel SOFC was partially methanated in the methanation unit shown
gas (e.g. syngas from a Shell gasifier) contains little methane. in Fig. 2. The input syngas is mixed with a cooled recycle
The highly exothermic methanation process also provides a stream to create a dilute feed stream at 250 °C that enters an
convenient method of high temperature fuel gas preheating, adiabatic catalytic methanation (chemical equilibrium) reactor.
potentially obviating costly heat exchangers and increasing the The recycle fraction also prevents to reach the temperature limit
efficiency of expanding the syngas in a turbine prior to feeding of 700 °C, at which temperature catalyst sintering would occur
it to the SOFC. Thus, while it may seem counterproductive to [21,22].
methanate the syngas upstream of the SOFC only to The use of only a single methanation reactor limits
immediately reverse that reaction within the SOFC, the somewhat the production of methane. In SNG plants using the
potential advantages are significant. This analysis seeks to full HICOM and TREMP processes, more elaborate
quantify the thermodynamic and economic ramifications of this configurations (with several reactors, inter-stage coolers and
“upstream methanation” strategy. recycle loops) are employed to maximize methane conversion.
However, a single reactor (which significantly increases the
METHODOLOGY methane content of the syngas) is consonant with our goal of
Aspen Plus chemical process design software is used to identifying cost-effective ways to improve SOFC performance
calculate the performance of (and mass and energy balances when using commercial, high temperature, entrained flow
around) all plant components. SOFC performance is simulated gasifiers. As discussed below, the methane conversion is
using a lumped 0-D model, and the cathode air requirement is almost complete in the TREMP case, but somewhat less in the
determined by the energy balance of the ‘hot-box’ [17,18]. An HICOM configuration where a second reactor might yield
integrated waste heat recovery steam cycle is designed and further benefits.
rigorously optimized for each case using the novel method of
Martelli [19]. SOFC Power Island
The SOFC power island (Fig. 1) is common to all three
PLANT CONFIGURATIONS cases, includes the SOFC modules, high-temperature heat-
The three IGFC plants studied here employ a commercial, exchangers for pre-heating the feed streams and cooling the
pressurized, O2-blown, entrained flow, dry-feed (Shell) coal exhaust flows, the anode oxy-combustor and “SOFC
gasifier whose processed syngas feeds an 800 °C pressurized turbomachinery”: a gas compressor for cathode air, expanders
SOFC. Post-SOFC CO2 capture is achieved by combusting the for the oxy-combusted anode exhaust and cathode exhaust,
anode exhaust in oxygen to create a stream of hot, pressurized respectively, and finally hot recirculators. The power core
CO2+H2O which is expanded and cooled to condense out the consists of SOFC modules pressurized to ~20 bar, matching the
water. The remaining (93% vol. pure) CO2 is dried and operating pressure of large industrial turbomachinary. Since
compressed in preparation for pipeline transport, and geologic the gasifier operates at ~38 bar, the syngas pressure is reduced
storage. (Note that the power required to compress CO2 to 150

