Anda di halaman 1dari 13

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

Available at www.sciencedirect.com

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

Development of the once-through hybrid sulfur process for nuclear hydrogen production
Yong Hun Jung, Yong Hoon Jeong*
KAIST (Korea Advanced Institute of Science and Technology), 335 Gwahangno, Yuseong-gu, Daejeon 305-701, Republic of Korea

article info
Article history: Received 24 March 2010 Received in revised form 28 July 2010 Accepted 30 July 2010 Available online 16 September 2010 Keywords: Hybrid sulfur Nuclear hydrogen Sulfur Sulfur combustion Sulfur dioxide Sulfuric acid

abstract
The Once-through Hybrid Sulfur (Ot-HyS) process, proposed in this work, produces hydrogen using the same Sulfur dioxide Depolarized water Electrolysis (SDE) process found in the original Hybrid Sulfur cycle (HyS). In the process proposed here, the Sulfuric Acid Decomposition (SAD) process in the HyS procedure is replaced with the well-established sulfur combustion process. First, a ow sheet for the Ot-HyS process was developed by referring to existing facilities and to the work done by the Savannah River National Laboratory (SRNL) under their reasonable assumptions. The process was then simulated using Aspen Plus with appropriate thermodynamic models. It was demonstrated that the Ot-HyS process has higher net thermal efciency, as well as other advantages, over competing benchmark processes. The net thermal efciency of the Ot-HyS process is 47.1% (based on LHV) and 55.7% (based on HHV) assuming 33.3% thermal-to-electric conversion efciency of a nuclear power plant with no consideration given to the work for the air separation. Hydrogen produced through the Ot-HyS process would be used as off-peak electricity storage, to relieve the burden of load-following and could help to expand applications of nuclear energy, which is regarded as a sustainable development technology. 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.

1.

Introduction

Humanity has been facing major energy challenges such as increasingly severe climate change, threats to energy security, and a global energy shortage that is especially dire in the developing world. Growing awareness of global warming has led to efforts to develop sustainable energy technologies benecial to the economy, social welfare, and the environment. Water-splitting nuclear hydrogen production is expected to help to resolve those challenges when high energy efciency and low cost hydrogen production become possible. Among sulfur-based water-splitting thermo-chemical cycles which are recognized as high priority candidates for research and development, the Hybrid Sulfur cycle (HyS), rst proposed

by Westinghouse Electric Corp., has been researched by Savannah River National Laboratory (SRNL) under the Nuclear Hydrogen Initiative (NHI) established by the U.S. Department of Energys Ofce of Nuclear Energy, Science and Technology (DOE-NE). SRNL has recently developed conceptual designs of HyS processes using sulfur dioxide depolarized electrolyzers now being developed and that are powered by an advanced nuclear reactor heat source such as Pebble Bed Modular Reactor (PBMR) [1,2]. The HyS cycle, also known as the Westinghouse Process, is a combined, all-uids cycle of the electrolysis and thermochemical processes with only two chemical reactions and only one additional element (sulfur) outside of hydrogen and oxygen [3e7]. In the HyS cycle, hydrogen and sulfuric acid are

* Corresponding author. Tel.: 82 42 350 3826; fax: 82 350 3810. E-mail addresses: moonrivery@kaist.ac.kr (Y.H. Jung), jeongyh@kaist.ac.kr (Y.H. Jeong). 0360-3199/$ e see front matter 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.ijhydene.2010.07.168

12256

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

produced by sulfur dioxide depolarized water electrolysis, the so-called SDE process. The presence of sulfur dioxide in addition to the water in SDE reduces the required cell potential; in other words, the electric energy required for this process is well below that required for conventional pure water electrolysis [8,9]. In order to supply sulfur dioxide and water to the SDE process again, the sulfuric acid product at the anode is recycled to the high-temperature, catalytic, vapor phase thermal decomposition of sulfuric acid (SAD), which is commonly found in all sulfur-based cycles as one of the most challenging technical issues and as the largest energy consumer in all sulfur-based cycles [10e12]. It is challenging to nd materials to decompose sulfuric acid that are both chemically and thermally stable in highly corrosive environments. A SiC bayonet decomposition reactor has been developed by Sandia National Laboratories (SNL) and experiments have been performed up to 900  C with no severe corrosion issues [13,14]. Although corrosion problems have been partly addressed, more emphasis is needed with respect to the scaling up process for industrial application. The availability and cost of the components for installations and plant maintenance may determine the overall success for commercial applications. This is one of the leading motives which have driven the development of the Ot-HyS process. In the Once-through Hybrid Sulfur process proposed in this work, however, this challenging issue, the high-temperature SAD process, is replaced by the sulfur combustion process, which is a well-established technology in manufacturing sulfur dioxide as a nal product or as an intermediate product in manufacturing sulfuric acid. Although the concept of the once-through process for hydrogen production has been proposed and studied before, under the name of the openloop thermo-chemical cycle, there has been no detailed analysis and there are many differences between the details of that process and the process described in this study [15e17]. The purpose of this study is to demonstrate that the proposed Ot-HyS process is an energy-efcient hydrogen production process that can help to resolve ongoing energy challenges. To achieve this purpose, the scope of the study was established as the development of a detailed ow sheet for the simulation rst and then, based on that, the estimation of the net thermal efciency of the Ot-HyS process through an Aspen Plus simulation with appropriate thermodynamic models. Finally, a study was conducted to compare the net thermal efciency of the proposed process with those of other benchmark processes for hydrogen production.

produce 1 mol of hydrogen and sulfuric acid, the process inputs are a certain amount of electric and thermal energy for the process as well as 1 mol of sulfur and oxygen or corresponding air and 2 mol of water. In the real process simulation, 3.83093 mol of water should be supplied in compensation for 1.79616 mol of water extracted together with 1 mol of sulfuric acid assuming 75 wt% of the acid concentration, and 0.03477 mol of water extracted together with waste purge. Differences in the reaction steps of the HyS and Ot-HyS processes are described in Fig. 1. Because the technically challenging SAD process is replaced with the well-established sulfur combustion process as stated above, the Ot-HyS process bears less technical challenges than the original HyS cycle does. Therefore, it is expected to advance the realization of large-scale nuclear hydrogen production by feeding an initial nuclear hydrogen stock. For environmental reasons, the sulfur feed for the SDE process can be supplied by a desulfurization process to recover sulfur during the processing of natural gas and petroleum rening. This kind of sulfur recovered for environmental

Hybrid sulfur cycle


H2O
Electricity H2

SO2 + 2H2O H2 + H2SO4


INPUTS : 1. Water 2. Heat ( 800 3. Electricity OUTPUTS : 1. Hydrogen ) 2. Oxygen 3. Waste Heat

SO2 + H2O

H2SO4(H2O)

O2 + SO2 + H2O H2SO4


O2
Heat ( 800 )

Once-through hybrid sulfur process


2H2O
Electricity H2

2.

