Вы находитесь на странице: 1из 10

N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3

Available online at www.sciencedirect.com

ScienceDirect

journal homepage: http://www.journals.elsevier.com/nuclear-


engineering-and-technology/

Original Article

TWO-DIMENSIONAL SIMULATION OF HYDROGEN IODIDE


DECOMPOSITION REACTION USING FLUENT CODE FOR
HYDROGEN PRODUCTION USING NUCLEAR TECHNOLOGY

JUNG-SIK CHOI a, YOUNG-JOON SHIN b, KI-YOUNG LEE b, and JAE-HYUK CHOI c,*
a
The Institute of Machinery and Electronic Technology, Mokpo National Maritime University, 91 Haeyangdaehak-ro, Mokpo-si, Jeollanam-do,
South Korea
b
Korea Atomic Energy Research Institute, Daedeok-Daero 989-111, Yuseong-gu, Daejeon, South Korea
c
Division of Marine Engineering System, Korea Maritime and Ocean University, 727 Taejong-ro, Yeongdo-gu, Busan, South Korea

article info abstract

Article history: The operating characteristics of hydrogen iodide (HI) decomposition for hydrogen pro-
Received 5 October 2014 duction were investigated using the commercial computational fluid dynamics code, and
Received in revised form various factors, such as hydrogen production, heat of reaction, and temperature distribu-
5 January 2015 tion, were studied to compare device performance with that expected for device devel-
Accepted 24 January 2015 opment. Hydrogen production increased with an increase of the surface-to-volume (STV)
Available online 27 March 2015 ratio. With an increase of hydrogen production, the reaction heat increased. The internal
pressure and velocity of the HI decomposer were estimated through pressure drop and
Keywords: reducing velocity from the preheating zone. The mass of H2O was independent of the STV
Computational fluid dynamics ratio, whereas that of HI decreased with increasing STV ratio.
Fluent code Copyright 2015, Published by Elsevier Korea LLC on behalf of Korean Nuclear Society.
Hydrogen iodide decomposition
reaction
Hydrogen production
Sulfureiodine cycle

1. Introduction were first postulated by Funk and Reinstrom [2] as the most
efficient way to produce fuels (e.g., hydrogen, ammonia) from
Hydrogen not only has potential for use as an alternative to stable and abundant species (e.g., water, nitrogen) using heat
fossil fuels, but also plays a key role in solving what is known sources.
as a trilemma (economic growth, energy use, and environ-
mental degradation) [1]. A number of thermochemical cycles  2H2O SO2 I2 / H2SO4 2HI

* Corresponding author.
E-mail address: choi_jh@kmou.ac.kr (J.-H. Choi).
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://
creativecommons.org/licenses/by-nc/3.0) which permits unrestricted non-commercial use, distribution, and reproduction in any me-
dium, provided the original work is properly cited.
http://dx.doi.org/10.1016/j.net.2015.01.006
1738-5733/Copyright 2015, Published by Elsevier Korea LLC on behalf of Korean Nuclear Society.
N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3 425