3 Copyright © 2012 by ASME


(by throttling in the ‘DIRECT’ case and expanding in Table 1 Performance results of the IGFC configurations studied
the ‘TREMP’ and ‘HICOM’ cases) upstream of the
SOFC.1 Both the anode and cathode exhausts are Plant design     DIRECT TREMP HICOM
partially recirculated using hot recirculators [24], Fuel input    
whose electric consumption is very small due to the Coal input, LHV     1827  1827  1827 
very low pressure drop through the SOFC (on both SOFC power unit    
sides). Global fuel utilization, FU     0.85  0.85  0.85 
The air compressor provides a pressurized cathode Local fuel utilization, FU*     0.70  0.80  0.85 
stream at the SOFC; the cathode exhaust is then Global air utilization, λ     4.49  1.73  2.36 
expanded in an uncooled turbine. The anode exhaust
Local air utilization, λ*     8.25  2.60  3.91 
is oxy-combusted, cooled to 850°C before expanding
Operating voltage   V  0.782  0.790  0.770 
to nearly atmospheric pressure in an uncooled gas
SOFC power   MWe  732.8  737.7  723.2 
turbine to generate additional power. Expansion
lowers further the temperature of the oxy-combusted Additional power / cathode side    
anode exhaust prior to final cooling/condensation to TIT   °C  800  800  800 
separate CO2 from H2O. (Cathode exhaust) expander power   MWe  783.2  277.7  394.8 
The SOFC is operated at 800°C; its operating Air compressor consumption  MWe  787.1  303.0  415.3 
voltage is calculated as following: GT net power   MWe  ‐3.9  ‐25.3  ‐20.5 
Additional power / anode side             
, , · . (1)
TIT   °C  850  850  850 
The Nernst potential is given by the reversible Anode steam‐CO2 expander   MWe  112.7  100.9  176.0 
thermodynamic potential generated by the O2 compressor   MWe  5.4  5.3  5.3 
(electrochemical) oxidation of H2 (and CO) as a Oxidant to the oxy‐combustor (ASU)  MWe  13.4  13.1  13.3 
function of temperature, pressure and gas
Power recovery from HRSC    
composition. The area specific resistance (ASR) of the
Additional power from net available steam   MWe  224.1  264.8  228.9 
fuel cell is taken to be 0.28 Ω cm2 [25] with an
Auxiliaries    
operating current density j of 0.64 A cm-2 and a fuel
utilization of 85%. Both anode and cathode Coal prep. & handling + slag handling  MWe  4.6  4.6  4.6 
recirculation are employed to provide a steam-to- Oxidant to the gasifier (ASU)  MWe  63.8  63.8  63.8 
carbon ratio of 2 at the inlet anode chamber and heat Syngas recycle compressor  MWe  0.1  0.1  0.1 
the air up to 650 °C at the cathode inlet, respectively2. AGR (H2S removal only for both  'DIRECT' and 'HICOM' 
The anode stream is pre-heated at 700 °C prior MWe  10.1  87.5  10.1 
cases, H2S and CO2 co‐removal for the 'TREMP' case) 
feeding the SOFC. As mentioned before, to avoid
significant reductions in Nernst voltage due to oxygen Methanator recirculator  MWe  0.0  1.1  0.5 
starvation, the cathode air flow is maintained so that CO2 compression (CO2 removed only post‐SOFC)  MWe  73.7  33.3  73.7 
the partial pressure of O2 in the cathode exhaust Syngas expander  MWe  0.0  6.9  17.2 
exceeds 10%. Summary    
Given the parameters above, for the IGFC cases Net DC power output    MWe  894.6  876.1  953.5 
considered the SOFC resulting power density is ~500
Inverter efficiency     0.96  0.96  0.96 
mW cm-2 (which is also the reference value for the
AC electrical efficiency, ηAC (LHV Coal)  47.0%  46.0%  50.1% 
pressurized SOFC modeled in the NETL reports
[7,26] providing cost analysis).
Heat Recovery Steam Cycle (HRSC)
1
In all three IGFC cases studied, a triple pressure level
Determining the optimal SOFC pressure level for each case was not + reheat Rankine cycle recovers heat from the anode and
attempted here. As shown in [23], the Nernst voltage increases sharply cathode turbine exhausts, as well as from the plant-wide
with pressures up to 5-7 bar; further pressurization is not necessarily
beneficial (generally a quite flat trend of the plant efficiency is
network of heat exchangers, including those in the gasifier
obtained above 10 bar). island and fuel cleaning/processing sections. The heat
2 integration between hot and cold process streams was
When operating an SOFC with exhausts’ recirculation, it is
convenient to express the reactants consumption (or their excess ratio, optimized separately for each plant configuration by following
which is essentially the inverse number) in both global and local the pinch analysis criteria [27]. The HRSC design was then
terms; so for example the global FU is the one actually seen by the optimized using the methodology developed by Martelli et al.
SOFC as a whole (i.e. with a control volume which include the [19,28], which allows to exploit the process excess heat with
recirculation loop), while the local FU refers to a control volume the maximum efficiency by taking into account all the possible
immediately around the SOFC module.