Background

SO2 + 2H2O H2 + H2SO4


INPUTS : OUTPUTS : 1. Recovered S 1. Hydrogen 2. Water & Oxygen 2. Sulfuric Acid 3. Electricity 4. Heat for Acid Concentration

The Ot-HyS process proposed in this work is one of the watersplitting nuclear hydrogen production processes and is based on the same SDE process found in the HyS cycle used to produce hydrogen. However, because the Ot-HyS process does not include the SAD process, sulfuric acid produced in the SDE process is just extracted from the whole process after it has reached the desired acid concentration. Sulfur dioxide for the continuous operation of the SDE process is prepared by the sulfur combustion process in which sulfur is burned in the presence of oxygen to produce sulfur dioxide. Therefore, to

Heat for Acid


Concentration

SO2

H2SO4(H2O)

SO2 S + O2
Recovered S

O2 (Air)

Fig. 1 e Reaction steps of the HyS and Ot-HyS processes.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

12257

reasons, the so-called recovered sulfur, is now the predominant source of sulfur worldwide and is expected to increase signicantly, even resulting in severe oversupply due to rising environmental concerns, expanded oil and gas operation and the development of oil sands [18,19]. Sulfur statistics are being developed in favor of the Ot-HyS process. In addition, the sulfur combustion itself is also favorable to the process in terms of the process efciency. The process efciency can be enhanced by recovering the great amount of high-temperature excess heat from the sulfur combustion process to supply process electricity as well as heat duty, because sulfur combustion is a highly exothermic chemical reaction releasing 297 kJ/mol of combustion heat and results in a large furnace temperature rise (generally above 1000  C). This high grade excess heat from the sulfur combustion process can make it possible to employ the advanced Power Conversion System (PCS) which generates process electricity at a high thermal-to-electric conversion efciency. The sulfuric acid product, which is another product of the Ot-HyS process, is also favorable to the process in terms of the sulfur statistics. It is reasonable to assume that nearly all the environmental sulfur is consumed after it is rst converted into sulfuric acid and sulfuric acid is the leading sulfur end-use in all forms. Although the long-term worldwide oversupply situation is likely to continue, sulfuric acid consumption will also continue to increase at a low rate [20,21]. Therefore, sulfuric acid produced by the Ot-HyS process instead of by a sulfuric acid manufacturing plant could meet this additionally rising sulfuric acid demand by feeding the severely increasing sulfur surplus recovered for environmental reasons. The sulfuric acid product from the Ot-HyS process could be supplied to the market or consumed directly near the plant site. If the sulfuric acid product is directly provided to neighboring consumers, it may be concentrated as necessary for individual customers. According to demand, the sulfuric acid product may not be concentrated at all. In this case, the process efciency could be increased by as much as the amount of reduced heat duty for the thermal sulfuric acid concentration process. In the sulfur furnace, a side reaction of sulfur dioxide with oxygen forms sulfur trioxide. This is an exothermic equilibrium reaction with a very slow reaction rate and is the reverse reaction of the high-temperature SAD (SO3 decomposition). In the sulfuric acid manufacturing plant, this reaction, or socalled Conversion process, is favorable and is positioned in the second stage after the rst sulfur combustion stage because sulfur trioxide is needed to produce sulfuric acid by being absorbed into water. Therefore, several catalyst beds, generally vanadium pentoxide (V2O5), which is only effective above its melting point of 400  C, are used to increase the reaction rate. The product of each bed should be cooled because sulfur trioxide is readily formed at temperatures of about 400e700  C. Both a catalyst and cooling are needed for reaction kinetics and equilibrium [20,21]. For the purpose of sulfur dioxide production in this work, however, the formation of sulfur trioxide is undesirable and should be suppressed. By maintaining a high sulfur dioxide concentration, sulfur trioxide formation can be suppressed due to the temperature dependency of the reaction equilibrium. If a high sulfur dioxide concentration can be achieved, then the

furnace temperature becomes high enough to suppress the formation of sulfur trioxide. In this work, for the simplication of the simulation, the formation of sulfur trioxide is disregarded. This is because, under the given conditions of this work, a high sulfur dioxide concentration and a high furnace temperature without any catalyst, only a negligible quantity of sulfur trioxide can be formed. Temperatures, oxygen concentration, and residence times in the typical air sulfur furnace make possible the formation of thermal NOx [20,21]. The formation of NOx has its maximum value at somewhere between 18e20% of sulfur dioxide concentration. Although special staged-combustion technology and gas recycling can be used to reduce NOx formation, nitrogen, NOx and Ot-HyS process may not be a good match. One of the advantages of the HyS cycle including the Ot-HyS process is that it has a small number of additional elements, chemical reactions, and separation steps. The HyS cycle has only sulfur as the additional element outside of hydrogen and oxygen. This advantage can be deteriorated by the addition of undesirable nitrogen and NOx if atmospheric air is directly used to burn the sulfur like in most sulfur burning acid plants. Undesirable nitrogen and NOx from sulfur combustion with the air eventually results in the increase in chemical reactions, separation steps, as well as the size of equipment because nitrogen makes up about 79% by volume of air. For that reason, especially in the Ot-HyS process, it is favorable that the oxygen used is as pure as possible. Oxygen may be prepared by an air separation system. In this work, for the simplication of the simulation, sulfur is burned in the presence of the pure oxygen with no consideration given to the work for the air separation. Additional work or cooling load for the separation process would result in a few percent loss of the net thermal efciency of the overall Ot-HyS process.

3.
3.1.

Analysis
Flow sheet development

To estimate the net thermal efciency of the proposed Ot-HyS through Aspen Plus, which is a renowned chemical process simulator, an appropriate ow sheet and thermodynamic models are required. The ow sheet for the Ot-HyS process is shown in the Figs. 2 and 3. Stream conditions and compositions are detailed in Table 1 and process net inow and outow are summarized in Table 2. The main ow sheet and thermodynamic model are based on SRNLs recent work. All the assumptions which were used in SRNLs work are also applied in this study [1,2]. PEM-type SDE is treated as a black box operated at 100  C and 20 bar under the development performance targets of the SRNL, such as a cell potential of 0.6 V versus a current density of 500 mA/cm2 and a sulfur dioxide conversion of 40%. The acid concentration of the SDE spent anolyte and the nal sulfuric acid product after the concentration process are set to be 50 wt% and 75 wt%, respectively. A nine-equilibrium staged vacuum distillation column including a partial condenser and a kettle reboiler is used to concentrate the spent anolyte using thermal energy.

12258

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

Fig. 2 e Flow sheet for the Ot-HyS process.