 2HI / H2 I2 process [22,23]. Commercial process simulators such as Aspen


 H2SO4 / H2O SO2 1/2 O2 Plus (Aspen Technology, Inc., Massachusetts, USA) [24e26]
and ESP (simulator program, Microsoft, USA) [27e29] have
In a closed cycle system, various related processes have been used to develop these thermodynamic models.
been proposed and have received a great deal of attention. The In addition, in recent years, interest in and research on nu-
sulfureiodine (SI) process involving thermochemical clear hydrogen technology, which involves direct decomposi-
hydrogen production using nuclear energy was proposed by tion of water to produce a large amount of hydrogen at high
General Atomics and this technology has been studied by temperatures in a nuclear reactor as the heat source of the
many researchers [3e5]. thermochemical cycle, has increased [30,31]. In particular,
Extensive research on hydrogen iodide (HI) decomposition following the successful continuous operation of a bench-scale
for hydrogen production has been carried out for experimental closed cycle gas turbine by the Japanese Atomic Energy Agency,
verification and measurement of several factors, such as con- SI thermochemical hydrogen production technology has taken
version efficiency and kinetics [6]. The decomposition of HI is center stage as one of the high-practical-potential hydrogen
the key to producing hydrogen in the SI cycle. The decompo- production technologies that could be coupled with the very
sition of HI in the absence of any catalysts is not efficient even high temperature reactor (VHTR) [32e36]. However, trans-
at 773.15 K; therefore, catalysts have been used to promote this ferring the high heat output of the gas coolant for VHTR to the SI
reaction [7]. Moreover, the rate of the homogeneous gas-phase thermochemical cycle involves the great challenge of devel-
reaction is considerably low at temperatures <700 K. Thus, it is oping a heat-exchanger material capable of withstanding high
desirable to use a catalyst to increase the reaction rate, which temperatures (1173 K) and pressures (approximately 50e70 kg/
has led to the study of platinum catalysts [8e10]. A decompo- cm2). In the HI decomposition reaction, which requires a lot of
sition test using 98% pure HI gas was carried out by Ilda [8] in heat in the SI cycle, if the decomposition efficiency of HI can be
1978 using a platinum supported on polytetrafluoroethylene maintained at high pressures, the design and development of
catalyst. Moreover, Wang et al. [11] investigated the effects of the heat exchanger and reactor can be flexible.
different supports (carbon nanotubes, active carbon, carbon In this study, an optimum decomposition reactor in the
molecular sieve, graphite, and Al2O3), masses of catalyst, and thermochemical VHTReSI cycle operating characteristics of
reaction temperatures on the decomposition of HI. Fresh and the HI decomposition reaction were investigated using the
used active carbon-supported platinum catalysts were also code. Several factors, such as hydrogen production, heat of
characterized. Furthermore, different types of catalysts the reaction, and temperature distribution, were studied to
including noble metals and support materials (active carbon compare the device performance with that expected for de-
[7,12], Pt [13], Ni [14], and bimetallic catalysts [6]) have been vice development.
reported to catalyze HI decomposition.
To date, much progress has been made on the efficiency of
2. Materials and methods
this process. However, insufficient research has been con-
ducted on the role of water vapor in the HI decomposition
2.1. Basic equation for the HI decomposition reaction
process. In addition, very few reports have been published on
the kinetic studies of HI decomposition in the presence of H2O.
The decomposition reaction of HI is as follows:
Assessment of the real potential of the HI decomposition re-
action requires a deep knowledge of the thermodynamic
behavior of hydrogen production systems. Improving the ef-
2HI / I2 H2 (1)
ficiency of hydrogen production and optimizing the reactor
requires an understanding of the thermodynamics of the In this study, the simulation was performed using the
process. The thermodynamic model for the HIx system used LangmuireHinshelwood-type rate equation on the catalytic
by Roth and Knoche was proposed by Neumann [15] in 1987, decomposition of HI in the presence of an active carbon
and was based mostly on total pressure measurements per- catalyst proposed by Oosawa et al. [13]
formed at Rheinisch-Westfa lische Technische Hochschule
(RWTH) Aachen [16] and on liquideliquid equilibrium data. In rHI kHI P RHI (2)
the simulation studies, hydrogen pressures higher than those
 
expected for homogeneous gas-phase HI decomposition in the p K PFe
xH2 xI2 1 I22
direct decomposition of HIeH2OeI2 solution were achieved xHI
RHI    (3)
[17]. Berndha user et al. [18] compared the calculated homo- 1 KI2 PxI2 Kp 1 KI2 PxI2
geneous gas-phase reaction results with those of direct HI
where
decomposition from HIeH2OeI2 solutions, and found that the
 
rate of direct HI decomposition was several orders of magni- 8; 250 cal$mol1
KHI 0:158 exp (4)
tude higher than that of the gas-phase reaction. Furthermore, RT
Lanchi et al. [19], Murphy and O'Connell [20], and Hadj-Kali
 
et al. [21], reviewed the phase equilibrium properties of the 16; 480:1 cal$mol1
KI2 5:086  1011 exp (5)
HIeI2eH2O (HIx) system and proposed new thermodynamic RT
models for describing the thermodynamic properties of this
system. Many researchers have attempted to develop appro- where Fe is the equilibrium conversion, rHI (mol/m3/s) is the
priate models for the HI decomposition reaction in the SI reaction rate, kHI (mol/m3/Pa/s) is the reaction rate coefficient,
426 N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3