4 Copyright © 2012 by ASME


integrations between the process heat sources/users and the two catalyst), but can withstand - and be able to catalytically
heat recovery steam generators (HRSGs) at the outlet of the convert into methane - a syngas having high CO content
cathode and anode turbines, respectively. The HRSC design provided that a sufficient amount of steam is co-fed to the
assumptions and the techno-economic constraints are the same reactor to avoid coking.3 In the ‘HICOM’ case, ~40% less
adopted by Martelli in [28] for the “coal to substitute natural methane is produced than in the ‘TREMP’ case, and although
gas” plant. the required cathode air flow is greatly reduced (compared to
Plant ‘IGFC-CCS-DIRECT’
The DIRECT case represents the baseline Table 2 Temperature, pressure, flow rate, and composition of selected streams in the IGFC
IGFC plant, with a Shell gasifier and SOFC- plant shown in Fig. 1
based hybrid power cycle. The hot raw syngas
Stream ID:  G13  G14  G16  G17  G20  G21  P3  P5  P17  P19 
leaving the gasifier is first quenched to 900°C
by mixing with cool (250°C) recycled syngas; Molar composition:                                   
the syngas temperature is further reduced to   H2O                       8.4%  8.4%  8.5%  34.9%  0.0%  28.3%  35.0%  55.0%  0.0%  0.0% 
250°C in syngas coolers that generate   O2                        0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  17.1%  13.3%
intermediate pressure steam. The cooled   AR                        0.9%  0.9%  0.9%  0.6%  0.9%  1.5%  0.6%  0.5%  1.0%  1.0% 
syngas is filtered to remove fly ash (which is   N2                        4.6%  4.6%  4.6%  3.4%  5.0%  8.2%  3.4%  2.8%  82.0%  85.7%
recycled back to the coal mill to boost the   H2                        25.0%  25.2%  25.2%  16.3%  68.0%  19.0%  16.3%  5.4%  0.0%  0.0% 
overall carbon conversion of the gasifier to   CO                        57.4%  57.9%  57.9%  7.1%  22.6%  1.3%  7.1%  3.0%  0.0%  0.0% 
99.8%), and then split to feed the quench   CO2                       0.0%  0.0%  0.0%  26.1%  0.5%  4.6%  26.1%  33.2%  0.0%  0.0% 
cooling recycle loop. The remaining syngas is   H2S                       0.8%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0% 
scrubbed with hot water to remove fugitive   CH4                       2.8%  2.8%  2.8%  11.5%  3.0%  37.0%  11.5%  0.0%  0.0%  0.0% 
particulates and water-soluble impurities, and   NH3                       0.0%  0.0%  0.0%  0.0%  0.0%  0.1%  0.0%  0.0%  0.0%  0.0% 
then sent to the AGR unit to remove sulfur   COS                       0.1%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0% 
compounds. The sweet syngas exiting the AGR
  HCN                       0.1%  0.1%  0.1%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0%  0.0% 
is directly fed to the SOFC power island. In
Total Flow  (kmol/s)     6.12  6.06  6.06  8.26  5.56  3.38  8.26  10.16  55.75  25.02 
case DIRECT, no methanation occurs before
Total Flow  (kg/s)       125.6  123.6  123.5  191.2  56.5  56.5  191.2  269.1  1606  717.2 
the SOFC plant.
Temperature (°C)          129  40  40  683  ‐39  603  700  800  650  800 
Plant ‘IGFC-CCS-TREMP’ Pressure    (bar)            34  31  34  31  34  31  20  20  20  20 
The “TREMP” case employs a Haldor
Topsoe TREMP methanator upstream the SOFC to increase the the DIRECT case), it is not at its lower limit. A multi-stage
methane content of the feed gas. The sour syngas exiting the methanator might achieve this result, although its economic
scrubber undergoes a partial water-gas shift (WGS) to produce viability is less clear.
syngas with a H2/CO molar ratio of 3.0 in preparation for Pressure losses  
methanation. The WGS unit uses a staged design to reduce Pressure losses are an important consideration, since
steam consumption [3,29,30]. Some of the entering syngas is cumulatively the many small pressure drops throughout a plant
combined with steam (to achieve a H2O/CO molar ratio of 2.1 can have significantly reduce its overall efficiency. A full
and thus avoid catalyst coking), heated to 250 °C and reacted in accounting would require a detailed knowledge of the geometry
the first catalytic WGS reactor. The remaining syngas bypasses of each plant component, information that is generally not
the reactor and is combined with its effluent and feed water available. Instead, we assessed and assigned pressure drops for
(again so that H2O/CO = 2.1 and T = 250 °C) before entering each component after carefully reviewing previous studies on
the final WGS reactor. The shifted syngas then enters a systems whose operating pressures similar to those used here
Rectisol-based absorption process to capture both CO2 and H2S. [4,5,32,33]. In the SOFC, an overall pressure loss of 2 bar was
The WGS and AGR units serve to provide a syngas whose assumed for both the air/cathode and the fuel/anode loops
composition is well suited to the TREMP methanation reactor. [5,32,33].
As a result, the methane content in the exiting syngas is quite
high (~37% vol.), with almost all the C in the incoming syngas PLANT PERFORMANCE
having been converted to CH4 in the methanator. Key results for the three IGFC plants are given in Table 1;
Plant ‘IGFC-CCS-HICOM’ note that the HICOM case is the most efficient plant, with a
As with the TREMP configuration, case HICOM (HIgh CO
Methanation) also employs a methanation unit upstream the 3
The amount of steam needed to prevent carbon formation is
SOFC generator to increase the CH4 content of the SOFC feed. determined by calculating full chemical equilibrium at the temperature
The HICOM catalyst is Ni-based [13,14,31] (as is the TREMP and pressure at the methanation reactor inlet, using the thermodynamic
properties of graphite as a model for those of coke.