In addition to these elements, the main ow sheet consists of an eight-equilibrium staged sulfur dioxide absorber, a threestaged sulfur dioxide recycle compressor, and an anolyte preparation tank. It also includes several knock-out drums, pumps, valves, and two vacuum ejectors. Many water coolers are required throughout the whole ow sheet under the assumption that the process cooling water is able to cool down the process streams to about 40  C. Two recuperative

preheaters are also required. They are positioned in a row before the vacuum distillation column. The rst recuperative preheater transfers heat from all the streams which have too high a temperature to just release into the cooling water. The main ow sheet was simulated using the OLI-MSE model, which has been found to work very well in the H2SO4eH2O system over the entire concentration range and to temperatures below 500  C [22].

Fig. 3 e Flow sheet for the SCHRS including PCS.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

12259

Table 1 e Ot-HyS ow sheet stream table. ID Phase H2O


C1 C2 C3 C4 C5 C6 C7 C8 C9 A1 A2 A3 A4 A5 A6 A7 A8 A9 A10 A11 A12 A13 A14 A15 A16 A17 A18 A19 A20 A21 A22 A23 A24 A25 A26 A27 A28 A29 A30 A31 A32 A33 A34 A35 A36 A37 A38 A39 A40 B1 B2 B3 B4 B5 B6 B7 B8 B9 B10 B11 S10 LV LV V V L L L L L L L L V LV V LV V V V L L V L L V V V L L V L L V L L L L LV L L V V V L L L e L L L LV L LV L LV LV LV L L L V 9 9 0.02185 0 8.97815 0.02185 1 10 10 26.4794 21.05779 21.05779 0.09086 0.09086 0.03731 0.03731 0.00267 0.0035 0.00617 0.00591 0.00591 0.00026 0.03162 0.03162 0.00595 0.02433 0.03028 0.02886 0.03477 0.00142 0.06825 0.06825 0.02403 0.01562 0.01562 0.0084 0.0084 0.00001 0.00001 0.00841 0 0.00116 0 0.00116 2.8031 6.39641 0 27.4794 27.4794 5.42161 5.42161 5.33075 5.33075 5.29344 5.29344 5.29344 5.29344 1.79616 3.4946 3.4946 0

Component mole ow [kmol/sec] H2SO4


0 0 0 0 0 0 0 0 0 4.88405 3.88405 3.88405 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 3.88405 3.88405 1 1 1 1 1 1 1 1 1 0 0 0

T [ C] S
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

P [bar]

SO2
0 0 0 0 0 0 0 0 0 1.5 1.19288 1.19288 0.28671 0.28671 0.01769 0.01769 0.00083 0 0.00083 0.00002 0.00002 0.00081 0.00012 0.00012 0.01838 0 0.01838 0.00041 0.00043 0.01797 0.00096 0.00096 0.30371 0.00065 0.00065 0.29972 0.29972 0.00334 0.00296 0.30268 0.00038 0.00852 0 0.00852 0 0.99527 0 2.5 2.5 0.30712 0.30712 0.02042 0.02042 0.00272 0.00272 0.00272 0.00272 0 0.0019 0.0019 1.00043

O2
0 0 0 0 0 0 0 0 0 0.00353 0.0028 0.0028 0.00072 0.00072 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.00072 0 0 0.00001 0.00001 0.00071 0 0.00001 0.00071 0.25797 0.25797 0 0 0.00071 0 0.00353 0.00353 0.00072 0.00072 0 0 0 0 0 0 0 0 0 0.25797

H2
1.00229 1.00229 1 1 0.00229 0 0 0.00229 0.00229 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Total
10.00229 10.00229 1.02185 1 8.98044 0.02185 1 10.00229 10.00229 37.75103 30.02158 30.02158 0.37829 0.37829 0.055 0.055 0.0035 0.0035 0.007 0.00593 0.00593 0.00107 0.03174 0.03174 0.02433 0.02433 0.04866 0.02927 0.0352 0.01939 0.06921 0.06921 0.32846 0.01627 0.01627 0.30813 0.30813 0.00406 0.00297 0.3111 0.00109 0.26766 0.25797 0.00968 2.8031 7.3924 0 37.75103 37.75103 7.72945 7.72945 7.35116 7.35116 7.29616 7.29616 7.29616 7.29616 3.79616 3.4965 3.4965 1.25841 100.00 76.96 76.96 40.00 76.96 40.00 40.00 73.19 73.20 100.00 100.00 80.00 83.87 40.00 80.18 40.00 38.66 169.99 94.01 40.00 40.04 40.00 40.00 41.24 40.00 169.99 93.41 40.00 40.02 40.00 40.00 41.20 40.00 40.00 41.11 40.00 41.25 40.00 40.00 41.24 40.00 42.44 40.00 42.44 40.00 75.12 84.12 78.77 100.00 83.87 83.87 80.18 80.18 66.92 79.72 94.41 123.62 38.66 38.86 75.00 20.00 20.00 20.00 20.00 20.00 20.00 20.00 20.00 21.00 20.00 20.00 20.00 1.01 1.01 0.30 0.30 0.09 7.91 0.30 0.30 1.01 0.30 0.30 21.00 0.30 7.91 1.01 1.01 1.01 1.01 1.01 21.00 1.01 2.78 21.00 7.65 21.00 21.00 21.00 21.00 21.00 21.00 21.00 21.00 21.00 21.00 21.00 21.00 20.00 1.01 1.01 0.30 0.30 0.11 0.11 0.11 0.13 0.09 21.00 21.00

(continued on next page)

12260

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

Table 1 (continued ) ID Phase H2O


A34 S0 S1 S11D S2 S3 S3D S3 S4 S5 S6 S7 S8 S9 R1 R2 R3 R4 R5 R6 S9D S9 S10 S11 V L V V V V V V V V V V V V L V V V M L V V V V 0 0 0 0 0 0 0 0 0 0 0 0 0 0 3.22867 3.22867 3.22881 3.22881 3.22867 3.22867 0 0 0 0

Component mole ow [kmol/sec] H2SO4


0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

T [ C] S
0 1.00043 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

P [bar]

SO2
0 0 0 4.00172 4.00172 5.00215 5.00215 5.00215 5.00215 5.00215 5.00215 5.00215 5.00215 5.00215 0 0 0 0 0 0 5.00215 5.00215 1.00043 4.00172

O2
0.25797 0 1.00043 1.0319 2.2903 1.28987 1.28987 1.28987 1.28987 1.28987 1.28987 1.28987 1.28987 1.28987 0 0 0 0 0 0 1.28987 1.28987 0.25797 1.0319 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

H2

Total
0.25797 1.00043 1.00043 5.03362 6.29202 6.29202 6.29202 6.29202 6.29202 6.29202 6.29202 6.29202 6.29202 6.29202 3.22867 3.22867 3.22881 3.22881 3.22867 3.22867 6.29202 6.29202 1.2584 5.03362 40.00 132.22 40.00 75.00 67.59 1078.99 1078.99 1078.99 895.44 807.90 189.65 280.10 100.00 75.00 41.78 540.00 333.58 540.00 60.07 40.00 75.00 75.00 75.00 75.00 21.00 10.34 21.00 21.00 21.00 10.34 10.34 10.34 10.34 10.34 10.34 21.00 21.00 21.00 160.00 160.00 40.00 40.00 0.20 0.20 21.00 21.00 21.00 21.00