P (Pa) is the system pressure, xHI, xI2, and xH2 are the mole line in Fig. 1B shows the multistage distillation column for the
fractions of HI, I2, and H2, respectively, KI2 is the absorption periodic acid solution and the HI decomposer.
equilibrium constant of iodine (1/Pa), and Kp is the equilibrium The HIeI2eH2O mixture gas was discharged from the
constant for the decomposition of HI. Further, Kp can be ob- condenser part of the multistage distillation column for the HI
tained from HSC 5.1 (chemistry-software, USA) [37] as follows: reaction, and flowed into the HI decomposer. Then, the H2eI2
 0:5  0:5 gas was generated after decomposition of the mixture gas. In
CeH2 CeI2 this study, the characteristics of the thermal decomposition
Kp (6) device (HI decomposer, Fig. 2A) were investigated using the
CeHI
CFD simulator model (Fig. 2B).
q
 e  e  Fig. 2A shows the HI decomposer model and the heating
Ce CH2 CI2
F 1  HI 1 
e
(7) device consisting of a sealed three-stage electric heater. A
CHI;in Kp CHI;in Hastelloy HC276 tube was inserted, and its outer diameter,
inner diameter, and height were 60.5 mm, 52.7 mm, and
Kp 4:515  103  4:519  105 T 1:437  107 T2 1500 mm, respectively. This tube was filled with active carbon
 8:662  1011 T3 1:828  1014 T4 (8) catalyst and Al2O3 Raschig ring. Each of the three electric
heaters had built-in heating elements (i.e., 3.65-kW Kanthal A-
where CeH2 (mol/m3), CeI2 (mol/m3), and CeHI (mol/m3) are the 1) and the HC276 tube was surrounded by a low-density
equilibrium concentrations of H2, I2, and HI, respectively, CHI,in ceramic fiber board for insulation. Moreover, a proportional-
(mol/m3) is the inlet concentration of HI, and T (K) is the re- integral-derivative controller was used to maintain the tem-
action temperature. perature of the three electric heaters at a set value. The
HIeI2eH2O mixture gas discharged from the condenser of the
multistage distillation flowed into the bottom of the inlet of
2.2. Model design the HI decomposer at an absolute pressure of 5 kgf/cm2
(506,625 Pa). The temperature gradually increased to reach the
The reactive distillation flow sheet developed by GA Company set temperature of the electric heater of the HI decomposer
(General Atomics, San Diego, CA, USA) [38] is shown in Fig. 1. and was maintained at this value. Meanwhile, helium (He) gas
Fig. 1A illustrates the HI decomposition section. The dashed or N2 gas was supplied to the HI decomposer to maintain the

Fig. 1 e Reactive distillation flow sheet in the SI cycle. (A) HI decomposition section. (B) Multistage distillation column and HI
decomposer. DI, deionized; HI, hydrogen iodide; SI, sulfureiodine.
N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3 427

Fig. 2 e Designed HI decomposer for the SI cycle and the CFD model. (A) Designed HI decomposer. (B) CFD domain.
CFD, computational fluid dynamics; HI, hydrogen iodide; ID, internal diameter; OD, outer diameter; SI, sulfureiodine.

pressure at 506,625 Pa, and thermal equilibrium was achieved size, and amount of catalyst. The effect of the surface-to-
using an external heating chamber of the HI decomposer. The volume (STV) ratio on hydrogen production was also consid-
mixture gas was supplied to the inlet under the conditions ered. When the reactor was developed, the STV ratio, defined
that the temperature reached the set point and the pressure as the specific surface of the catalyst, was approximately 0.4 in
was reliably maintained at 5 kgf/cm2. the preliminary simulation. The effects of setting the STV
ratio at 0.3, 0.4, and 0.5 on hydrogen production and reaction
heat were investigated.
2.3. CFD analysis methods and boundary conditions of
the CFD domain
STV ratio A/V
The CFD simulation model of the HI decomposer is presented
in Fig. 2B. Size, properties, and boundary conditions identical where A is the surface area of the pore walls and V is the
to those of the actual operating system were applied, and then unit volume.
the CFD analysis was conducted using the Fluent program. For efficient CFD analysis, the following criteria were
Moreover, to describe the actual operating system, a user- assumed:
defined function was applied. At t 0 and 464.85 K
(191.7  C), the HIeI2eH2O mixture gas with flow rates of  The temperature of the heater was always maintained at
0.008 mol/s HI, 0.0002 mol/s I2, and 0.020 mol/s H2O was the set temperature.
injected into the HI decomposer at an absolute pressure of  N2 gas was fully filled inside the HI decomposer, as the
5 kgf/cm2. Maintaining a fixed porosity of 0.4, the tempera- initial condition of the simulation.
tures of the electric heaters were set at 773.15 K (500 C),  The backward (reverse) reaction was not considered.
873.15 K (600 C), and 973.15 K (700 C). The porosity value was  The absolute pressure and temperature were set at 5 kgf/
approximately predicted by considering several parameters, cm2 (506,625 Pa) and 673.15 K (400 C), respectively, and
such as the reactor size, volume, catalyst density, catalyst then the mixture gas was injected.
428 N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3

Fig. 3 e Estimates of hydrogen production according to changes in heater temperature. (A) Heater set value 773.15 K. (B)
Heater set value 873.15 K. (C) Heater set value 973.15 K. STV, surface-to-volume ratio.