5 Copyright © 2012 by ASME


LHV efficiency of 50.1%. (Table 2 provides a data on selected The design and capital cost of SOFC modules (Table 4)
streams in the gasifier island, gas cleaning/processing section were calculated following NETL [7,26]; additional costs for
and the SOFC power island.) The performance of the TREMP piping and insulation costs were taken from Siemens
plant suffers from the burden of WGS and CO2 removal prior to Westinghouse [38]. The highly modular SOFC unit is
methanation; the auxiliary power for the AGR (Table 1) is designed as follows. Each pressurized module consists of 64
much higher in the TREMP case than for the other cases blocks of stacked planar cells with metallic interconnectors that
because the CO2 load far exceeds that of H2S. serve as both bipolar plates and gas distributors. Each cell (i.e.
The SOFC air utilization factors listed in Table 1 show that the fundamental unit of a stack) has an active area of 550 cm2,
upstream methanation clearly reduces the amount of cathode air and each block contains 96 cells, or 5.28 m2 of active cell area.
– and parasitic compression power – needed to cool the SOFC. A module of 64 blocks produces ~2 MWe. (The entire SOFC
In fact, the cathode fresh air intake is found to be inversely unit comprises slightly more than 400 modules for a total
proportional to the methane content present in the syngas output around 730 MWe.) Each module is enclosed in its own
feeding the fuel cell. Smaller air flow also leads to lower capital insulated pressure vessel, with an AC/DC inverter,
costs for the SOFC and cathode air turbomachinery. turbomachinary and heat-exchangers.
Interestingly, the net electricity consumption of the air The cost for high-temperature heat-exchangers was
compressor/turbine group resulted decreased in the DIRECT evaluated following Turton et al. [39,40] and cost data reported
case, where cathode air consumption is the highest; the in [41,42]; cost factors of 10 and 20 (relative to carbon steel
efficiency gain due to the increased air flow occurs as more heat-exchangers) were used for heat-exchangers made of high-
waste heat, generated internally the SOFC, is removed from the temperature alloys (for gas temperatures below 850°C) and of
cathode air and then converted to power in the cathode turbine. metal/ceramic alloys (for gas temperatures < 1450 °C and with
The incoming syngas also cools the SOFC to some degree since wall temperatures <1050 °C), respectively.
it enters at 700 °C and exits at 800 °C. In the HICOM case, the
electricity generated by the anode expander is notably higher Integrated pres. SOFC blocks, $/kWe  

112 
than in the other plants because of its enhanced anode flow4, Pressurized enclosure, $/kWe  200 
which also removes a significant part of the SOFC waste heat. DC/AC Inverter, $/kWe     82 
Piping and insulation, $/kWe     125 
Transport and placement, $/kWe  12 
PLANT ECONOMICS Foundations at the site, $/kWe  37 
The economic viability of each plant was estimated using Manufacturing cost, $/kWe     30 
the EPRI TAG revenue requirement methodology [34] in Bare‐Erected‐Cost (BEC), $/kWe  568 
conjunction with the capital cost database of Kreutz et al. [35].

Total‐Plant‐Cost (TPC) , $/Kwe  657 
Economic parameters used to estimate the cost of producing † 2
SOFC cost calculated with a power density of 500 mW/cm  and specific 
2
electricity are given in Table 3. All capital costs are escalated to active area cost of 0.054 $/cm . 
mid-2007 US dollars using the Chemical Engineering Plant ‡
Engineering and contract fees add to the the SOFC module BEC to yield 
Cost Index [36,37]. the TPC. Contingencies fees are not accounted for as the SOFC cost 
presented here represents a target value. 
Coal price, $/GJ HHV       2.04 
Capacity factor   85%  Table 4 SOFC module cost breakdown and main assumptions.
Capital charge rate (CCR), % per yr  14.4% 
Interest during construction (IDC), fraction of TPC  11.4%  Spare SOFC capacity is assumed to be installed along with
Construction period, yr  3  nominal capacity to compensate for degradation of the latter
O&M, fraction of TPC per yr  4%  [26]. The sparing strategy is such that the SOFC power output
CO2 transport & storage, $/tonne  6.1  and efficiency are kept almost constant over the time; as the
SOFC module lifetime, yr   5  cell area-specific-resistance (ASR) increases steadily (at a rate
SOFC degradation rate, ASR % increase per 1000h  0.2  of 0.2% per 1000h), the cell current density is decreased at
Interest during construction (IDC) is based on a 3‐year construction schedule  regular intervals (e.g. every 1000 h) by enabling spare cell
with equal, annual payments, and a discount rate of 10%/yr.  capacity in order to maintain a nearly constant operating
The capital charge rate is applied to total plant cost (TCP) + IDC.  voltage and plant power output. We assume that 100% of the
Table 3 Techno-economic assumptions employed in this study. initial SOFC capacity is replaced every 5 years. With an
assumed degradation rate of 0.2% per 1000h, the overall stack
degradation at the end of life will be ~7%. Consequently, the
4
installed spare capacity will be also ~7% of nominal capacity.
The mass flow of the SOFC anode feed is largest in the HICOM case
due to the addition of steam prior to methanation. Although steam is
also added in the TREMP case, CO2 is removed in the AGR.