The biggest difference between the Ot-HyS and HyS ow sheet comes from the replacement of the SAD process with the sulfur combustion process. A detailed ow sheet for the sulfur combustion process and the Sulfur Combustion Heat Recovery System (SCHRS), including the PCS (in this case, the Rankine cycle) was developed by referring to existing facilities. The process was simulated using an ideal model, which is generally used for the sulfur combustion system at hightemperature gas phase, and the STEAMNBS model, which is the recommended property model for the steam cycles including compressors and turbines. The pure oxygen sulfur combustion process was developed by referring to the liquid sulfur dioxide manufacturing plant developed by Quimetal in Chile [20]. In this conguration, molten sulfur fed to the furnace is burned in the presence of a gas mixture containing oxygen and sulfur dioxide. This gas mixture consists of the following elements: rst, an internally recycled stream to maintain the moderate operating temperature of the sulfur furnace (1079  C); second, an externally recycled oxygen stream from the sulfur dioxide absorber; and last, an additional feed stream of dried pure oxygen to make up for the decient oxygen feed. A small amount of sulfur dioxide loss purged through the ejector blow-downs is compensated in the sulfur combustion process by feeding more molten sulfur and oxygen of the same amount. The resulting combustion gas is regulated to 79.5% of sulfur dioxide and 20.5% of oxygen. The furnace temperature of the sulfur combustion with pure oxygen at this level of high gas strength will become very high and therefore cause some problems. However, the sulfur furnace can be maintained at a moderate operating temperature just by recycling a large amount of cold combustion gas to the sulfur furnace again.

The resulting combustion gas is cooled in the SCHRS inside the sulfur furnace, for the process heat duty of the vacuum distillation columns reboiler, and for the electricity generation in the PCS. After that, it really leaves the sulfur furnace. As a PCS, a simple steam Rankine cycle was used to convert the recovered heat of sulfur combustion for electricity. The maximum steam pressure and temperature of the Rankine cycle were set to be 160 bar and 540  C, respectively, with reference to the 200 MW units at Western Powers Kwinana Power Station [23]. Steam is superheated and reheated once each and is expanded twice in the high pressure turbine and low pressure turbine which have 90% of the isentropic efciencies. The back pressure of the low pressure turbine was set as 0.2 bar for the condensation pressure and temperature limitations under the assumption that the overall process streams can be cooled to 40  C with process cooling water. The back pressure of the high pressure turbine was set as 40 bar for the vapor fraction limitation of the low pressure turbine downstream. At this pressure, the vapor fraction leaving the low pressure turbine is high enough (about 95%). After pressurization to 21 bar using a fresh SO2 compressor, the combustion gas leaving the sulfur furnace is cooled to 100  C for the process heat duty of the second preheater, and is then cooled again to 75  C, which is slightly above its boiling point, using process cooling water. Because energy transferred to the stream by the fresh SO2 compressor should be recovered as much as possible, the process heat duty for the second preheater is recovered after the fresh SO2 compressor. After that, the majority of the combustion gas is recycled to the sulfur furnace to avoid a high furnace operating temperature. It does not affect the gas concentration because it is the same concentration as the gas existing in the sulfur furnace.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

12261

Table 2 e Ot-HyS ow sheet net inow and outow. ID Phase H2O


1. SO2 depolarized electrolysis Cathode inow Cathode outow Cathode reaction Anode inow Anode outow Anode reaction SDE net reaction (Outow-Inow) 2. Ot-HyS except sulfur combustion Hydrogen production Sulfuric acid production Oxygen separation Waste purge Total outow Additional water for SDE Sulfur combustion gas Steam for vacuum ejector #1 Steam for vacuum ejector #2 Make-up anolyte Total inow Net ow (Outow-Inow) 3. Ot-HyS including sulfur combustion Hydrogen production Sulfuric acid production Waste purge Total outow Sulfur feed for sulfur combustion Oxygen feed for sulfur combustion Additional water feed for SDE Steam for vacuum ejector #1 Steam for vacuum ejector #2 Make-up anolyte Total inow C9 C1 Diff. A40 A1 Diff. Diff. C4 B9 A34 A20 V L V L L M 10 9 1 27.4794 26.4794 1 2 0 1.79616 0 0.03477 1.83093 1 0 0.02433 0.0035 2.8031 3.83093 2 V L L 0 1.79616 0.03477 1.83093 0 0 1 0.02433 0.0035 2.8031 3.83093 2

Component mole ow [kmol/sec] H2SO4


0 0 0

T [ C] P [bar] S Total
10.00229 73.20 10.00229 100.00 0 26.8 37.75103 78.77 37.75103 100.00 0 21.23 21.00 20.00 1 21.00 20.00 1

SO2
0 0 0 0 0 0

O2

H2
0.00229 1.00229 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

L L

3.88405 2.5 4.88405 1.5 1 1 1 0 1 0 0 1 0 0 0 0 0 0 1 0 1 0 1 0 0 0 0 0 0 0 1 1 0 0 0 0.00043 0.00043 0 1.00043 0 0 0 1.00043 1 0 0 0.00043 0.00043 0 0 0 0 0 0 0

0.00353 0 0.00353 0 0 0 0 0 0 0.25797 0 0.25797 0 0.25797 0 0 0 0.25797 0 0 0 0 0 0 1.00043 0 0 0 0 1.00043 1 1 0 0 0 1 0 0 0 0 0 0 1 1 0 0 1 0 0 0 0 0 0 0

1 40.00 3.79616 123.62 0.25797 40.00 0.0352 40.02 5.08933 1 40.00 1.25841 75.00 0.02433 169.99 0.0035 169.99 2.8031 40.00 5.08934

20.00 0.13 21.00 1.01

C7 S10 A17 A9 A36

L V V V L

20.00 21.00 7.91 7.91 21.00

Diff. C4 B9 A20

1 40.00 20.00 3.79616 123.62 0.13 0.0352 40.02 1.01 4.83136 e e 1.00043 1.00043 1 0.02433 0.0035 2.8031 5.83179 132.22 10.34 40.00 21.00 40.00 20.00 169.99 7.91 169.99 7.91 40.00 21.00 e e

S0 S1 C7 A17 A9 A36

L L L V V L

1.00043 0 0 0 0 0 1.00043 1.00043

Ot-HyS net reaction (Outow-Inow) Diff.

0.00043 1.00043 1

The remaining amount of the combustion gas consists of 79.5% sulfur dioxide and 20.5% oxygen and is nally fed to the bottom stage of the SO2 absorber. After being separated from the sulfur dioxide in the SO2 absorber, the oxygen is dried and then recycled to the sulfur combustion process again. Together with the recovered anolyte and water, dissolved sulfur dioxide from the combustion gas makes the fresh anolyte. It is supplied to the SDE to produce hydrogen.