3. Results and discussion was produced at 1.11e1.12 mol/h, 1.98e1.99 mol/h, and
3.09e3.10 mol/h at STV ratios of 0.3, 0.4, and 0.5, respectively.
3.1. Effect of the heater temperature on hydrogen As shown in Fig. 3AeC, hydrogen production was more
production strongly affected by the STV ratio than by the heat supplied by
the heater. The inner temperature of the HI decomposer was
To investigate the effects of the heater temperature on higher than the temperature for catalyst activation
hydrogen production, hydrogen production at HI decomposer (T > 623.15 K) initially, and the inlet mass flow rate was the
temperatures of 773.15 K, 873.15 K, and 973.15 K were esti- same regardless of the heater temperature.
mated at STV ratios of 0.3, 0.4, and 0.5, as shown in Fig. 3.
Under each condition, the amount of hydrogen produced was 3.2. Reaction heat prediction with different STV ratios
monitored at 10-minute intervals and the total operation time
was 180 minutes. Fig. 4AeC presents the heat of the reaction at STV ratios 0.3,
Fig. 3A shows the hydrogen production measured at 0.4, and 0.5. The reaction was endothermic. As shown in
773.15 K (heater temperature). Up to 20 minutes after the Fig. 4A, at an STV ratio of 0.3, the reaction heat was measured
operation of the HI decomposer (i.e., after the start of the at different heater temperatures. A similar consumption of
decomposition process), no hydrogen production occurred reaction heat was observed regardless of the heater temper-
regardless of the STV ratio. From 20 minutes to 120 minutes, ature. After 60 minutes, the reaction heat required to produce
the amount of hydrogen production varied according to the hydrogen was 0.0827e0.0831 kW/h, whereas after 120 mi-
STV ratio: the hydrogen production was 1.11e1.12 mol/h, nutes, when the hydrogen production was constant, it was
1.98e1.99 mol/h, and 3.09e3.10 mol/h at the STV ratios of 0.3, 0.1116e0.117 kW/h. Therefore, the reaction heat increased
0.4, and 0.5, respectively. As presented in Fig. 3B, at 873.15 K, with the increase of hydrogen production regardless of the
the amount of hydrogen produced was very low for 20 mi- temperature of the heater. Fig. 4B, C shows similar results at
nutes, but after 120 minutes, it increased according to the STV STV ratios of 0.4 and 0.5, respectively. At an STV ratio of 0.4,
ratio. After 120 minutes, hydrogen production was the reaction heat was 0e0.1604 kW/h regardless of the heater
1.11e1.12 mol/h, 1.98e1.99 mol/h, and 3.09e3.10 mol/h at STV temperature; however, it increased with increasing hydrogen
ratios of 0.3, 0.4, and 0.5, respectively. Fig. 3C shows hydrogen production. At an STV ratio of 0.5, the reaction heat was
production at 973.15 K. Similar to the reaction at 773.15 K and 0e0.2147 kW/h and showed a similar correlation with
873.15 K, the chemical reaction occurred, and then, hydrogen hydrogen production as that observed at the 0.4 STV ratio.

Fig. 4 e Reaction heat according to different STV ratios. (A) STV ratio 0.3. (B) STV ratio 0.4. (C) STV ratio 0.5. STV, surface-to-
volume ratio.
N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3 429

Fig. 5 e Inner temperature of HI decomposer according to temperature changes of heater. (A) Heater set value 773.15 K
with different STV ratios at 160 minutes. (B) STV ratio of 0.3 with different temperatures at 160 minutes. (C) Heater set
value 773.15 K with different STV ratios at 160 minutes. (D) STV ratio of 0.3 with different temperatures at 160 minutes.
HI, hydrogen iodide; ITC, inner thermocouple; STV, surface-to-volume ratio.
430 N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3

Fig. 5 e (continued).