6 Copyright © 2012 by ASME


Plant economic performance SOFC, and whose cost is included in the ‘Reactors’ cost item in
A summary of the economic performance of the three IGFC Table C1.
plants is given in Table 5. The total plant cost (TPC), or
“overnight construction cost”, given in Table 3 for each case, Plant component  Overnight cost with BOP (M$) 
includes engineering and overhead, general facilities, balance    'DIRECT'  'TREMP'  'HICOM' 
of plant, and both process and project contingencies. For SOFC power core  497  500  490 
comparison and reference, the results for a Shell-based IGCC- Oxy‐combustor   16  16  16 
CCS power plant are also shown.5 Note that, at an assumed Turbomachinary (within the power island)  159  82  122 
“state-of-the-art”/target SOFC module cost of 657 $/kW AC, all Heat‐exchangers (air recuperator and fuel 
44  38  39 
pre‐heater) 
three IGFC plants produce decarbonized electricity at costs
Reactors (cleaning and fuel processing 
significantly below that of IGCC-CCS. As expected from its 167  282  167 
upstream of  the SOFC) 
superior thermodynamic performance, the HICOM case is seen
Gasifier island (including ASU)  725  724  724 
to achieve the lowest $/kW and LCOE of all plants.
CO2 compression  89  52  89 
Heat recovery steam cycle (HRSC)  199  231  203 
      Electricity   η (LHV) TPC  LCOE 
Total Plant Cost (TCP), 2007M$  1,897  1,926  1,851 
   MWe     M$  $/kWe  $/MWh
IGCC‐CCS  517  33.5%  1,424  3,069  109.6  Table 6 Capital costs sheet for the IGFC cases analyzed in this
† study.
IGFC‐CCS‐DIRECT   859  47.0%  1,897  2,460  85.6 

IGFC‐CCS‐TREMP   841  46.0%  1,926  2,551  89.8 
The ‘TREMP’ case has then lower capital costs for the

IGFC‐CCS‐HICOM   915.4  50.1%  1,851  2,253  78.9  turbomachinary, due to the reduced flows of syngas and

SOFC cost at 657$/kW AC              cathode air feeding the SOFC, but higher for the HSRC, due to
Table 5 LCOE of IGFC plant in this study. a more extended network of HXs.
Effect of the power density on the LCOE of the ‘HICOM’ case
Table 6 provides a more detailed comparison of the Reductions in the cost of SOFC modules could result from
electricity cost components between the most promising IGFC increases in single-cell power density, since higher power
case (HICOM) and the reference IGCC-CCS plant. Even density translates to less active area required. Fig. 3 shows the
though the overnight cost of IGCC-CCS is 25% less than the HICOM case LCOE for assumed SOFC power densities
‘HICOM’ case, the much higher efficiency of the latter plant varying from half the state-of-the-art value of 500 mW cm-2
(true for all IGFC cases) yields lower specific ($/MWh) costs in (the value used in all cost results shown here) up to 1500 mW
every category. cm-2, with the specific cost of the cells fixed at $0.054 cm-2 to
Electricity Cost Components,   be consistent with NETL assumptions [7,26].

2007 $/MWh:  IGFC‐CCS‐HICOM   IGCC‐CCS
Installed capital (at 14,38% of TPI)  43.48  59.22 
O&M (at 4% of TPC per yr)    10.86  14.79 
O&M SOFC modules    3.28  ‐ 
Coal (at 2.04/5 $/GJ, HHV)  15.34  22.90 
CO2 emissions (at 50 $/tonne CO2)  2.07  6.80 
CO2 disposal (at 6.1 $/tonne CO2)  3.92  5.94 
Levelized cost of electricity (LCOE)  78.94  109.65 

SOFC cost at 657$/kW AC         

Table 6 Comparison of the electricity cost components between


the ‘HICOM’ case and the reference IGCC case.

In Table 6 the detail of total capital costs for the three IGFC
cases analyzed is given. Notably, the SOFC power island
contributes nearly one-third of the total plant cost. Compared to
Fig. 3 Effect of the SOFC power density on the LCOE for the
the other two other IGFC cases analyzed, the ‘TREMP’ case ‘HICOM’ case.
has the highest installed capital essentially because of the
expensive Rectisol unit that capture CO2 and H2S before the The figure clearly shows that the LCOE can be reduced
only modestly with further development of SOFC power
5
density (which essentially means better engineered fuel cell
The thermodynamic performance of the IGCC-CCS plant considered electrodes, materials, etc.). Today’s IGFC-CCS would be much
here equals that reported in [3]. However, the LCOE was calculated
more costly than IGCC-CCS, but at a SOFC module cost of
using the economic/financial assumptions used in this study.

7 Copyright © 2012 by ASME


~1400 $/kW, the LCOE of the IGFC ‘HICOM’ case would also gratefully acknowledged for their support and fruitful discussions
equal that of IGCC-CCS. around the topics here presented.