4.
4.1.

Results and discussion


Energy requirement and efciency

The specications on the SCHRS, including the Rankine cycle, are presented in Table 3. In the HXC-RB inside the sulfur furnace, 60.1 MWth/kmol-H2 heat of sulfur combustion is recovered for the heat duty of the vacuum distillation

columns reboiler. After the fresh SO2 compressor, in the HXC02 outside the sulfur furnace, 51.2 MWth/kmol-H2 heat of sulfur combustion is recovered for the heat duty of the second preheater. Of the 215.7 MWth/kmol-H2 heat of sulfur combustion supplied from the sulfur furnace through steam generators RSG-01and RSG-02 to the Rankine cycle, 37.3% is converted to the work which is produced in the high and low pressure turbines (RTB-HP, RTB-LP) minus the work required for the condensate/feed pump (RPP). Under the assumption that there is no conversion loss from work to electricity, this corresponds to 80.5 MWe/kmol-H2 of electric power and is provided to the main ow sheet. The rest of the heat supplied to the Rankine cycle, 135.2 MWth/kmol-H2, is thus rejected from the process by means of process cooling water. There are a total of 22 heat exchanger blocks in the whole ow sheet. Among them, 2 are steam generators for the Rankine cycle and 1 is a water cooler used for the Rankine cycle condenser. There are also another 11 water coolers using

12262

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

Table 3 e Thermal-to-electric conversion efciency of the Rankine cycle in the SCHRS. Steam Rankine cycle -PCS (Power Conversion System)
Heat input to Rankine cycle RSG-01 RSG-02 Work output from Rankine cycle RTB-HP RTB-LP Work input to Rankine cycle (RPP) Net work output from Rankine cycle Heat rejected from Rankine cycle (RCD) Thermal-to-electric conversion efciency

Heat or work [MWth] or [MWe]


215.7 28.3 187.4 81.8 21.0 60.8 1.2 80.5 135.2 37.33%

process cooling water. The remaining 8 heat exchangers are divided into hot sides and cold sides as if there are actually 3 heat exchangers. However, the enthalpy difference of the each stream might not be directly used for the heat transfer. Instead, the hot side streams might be cooled with water, which might then be used to heat the cold side streams as an intermediate heat transfer medium. The rst preheater, HX-01, consists of HXC-01A, HXC-01B, HXC-01C, which are its hot sides being cooled and HXH-01 which is its cold side being heated. As in the case of HX-01, the second preheater, HX-02, and the reboiler of the vacuum distillation column also consist of HXC-02, HXH-02 and HXC-RB, HXH-RB, respectively. Process heat duties and cooling loads are detailed in Table 4. All the heat duties are required only for the acid concentration process and correspond to 190.7 MWth/kmolH2. The heat duty for the rst and second preheaters (HX-01, HX-02) accounts for 68.5% (130.6 MWth/kmol-H2) of the heat duties, while that of the vacuum distillation columns reboiler accounts for the remaining 31.5% (60.1 MWth/kmol-H2). Only 41.7% (79.5 MWth/kmol-H2) for HX-01 comes from process internal heat recuperation of the main ow sheet while 58.3% (111.2 MWth/kmol-H2) for HX-02 and the reboiler comes from the heat of the sulfur combustion. There are a total of 12 water coolers using process cooling water in the whole ow sheet except for a condenser for the Rankine cycle, as mentioned above. The amount of waste heat rejected by cooling water is 194.4 MWth/kmol-H2. The cooling load for the condenser of vacuum distillation column accounts for 86.4% (167.9 MWth/kmol-H2). If the cooling load for the Rankine cycle condenser (135.2 MWth/kmol-H2) is considered, then the total amount of waste heat becomes 329.6 MWth/kmol-H2. In this case, the cooling load for the vacuum distillation columns condenser accounts for 50.9%, while that for the Rankine cycle condenser accounts for 41.0% in total. Electric power requirements are listed in Table 6. A total of 142.4 MWe/kmol-H2 is required in the process. 81.3% (115.8 MWe/kmol-H2) of it is consumed by SDE and 16.7% (23.8 MWe/kmol-H2) by fresh SO2 compressor COeSO2. The approximate remainder of about 2.0% comes from SO2 recycle compressor CO-01 and other pumps. It should be noted that the actual pumping power will be higher due to frictional losses ignored during the simulation. Of the total 142.4 MWe/

kmol-H2 electricity required for the process, 56.5% (80.5 MWe/ kmol-H2) is supplied by the Rankine cycle in SCHRS while only 61.9 MWe/kmol-H2, or about 43.5%, is supplied by nuclear energy. If a thermal-to-electric conversion efciency of the nuclear power plant of 33.3% is assumed, the electric power supplied by nuclear energy corresponds to a heat input of 185.7 MWth/kmol-H2. Considering 215.7 MWth/kmol-H2 of sulfur combustion heat recovered by the Rankine cycle, the total equivalent heat requirement corresponding to the total electric power requirement is 401.4 MWth/kmol-H2. A small quantity of steam is also required for the two-stage vacuum ejector. Stream A17 for the rst-stage steam ejector and stream A9 for the second-stage steam ejector carry a small amount of steam under the one-to-one molar entrainment ratio. The total amount of steam is about 0.02783 kmol/s, which corresponds to a 1.31 MWth/kmol-H2 heat duty. The heat duty for the steam was estimated by the enthalpy difference between the boiler feed water at 40  C and the 7.91 bar steam, as shown in Table 5. Estimated heat for this steam generation is less than one-percent. Therefore, it hardly affects the efciency estimation. As shown in the Table 7, the total heat and equivalent heat supplied to the Ot-HyS by the SCHRS as well as by the nuclear power plant is 513.9 MWth/kmol-H2. Therefore, the net thermal efciency is 47.1% based on the lower heating value (LHV) of hydrogen, at 242 MWth/kmol-H2 and 55.7% based on the higher heating value (HHV) of hydrogen, at 286 MWth/ kmol-H2 with no consideration given to the work for the air separation. Schematic conguration and major simulation results of the Ot-HyS process are shown in Fig. 4.