REFERENCES
CONCLUSIONS [1] Hassmann, K., 2001, “SOFC power plants, the Siemens-
The results show that methanation upstream the SOFC is Westinghouse approach”, Fuel Cells, 1(1), pp. 78-84.
favorable only under certain plant design configurations. The [2] Singhal, S.C., Progress in tubular solid oxide fuel cell
more straight-forward is the syngas pathway toward the SOFC, technology, SOFC-VI Symposium, ECS Proceedings Volume,
the better are both the performance and the economic viability 1999.
of the plant, as shown by the efficiency and economic [3] DOE/NETL-2010/1397 report, 2010, Cost and Performance
performance obtained with the HICOM plant. Baseline for Fossil Energy Plants Volume 1: Bituminous Coal
The high capital cost fraction for the SOFC power island is and Natural Gas to Electricity.
generally offset by its high conversion efficiency and low cost [4] Franzoni, A., Magistri, L., Traverso, A. and Massardo,
of CO2 capture. At an SOFC module cost of 1400 $/kW, the A.F., 2008, “Thermoeconomic analysis of pressurized
LCOE of the IGFC ‘HICOM’ case would break-even with that hybrid SOFC systems with CO2 separation”, Energy,
of IGCC-CCS. 33(2), pp. 311-320.
This work suggests that IGFC-CCS plants with commercial [5] Park, S.K., Kim, T.S., Sohn, J.L., Lee, Y.D., 2011, “An
entrained flow gasifiers have the potential to provide an integrated power generation system combining solid oxide
interesting option as future power generators in a carbon- fuel cell and oxy-fuel combustion for high performance
and CO2 capture”, Applied Energy, 88(4) pp.1187-1196.
constrained world.
[6] Grol, E., DOE/NETL-40/080609 report, 2009, “Systems
Analysis of an Integrated Gasification Fuel Cell Combined
NOMENCLATURE
Cycle”.
AGR = acid gas removal [7] DOE/NETL-2011-1482 report, 2011, “Analysis of
ASU = (standalone, cryogenic) air separation unit. Integrated Gasification Fuel Cell Plant Configurations”.
BOP = balance-of-plant
COE = cost of electricity [8] Spallina, V., Romano, M.C., Campanari, S., Lozza, G.,
CCS = carbon capture and storage 2011, “Thermodynamic Analysis and Optimization of IT-
DOE = (US) Department of Energy SOFC-Based Integrated Coal Gasification Fuel”, J. Fuel
FU = (global) fuel utilization Cell Sci. Tech., 8(4).
GT = gas turbine [9] Romano, M.C., Spallina V., Campanari, S., 2011, “Integrating IT-
HHV = higher heating value SOFC and gasification combined cycle with methanation reactor
HRSC = heat recovery steam cycle and hydrogen firing for near zero-emission power generation
HRSG = heat recovery steam generator from coal”, Energy Procedia, 4, pp. 1168-1175.
HX = heat exchanger
[10] Spallina, V., Romano, M.C., Campanari, S., Lozza, G., 2011, “A
IGCC = integrated gasifier combined cycle
SOFC-Based Integrated Gasification Fuel Cell Cycle With CO2
IGFC = integrate gasifier fuel cell
LCOE = levelized cost of electricity Capture, J. Eng. Gas Turbines Power, 133(7).
LHV = lower heating value [11] Li, M., Rao, A.D., Brouwer, J., Samuelsen, G.S., 2010, “Design
MDEA = methyldiethanolamine of highly efficient coal-based integrated gasification fuel cell
NETL = National Energy Technology Laboratory power plants, Journal of Power Sources, 195(17), pp. 5707-5718.
O&M = operation and maintenance [12] Rao, A.D., 1991, “Reactor expander topping cycle”, U.S.
SNG = substitute natural gas Patent No. 4,999,993.
SOFC = solid oxide fuel cell
[13] Twigg, M., 1996, “Catalyst Handbook”, 2nd edition, pp.
TIT = turbine inlet temperature
TPI = total plant investment 374-375.
TPC = total plant cost [14] Udengaard, N.R., Olsen, A., Wix-Nielsen, C., 2006, “High
WGS = water gas shift Temperature Methanation Process – Revisited”, 23rd
λ = (global) cathode air excess ratio Annual International Pittsburgh Coal Conference,
September 2006.
ACKNOWLEDGMENTS [15] Föger, K., 2010, “BlueGen – Ceramic Fuel Cells First Product for
Andrea Lanzini is grateful to the US/Italy Fulbright Commission Commercial Roll-out”, Fuel Cell Seminar presentation 2010, San
for the grant awarded to support him in spending a research period in Antonio, TX (USA), October 2010.
Princeton University, within the Energy Group of the Princeton [16] Bose, D., Batawi, E.E., Couse, S., Hickey, D., Mcelroy, J., 2007,
Environmental Institute, as a visiting graduate student. “Solid Oxide Fuel Cell system with internal reformation”, Patent
Prof. Robert Socolow and the other senior members of the No. WO 2008/123968.
Princeton Energy Group (Dr. Bob Williams and Dr. Eric Larson) are [17] Lanzini, A., Leone, P., 2010, “Experimental investigation
of direct internal reforming of biogas in solid oxide fuel

8 Copyright © 2012 by ASME


cells”, International Journal of Hydrogen Energy, 35(6), [33] Adams, T.A., Barton, P.I., 2010, “High-efficiency power
pp. 2463-2476. production from natural gas with carbon capture”, Journal
[18] Lanzini, A., Santarelli, M., Orsello, G., 2010, "Residential of Power Sources, 195(7), pp. 1971-1983.
Solid Oxide Fuel Cell Generator Fuelled by Ethanol: Cell, [34] Electric Power Research Institute (EPRI). Technical
Stack and System Modelling with a Preliminary Assessment Guide, Volume 1: Electricity Supply. Report
Experiment", Fuel Cells, 10(4), pp. 654-675. number TR-102276-V1R7, 1993.
[19] Martelli, E., Amaldi, E., Consonni, S., 2011, “Numerical [35] Kreutz, T.G., Larson, E.D., Liu, G., Williams, R.H.
optimization of heat recovery steam cycles: Mathematical Fischer-Tropsch Fuels from Coal and Biomass. 25th
model, two-stage algorithm and applications”, Computers Annual International Pittsburgh Coal Conference,
and Chemical Engineering, 35(12), pp. 2799-2823. Pittsburgh (PA), USA, 2008.
[20] Martelli, E., Kreutz, T.G., Carbo, M., Consonni, S., Jansen, [36] Vavatuk, W.M., “Updating the CE Plant Cost Index”,
D., 2011, “Shell coal IGCCS with carbon capture: Chemical Engineering, 62-70, Jan. 2002.
Conventional gas quench vs. innovative configurations”, [37] “Chemical Engineering” magazine, monthly. See
Applied Energy, 88(11), pp. 3978-3989. http://www.che.com/pci/.
[21] Rostrup-Nielsen, J.R, Pedersen, K., Sehested, J., 2007, [38] Lundberg, W.L., Israelson, G.A., Holmes, R.A., Zafred,
“High temperature methanation”, Applied Catalysis A: P.R., King, J.E., Kothmann, R.E, 2000, “Pressurized solid
General, 330, pp. 134–138. oxide fuel cell/gas turbine power system”, Siemens
[22] Pedersen, K., Skov, A., Rostrup-Nielsen, J.R., 1980, Westinghouse.
“Catalytic aspects of high temperature methanation”, ACS [39] Turton, R., et al., 2003, “Analysis, Synthesis, and Design
Fuel Chem. Div. Preprints, 25(2), pp. 89-100. of Chemical Processes”, Third Edition, Prentice Hall.
[23] Lanzini, A., 2011, “Study of solid oxides fuel cell systems [40] Green, D.W., 2007, “Perry's Chemical Engineers'
running on poly-fuel mixtures”, Ph.D. Thesis, Department Handbook”, Eight Edition, McGraw-Hill.
of Energy, Politecnico di Torino, Torino, Italy. [41] Kuramochi, T., Turkenburg, W., Faaij, A., 2011,
[24] Agrawal, G., 2009, “Advances in Fuel Cell Blowers”, 10th “Competitiveness of CO2 capture from an industrial solid
Annual SECA Workshop, Pittsburgh, PA (USA). oxide fuel cell combined heat and power system in the
[25] Milewski, J., Świrski, K., Santarelli, M., Leone, P., 2011, early stage of market introduction”, Fuel, 90(3), pp. 958-
“Advanced Methods of Solid Oxide Fuel Cell Modeling”, 973.
First Edition, Springer. [42] DOE/NETL-2002/1169 report, 2002, “Process equipment
[26] DOE/NETL-2011/1486 report, 2011 “Analysis of Natural cost estimation: final report.
Gas Fuel Cell Plant Configuration”'. [43] Guangjian, L., Larson, E.D., Williams R.H., Kreutz T.G,
[27] Linnhoff, B., Hindmarsh, E., 1983, “The pinch design Guo, X., 2011, “Making Fischer−Tropsch Fuels and
method for heat exchanger networks”, Chemical Electricity from Coal and Biomass: Performance and Cost
Engineering Science, 38(5), pp. 745-763. Analysis”, Energy Fuels, 25(1), pp. 415–437.
[28] Martelli, E., 2010, “Numerical optimization of heat [44] Haldor Topsøe, 2009, “From solid fuels to substitute
recovery steam cycles for highly integrated energy natural gas (SNG) using TREMP™”, www.topsoe.com
systems”, Ph.D. Thesis, Department of Energy, Politecnico (last accessed October 31, 2011)
di Milano, Milano, Italy. [45] Gerdes, K., Grol, E., Kearins D., Newby, R., DOE/NETL-
[29] Carbo, M.C., Boon, J., Jansen, D., van Dijk, H.A.J., 2009/1361, 2009, “Integrated Gasification Fuel Cell
Dijkstra, J.W., van den Brink, R.W., Verkooijen, A.H.M., Performance and Cost Assessment”.
2009, “Steam demand reduction of water–gas shift reaction [46] Schilke, P.W., 2004, “Advanced Gas Turbine Materials
in IGCC power plants with pre-combustion CO2 capture”, and Coatings”, GE energy.
International Journal of Greenhouse Gas Control, 3(6), pp.
712-719.
[30] Carbo, C., Jansen, D., Boon, J., Dijkstra J.W., van den
Brink, R.W. Verkooijen, A.H.M., “Staged water-gas shift
configuration: Key to efficiency penalty reduction during
pre-combustion decarbonisation in IGCC”, Energy
Procedia, 1(1), pp. 661-668.
[31] Penner, S.S., 1987, “Coal Gasification: Direct
Applications and Syntheses of Chemicals and Fuels”,
University of California, San Diego.
[32] Trasino, F., Bozzolo, M., Magistri, L. and Massardo, A.F.,
2011, “Modeling and Performance Analysis of the Rolls-
Royce Fuel Cell Systems Limited: 1 MW Plant”, Journal
of Engineering for Gas Turbines and Power-Transactions
of the ASME, 133(2).