Table 4 e Required process heat in the Ot-HyS process. Block ID Heat Temperature [ C] Heat exchanged with : [MWth] Inlet Outlet
3.6 1.5 0.3 1.4 2.1 8.2 0.1 2.0 0.2 17.7 15.5 46.2 51.2 60.1 79.5 51.2 60.1 167.9 7.1 83.87 80.18 94.01 93.41 137.99 137.76 143.76 76.96 42.44 100.00 84.12 100.00 280.10 1079.82 66.92 79.72 e e 100.00 40.00 40.00 40.00 40.00 40.00 40.00 40.00 40.00 40.00 76.96 78.77 80.00 100.00 896.29 79.72 94.41 123.62 38.66 75.00 Cooling Water Cooling Water Cooling Water Cooling Water Cooling Water Cooling Water Cooling Water Cooling Water Cooling Water HXH-01 HXH-01 HXH-01 HXH-02 HXC-RB HXC-01A, B, C HXC-02 HXC-RB Cooling Water Cooling Water

CL-01 CL-02 CL-03 CL-04 CO-01/Stage 1 Cooler CO-01/Stage 2 Cooler CO-01/Stage 3 Cooler DR-01 DR-02 HXC-01A HXC-01B HXC-01C HXC-02 HXC-RB HXH-01 HXH-02 HXH-RB (TO-01 Reboiler) CL-CD (TO-01 Condenser) CL-SO2

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

12263

Table 5 e Steam generation and feed for the vacuum ejector. Total Flow [kmol/sec] Steam Pressure [bar] Steam Water
7.91 7.91 7.91

Temperature [ C] Steam
169.99 169.99 169.99

Enthalpy [kJ/sec] Steam


832.09 5783.68 6615.77 1.31

Water
40 40 40

Water
996.395 6925.75 7922.15

1st-stage vacuum ejector 0.0035 7.91 2nd-stage vacuum ejector 0.02433 7.91 Total 0.02783 7.91 Enthalpy difference between steam and boiler feed water, MWth/kmol-H2

4.2.

Discussion

Using Aspen Plus with appropriate thermodynamic models, simulation of the Ot-HyS process successfully demonstrates that it is an energy-efcient hydrogen production process. The process produces 1 kmol/s of hydrogen and 75 wt% dilute sulfuric acid containing 1 kmol/s of sulfuric acid while consuming 1 kmol/s of sulfur and oxygen with 3.83 kmol/s of water. It also consumes 513.9 MWth/kmol-H2 of the total heat and equivalent heat supplied by the SCHRS (including the Rankine cycle) as well as by the nuclear power plant with its thermal-to-electric conversion efciency of 33.3%, resulting in a process net thermal efciency of 47.1% (based on LHV) and 55.7% (based on HHV). If a nuclear power plant provides electricity with a higher thermal-to-electric conversion efciency of about 42.0%, as can a PBMR, then the process net thermal efciency would be 50.9% (based on LHV) and 60.1% (based on HHV). To investigate how much more efcient the Ot-HyS process is, conventional water electrolysis coupled to nuclear power plant and SRNLs PBMR/HyS cycle were chosen as the benchmarks, because the Ot-HyS process is based on water electrolysis technology and the HyS cycle. Alkaline electrolysis and PEM electrolysis can produce hydrogen at high pressure using 387 MWe/kmol-H2 (4.8 kWh/Nm3) [24] and 355 MWe/kmol-H2 (4.4 kWh/Nm3) [25]. Combining these with a conventional nuclear power plant with a thermal-to-electric conversion efciency of 33.3% brings 24.6% and 26.9% of net thermal efciencies based on HHV. Their total equivalent heat requirements are 1161.1 MWth/kmol-H2 and 1065.1 MWth/ kmol-H2, respectively. If they are combined with a PBMR

nuclear power plant that has a thermal-to-electric conversion efciency of 42.0%, then their HHV basis net thermal efciencies are 31.0% and 33.8%, with total equivalent heat requirements of 921.4 MWth/kmol-H2 and 845.2 MWth/kmolH2, respectively. Based on SRNLs work, its PBMR/HyS cycle has HHV basis net thermal efciencies of 40.6%, excluding the power needed for the helium circulators. That of the reference 500-MWth PBMR/HyS cycle, including the electric power requirements of the helium circulator, NHSS, PGS and BOP, is even lower (36.7%) Net thermal efciency benchmarking results are summarized in Table 8. In conclusion, the Ot-HyS process is a signicantly energy-efcient hydrogen production process. For a simple comparison, based on HHV and a nuclear power plant with a thermal-to-electric conversion efciency of 42.0%, the net thermal efciency of the Ot-HyS process is about 25.0e30.0% higher than that of conventional water electrolysis and about 20.0% higher than that of SRNLs PBMR/HyS cycle. Although we already knew that the Ot-HyS process is more

Table 7 e Net thermal efciency of the Ot-HyS process. Once-through hybrid sulfur process
Total cooling load Recuperative heat Net cooling load Supplied by sulfur combustion Supplied by cooling water Total heat requirement Recuperative heat Net heat requirement Supplied by sulfur combustion Supplied by nuclear energy Heat load for vacuum ejector steam generation Total electricity requirement Supplied by sulfur combustion Efciency Equivalent heat requirement Supplied by nuclear energy Efciency Equivalent heat requirement Total equivalent heat requirement LHV of H2 HHV of H2 Net thermal efciency (based on LHV of H2) Net thermal efciency (based on HHV of H2)

Heat or work [MWth] or [MWe]


385.1 79.5 305.6 111.2 194.4 190.7 79.5 111.2 111.2 0.0 1.3 142.4 80.5 37.33% 215.7 61.9 33.33% 185.7 401.4 242 286 47.09% 55.65%

Table 6 e Required process electricity in the Ot-HyS process. Block ID


PEM SDE CO-SO2 CO-01/Stage 1 CO-01/Stage 2 CO-01/Stage 3 PP-01 PP-02 PP-03 PP-04 PP-05 PP-06 PP-07 Total

Work [MWe]
115.8 23.8 1.3 1.2 1.5.E-02 2.2.E-02 0.2 2.6.E-05 4.0.E-03 8.6.E-03 1.9.E-03 2.6.E-02 142.4

12264

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

Fig. 4 e Schematic conguration of the Ot-HyS process with its major simulation results.

energy-efcient than benchmark processes because of its characteristics, the focus of this benchmarking is on the quantication of how much more efcient the Ot-HyS process is. Hydrogen produced from off-peak electricity by the Ot-HyS process can be stored and reconverted into electricity again during peak-load time by a fuel cell. Using heat from the sulfur combustion that includes 80.5 MWe/kmol-H2 of electric energy, 61.9 MWe/kmol-H2 of nuclear electricity generation is rst converted into 1 kmol/s of hydrogen production as energy storage. Then, based on the LHV of hydrogen, this 242 MWth/ kmol-H2 of heat is reconverted to electricity corresponding to 72.6 MWe/kmol-H2, 121 MWe/kmol-H2 and 169.4 MWe/kmolH2 under fuel cell efciencies of 30.0%, 50.0% and 70.0%, respectively (85.8 MWe/kmol-H2, 143.0 MWe/kmol-H2 and 200.2 MWe/kmol-H2, based on HHV). Nuclear power plants, which have high capital cost, and low operating cost are generally operated continuously generating the base load because it is the most economic mode. However, it is also technically the simplest way, since nuclear power plants cannot alter power output as readily as other power plants. Although some plants being built today have load-following capacity, this is still a complex problem. With hydrogen

produced through the Ot-HyS process, nuclear energy, which is a sustainable development technology, can be applied to a greater range of applications. If 69,000 thousand metric tons of all the sulfur annually produced in all forms worldwide were supplied to the Ot-HyS process, about 4313 thousand metric tons of hydrogen could be produced annually from the Ot-HyS process based on one-to-one molar ratio. If electricity were produced from the stored hydrogen 24 h a day, 365 days a year, the produced electricity would be equivalent to 11.6 GWe (LHV basis, 70% of fuel cell efciency is assumed). If the stored hydrogen were used just for 3 h of peak-load time, the produced electricity would be equivalent to 92.7 GWe. The electricity produced in 2007 was about 16.42 trillion kWh. This is equivalent to 1900.5 GWe. Considering 92.7 GWe of supply from stored hydrogen, 4.9% of electricity can be supplied from hydrogen reservoir. The effect of the thermal-to-electric conversion efciency of the PCS in SCHRS on net thermal efciency is shown through graphs in Fig. 5. Obviously, the net thermal efciency increases as the thermal-to-electric conversion efciency of the PCS in SCHRS increases. If it increases up to 42% using more efcient PCS in SCHRS, and if that of the nuclear power plant is also 42%, then the net thermal efciency of the Ot-HyS process would be

Table 8 e Net thermal efciency benchmarking. Benchmark (based on 1 kmol/s-H2) Ot-HyS SRNLs PBMR/HyS [1,2] Conventional water electrolysis Alkaline electrolysis [24]
Heat requirement [MWth] Electricity requirement [Mwe] Thermal-to-electric conversion efciency of NPP Equivalent heat requirement [MWth] Total heat requirement [MWth] Net thermal efciency (LHV basis) Net thermal efciency (HHV basis) 112.5 142.4 33.33% 401.4 513.9 47.09% 55.65% 417.1 120.9 42.00% 287.9 705.0 34.33% 40.57% 0.0 387.0 33.33% 1161.1 1161.1 20.84% 24.63%

PEM electrolysis [25]


0.0 355.0 33.33% 1065.1 1065.1 22.72% 26.85%

42.00% 363.0 475.6 50.88% 60.14%

42.00% 921.4 921.4 26.26% 31.04%

42.00% 845.2 845.2 28.63% 33.84%

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

12265

Fig. 5 e The effect of the PCS efciency on the net thermal efciency of the Ot-HyS process.

53.7% (LHV basis) and 63.5% (HHV basis). Because of the great amount of high-temperature excess heat from the sulfur combustion process, there is much room for the use of more efcient PCS such as the supercritical CO2 Brayton cycle rather than the simple Rankine steam cycle used in this work. In the absence of a detailed model for its phenomenon, SDE has been treated as a black box using the development performance targets of SRNL. The SDE is assumed to operate at a xed cell potential of 600 mV versus current density of 500 mA/cm2, with a xed SO2 conversion of 40%, and at a xed 50 wt% H2SO4 product. Among these, the cell potential is directly related to the net thermal efciency of the Ot-HyS process. Electric energy required for the SDE, which is the only

Fig. 6 e The effect of the SDE electricity requirement on the net thermal efciency of the Ot-HyS process.

remaining technical challenge of the Ot-HyS process, accounts for the greatest portion of the total required energy. In other words, it has the most dominant effect on the net thermal efciency. Based on the target energy demand of 150% for the SDE, the net thermal efciency will be signicantly reduced to 35.2% (LHV basis) and 41.6% (HHV basis), as shown in Fig. 6. In the past year, the SDE research project was focused on demonstration of the SDE operation without sulfur layer build-up inside of the MEA, which is a phenomenon that reduces performance as well as limits operating lifetime. In the past year, SRNL has demonstrated successful long time operation of the single cell SDE without sulfur build-up limitations using a new operating method. The test using the new sulfur-limiting operating method was conducted with a Naon membrane, which has a 54 cm2 active cell area, at 80  C and 172 kPa of the cell operating conditions. The cell potential was stable at approximately 760 mV, which is beyond the development performance target of 600 mV, while the current density was maintained at the development performance target of 500 mA/cm2. However, if this method is applied for more advanced membranes under higher temperature and pressure in the future conguration, the target cell potential of 600 mV can be achieved. Although a detailed heat transfer analysis has not yet been conducted, temperature differences between the hot sides and cold sides were set high enough to assume that all heat transfers can be achieved as designed for the purpose of the process net thermal efciency estimation. An additional pumping power will be required to compensate the ignored pressure drops through the process. Additional energy will also be required to pump process cooling water that allows a total of 329.6 MWth/kmol-H2 waste heat rejection to and from the cooling tower. The amount of additionally required electric energy also will not be signicant and will cause no signicant

12266

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

loss of net thermal efciency. However, its ability to cool down the process stream to about 40  C should be available yearround by maintaining the temperature of process cooling water at 30  C or lower. Otherwise, some of the coolers, like the partial condenser of the vacuum distillation column, would need to be refrigerated condensers. In the SRNLs PBMR/HyS report, hydrogen price from the PBMR/HyS plant was calculated as $6.18/kg H2 based on Shaws estimation and $5.34/kg H2 based on SRNL/Icaruss estimation [2]. In the CEAs recent work, hydrogen price from its HyS cycle conguration was calculated as V6.6/kg H2 [26]. Although no detailed economic feasibility study was conducted here, economics of the Ot-HyS plant has some positive and negative economic potentials that can be easily found when compared with that of the PBMR/HyS plant. In terms of O&M cost, there is no need for nuclear fuels, but the Ot-HyS plant needs sulfur as well as about 3.8 times more water as the process feed, than the PBMR/HyS plant does. The Ot-HyS plant lost the oxygen byproduct credit, but now can take 75 wt% sulfuric acid by-product credit, instead. In addition to that, the Ot-HyS plant needs less net electric power from grid than the PBMR/HyS plant does. From the view point of the capital cost, Ot-HyS plant could save costs for the SAD and PBMR systems which are more expensive than the sulfur combustion system.

efciency, the net thermal efciency would be 53.7% (LHV basis) and 63.5% (HHV basis). The progress of development for the SDE cell potential has a dominant effect on the net thermal efciency. The SDE is the only remaining technical challenge of the Ot-HyS process. In order for the Ot-HyS process to maintain its competitiveness, the development performance targets of SRNL should be achieved as anticipated. Hydrogen produced through the energy-efcient Ot-HyS process would be used as the off-peak electric energy storage, to relieve the burden of load-following and would help to expand applications of nuclear energy. Because of its competitive advantages, such as a higher net thermal efciency, less technical challenge and favorable sulfur statistics, the Ot-HyS process could play an important role in securing a bridge to the sustainable energy future during the short-term transitional period. Although it was demonstrated that the Ot-HyS process is an energy-efcient hydrogen production process, it is the unit cost of H2 production ($/kg-H2) that will ultimately determine whether the Ot-HyS process is feasible. If hydrogen produced by the OtHyS process is also economically favorable, then the competitive edge of the Ot-HyS process could be further secured. A detailed economic feasibility study will be the main topic of our future work.