9 Copyright © 2012 by ASME


ANNEX A

Table A1 Process design parameter assumptions for ASPEN Plus.


Unit  Description 
Cryogenic air separation for the oxy‐combustor: 99.5 (mol%) O2 production [3,7]; power (kWh/t pure O2 at atmospheric pressure) = 26.32 
Air Separation Unit (ASU) 
[3].  
Coal milling & handling + 
Electricity requirement = 0.24% of input coal (MWel/MW(HHVcoal)) [3]. 
ash handling  
Fresh air composition  Air composition (% vol.): O2, 20.95%; N2, 78.12%; Ar =0.93% . 
Coal composition  Illinois #6 bituminous coal, no. 2 [3]. 
Shell gasifier with convective syngas coolers (T = 1391 °C, p = 38.5 bar). Losses: 1.5% of the inlet coal HHV as heat loss at the membrane 
Gasifier 
wall; 1% of scrubbed syngas for coal drying. 
Physical CO2 and sulfur absorption with chilled methanol at ‐40 °C; syngas recycle is applied to minimize the amount of fuel co‐absorbed 
Rectisol unit 
with the CO2. CO2 is then separated and compressed from a cascade of four flash separators [20]. 

2‐stage WGS adiabatic reactor with bypass line and an inlet  steam/CO ratio of 2.1 at each stage; the outlet syngas has an imposed  H2/CO 
WGS unit 
ratio of 3 as requested downstream by the TREMP™ catalyst of the methanator [44]. The inlet syngas temperature in the reactor is 250 °C.  

Methanator with TREMP™  Single adiabatic reactor with inter‐cooled recycle with an inlet H2/CO ratio of 3. The catalyst is the Ni‐based one used in the TREMP™ process 
catalyst   by Haldor TOPSOE. The inlet syngas temperature in the reactor is 250 °C. 
Methanator with HICOM  High CO direct methanation (HICOM) Ni‐based catalyst employed in single adiabatic reactor with inter‐cooled recycle; an inlet steam/CO 
catalyst  ratio of 1.5 is used to avoid coking. The inlet syngas temperature in the reactor is 250 °C. 
Cathode air compressor  Not‐intercooled: ηis  = 0.83; ηmech+el  = 0.95. 
Cathode turbine  Not‐intercooled: ηis  = 0.89; ηmech+el  = 0.95. 

Anode expander  Steam/CO2 expander  (un‐cooled, with TIT max. set to 850 °C): ηis  = 0.85; ηmech  = 0.95. 
Oxy‐combustor  Oxygen excess stoichiometry: 1.01 mol [45] 
4 stages inter‐cooled at 30 °C with liquid water knock‐out. Pressure ratio of each stage is β = 1.21 and discharge pressure is of 43 bar. ηpol  = 
CO2 compression train 
0.73; ηmech+el  = 0.95. Last stage (from 43 to 150 bar): ηis  = 0.75; ηmech+el  = 0.95. 

‘DIRECT’ case optimized configuration: 3 pressure levels with reheat; LP, MP and HP levels are 8.4, 47.2 and 139.5 bar, respectively; SH and 
RH temperatures (LP, MP/RH and HP/SH) are 171, 440 and 383 °C; steam mass flow rates (LP, MP and HP) are  27.49, 186.52 and 5.53 kg/s.
Heat recovery steam cycle  ‘ TREMP’ case optimized configuration: 3 pressure levels with reheat; LP, MP and HP levels are 4.0; 32.3 and 120.0, respectively; SH and RH 
(HRSC)  temperatures (LP, MP/RH and HP SH) are 270, 520.00 and 580.00; steam mass flow rates (LP, MP and HP) are  37.80 and 201.39 kg/s.  
‘HICOM’ case optimized configuration: 3 pressure levels with reheat; LP, MP and HP levels are 4.0, 58.8 and 143.1 bar, respectively;  SH and 
RH temperatures (LP, MP/RH and HP SH) are 142, 437 and 477 °; steam mass flow rates (LP, MP and HP) are  14.27, 21.51 and 202.84 kg/s. 


Due to limitations on the blade material, uncooled turbines cannot withstand temperatures above around 850 °C [46]. 

10 Copyright © 2012 by ASME

You might also like