5.

Conclusion Acknowledgement
This research was supported by the WCU (World Class University) program through the National Research Foundation of Korea funded by the Ministry of Education, Science and Technology (R33-2008-000-10047-0).

This study was conducted for the purpose of demonstrating that the proposed Once-through Hybrid Sulfur (Ot-HyS) process, which is competitive due to the fact that it is based on technology of the SDE of the Hybrid Sulfur cycle (HyS) and on the well-established sulfur combustion process, is an energyefcient nuclear hydrogen production process that can help to resolve the major energy challenges we are facing. In order to demonstrate that, a ow sheet for the sulfur combustion process, including its heat recovery system, was developed by referring to existing facilities. By combining this with the main ow sheet developed for the HyS cycle by SRNL with small modications, detailed ow sheet for the Ot-HyS process was completed. After that, the ow sheet developed for the Ot-HyS process was simulated using Aspen Plus with appropriate thermodynamic models. Finally, using the simulation result, the net thermal efciency of the Ot-HyS process was benchmarked with those of others and discussed. It was found that the net thermal efciency was 47.1% (based on LHV) and 55.7% (based on HHV) assuming 33.3% thermal-to-electric conversion efciency of the nuclear power plant, and also 50.9% (based on LHV) and 60.1% (based on HHV) assuming 42.0% thermal-to-electric conversion efciency of the nuclear power plant. The Rankine cycle in the SCHRS (Sulfur Combustion Heat Recovery System) was applied with 37.3% thermal-to-electric conversion efciency. This efciency was also signicantly higher than those of the benchmarking hydrogen production processes such as nuclear-powered conventional water electrolysis and SRNLs PBMR/HyS cycle. There is much room for the use of a more efcient PCS (Power Conversion System). If both the nuclear power plant and PCS in the SCHRS have 42% thermal-to-electric conversion

references

[1] Gorensek MB, Summers WA. Hybrid sulfur owsheets using PEM electrolysis and a bayonet decomposition reactor. Int J Hydrogen Energy 2009;34(9):4097e114. [2] Gorensek MB, Summers WA, Bolthrunis CO, Lahoda EJ, Allen DT, Greyvenstein R. Hybrid sulfur process reference design and cost analysis. Savannah River National Laboratory, http://dx.doi.org/10.2172/956960; June 12, 2009. Report no. SRNL-L1200e2008e00002. [3] Brecher LE, Wu CK. Electrolytic decomposition of water. Westinghouse Electric Corp. US Patent No. 3888750; 1975. [4] Farbman GH. The conceptual design of an integrated nuclear-hydrogen production plant using the sulfur cycle water decomposition system. NASA contractor report, NASA-CR-134976; 1976. [5] Lu PWT, Ammon RL, Parker GH. A study on the electrolysis of sulfur dioxide and water for the sulfur cycle hydrogen production process. NASA contractor report, NASA-CR163517; 1980. [6] Parker GH. Solar thermal hydrogen production process. Final report from Westinghouse Electric Corp to US DOE, DOE/ET/ 20608-1; 1983. [7] Brecher LE, Spewock S, Warde CJ. The Westinghouse sulfur cycle for the thermochemical decomposition of water. Int J Hydrogen Energy 1977;2(1):7e15.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 2 2 5 5 e1 2 2 6 7

12267

[8] Lu PWT, Ammon RL. Sulfur dioxide depolarized electrolysis for hydrogen production: development status. Int J Hydrogen Energy 1982;7(7):563e75. [9] Lu PWT. Technological aspects of sulfur dioxide depolarized electrolysis for hydrogen production. Int J Hydrogen Energy 1983;8(10):773e81. [10] Carty R, Cox K, Funk J, Soliman M, Conger W, Brecher L, et al. Process sensitivity studies of the Westinghouse sulfur cycle for hydrogen generation. Int J Hydrogen Energy 1977;2(1): 17e22. [11] Farbman GH. Hydrogen production by the Westinghouse sulfur cycle process: program status. Int J Hydrogen Energy 1979;4(2):111e22. [12] Jeong YH, Kazimi MS. Optimization of the hybrid sulfur cycle for nuclear hydrogen generation. Nucl Technol 2007;159: 147e57. [13] U.S. DOE. II.H.1 sulfur-iodine thermochemical cycle, DOE Hydrogen Program FY. 2008 Annual Progress Report; 2009. [14] U.S. DOE. Sulfur-iodine thermochemical cycle, Proc. DOE Hydrogen Program. 2009 Annual Merit Review; May, 19 2009. [15] Conger WL. Open-loop thermochemical cycles for the production of hydrogen. Int J Hydrogen Energy 1979;4(6): 517e22. [16] Abdel-Aal HK. Opportunites of open-loop thermochemical cycles: a case study. Int J Hydrogen Energy 1984;9(9): 767e72. [17] Zhou Junhu, Zhang Yanwei, Wang Zhihua, Yang Weijuan, Zhou Zhijun, Liu Jianzhong, et al. Thermal efciency evaluation of open-loop SI thermochemical cycle for the

[18] [19] [20] [21] [22]

[23]

[24]

[25]

[26]

production of hydrogen, sulfuric acid and electric power. Int J Hydrogen Energy 2007;32(5):567e75. Ober JA. 2009 Mineral commodity summaries. U.S. Department Of the Interior; Jan 2009. U.S. Geological Survey. Ober JA. 2007 Minerals yearbook-SULFUR. U.S. Department Of the Interior; July 2009. U.S. Geological Survey. Louie DK. Handbook of sulphuric acid manufacturing. 2nd ed. DKL Engineering, Inc; 2008. DKL Engineering, Inc.. Sulphuric acid on the WebTM. DKL Engineering, Inc., http://www.sulphuric-acid.com/; 2008. Wang P, Anderko A, Springer RD, Young RD. Modeling phase equilibria and speciation in mixed-solvent electrolyte systems: II. Liquideliquid equilibria and properties of associating electrolyte solutions. J Mol Liquids 2006;125:37e44. Kwinana Power Station, from Wikipedia, the free encyclopedia, http://en.wikipedia.org/wiki/Kwinana_Power_ Station,_Western_Australia; 2009. Norsk Hydro Electrolysers AS. High pressure electrolysers: 10e60 Nm3/h, http://www4.hydro.com/electrolysers/en/ products/range/high_pressure_electrolyser/; Feb, 25 2008. Norsk Hydro Electrolysers AS. Product development e the shape of things to come, http://www4.hydro.com/ electrolysers/en/products/product_development/; February 25 2008. Leybrosa J, Saturnina A, Mansillab C, Gilardic T, Carles P. Plant sizing and evaluation of hydrogen production costs from advanced processes coupled to a nuclear heat source: Part II: Hybrid-sulphur cycle. Int J Hydrogen Energy 2010;35(3):1019e28.

Anda mungkin juga menyukai