As shown in Fig. 4AeC, the reaction heat used in the HI ITC 2, 777e794 K at ITC 3, 790e805 K at ITC 4, and 789e820 K at
decomposition reaction for hydrogen production was corre- ITC 5, regardless of the STV ratio.
lated to the specific surface area of the catalyst and it had a Especially, at an STV ratio of 0.3, the temperatures of ITC
direct relation with the production of hydrogen. However, the 1, ITC 2, and ITC 3 at 973.13 K were higher than the tem-
temperature changes of the HI decomposer had no direct ef- peratures at 773.15 K and 873.15 K for the same STV ratio
fect. The mixture gas was introduced into the HI decomposer (Fig. 5B). Because the decomposition is an endothermic re-
when the heat supplied from the heater passed through the action, it seems that less reaction heat was supplied, and
decomposer completely and the temperature was higher than therefore the inner temperature was higher at 973.15 K than
the activation temperature (i.e., 673.15 K). Moreover, the in the other cases. Using different STV ratios, there was no
temperature of the mixture gas from the preheating zone was significant difference in all of the temperatures. However,
increased up to the temperature needed for the catalyst different trends were exhibited at various temperatures
decomposition reaction. Therefore, to enhance thermal effi- even if the differences were small. At 973.15 K, the inner
ciency, lowering the temperature of the heater was more ad- temperature was found to be higher than in the other cases.
vantageous at the STV ratio. The main reaction takes place well with the reaction tem-
perature or more operation conditions, so it is not necessary
to supply the over temperature.
3.3. Dependence of the HI decomposer inner temperature
distribution on the heater temperature
3.4. Inner velocity and pressure of HI decomposer
The inner temperature distribution of the HI decomposer was according to temperature change of heater
monitored at different temperatures (773.15 K, 873.15 K, and
973.15 K) at various STV ratios (0.3, 0.4, and 0.5) during an In Fig. 6AeC, the static pressure and velocity magnitude at
operation time of 160 minutes (Fig. 5AeC). Moreover, the STV ratios of 0.3e0.5 after 160 minutes of operation are pre-
temperatures are presented in the form of area-weighted av- sented at three different heating temperatures. The static
erages at inner thermocouple (ITC) 1e5. The temperature at pressure and velocity magnitude are area weighted and
the center of the catalyst zone was lower than that at the averaged. As shown for positions AeE, the static pressure was
preheating zone for all STV ratios (Fig. 5A). The heat was used 2.71 Pa at position A (0 mm), 26.95 Pa at position B (10 mm),
to catalyze the decomposition reaction. The temperatures of 18.66 Pa at position C (510 mm), 2.41 Pa at position D
the outlet part, which discharged the mixture gas, and the (1490 mm), and 2.25 Pa at position E (1500 mm). The mixture
temperature of the zone above the catalyst zone were rela- gas flowed to the preheating zone filled with the Al2O3 Raschig
tively higher than that of the catalyst zone. The temperature ring and it collided with the solid Al2O3 Raschig ring, thus
distribution was as follows: 778e801 K at ITC 1, 779e792 K at generating high pressure. Therefore, a pressure drop was
N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3 431

Fig. 6 e Pressure and velocity of HI decomposer according to temperature changes of the heater. (A) STV ratio 0.3, (B) STV
ratio 0.4. (C) STV ratio 0.5. Heater set at 773.15 K, 873.15 K, and 973.15 K. HI, hydrogen iodide; STV, surface-to-volume ratio.

evident from the increase of the pressure value. When the As shown for AeE, the velocity magnitude also had no cor-
mixture gas from the preheating zone flowed to the catalyst relation with the heater temperature and STV ratio. The ve-
zone, a pressure drop was generated but the magnitude was locity was 0.016 m/s at position A (0 mm), 0.04 m/s at position B
smaller than in the previous case. (10 mm), 0.016 m/s at position C (510 mm), 0.016 m/s at position
432 N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3

D (1490 mm), and 0.14 m/s at position E (1500 mm). At the outlet references
at position E, the velocity of the mixture gas was highest
(0.14 m/s). In the empty space (i.e., AeB), the velocity of the
mixture gas was 0.016e0.04 m/s, but it slowed down to 0.016 m/ [1] J.-M. Kim, J.-E. Park, Y.-H. Kim, K.-S. Kang, C.-H. Kim, C.-
s when the gas flowed to the BeD region that was filled with the S. Park, K.-K. Bae, Decomposition of hydrogen iodide on Pt/C-
solid. The gas collided with the solid upon decreasing the flow based catalysts for hydrogen production, Int. J. Hydrogen
rate, and then the resulting mixture gas was well dispersed. Energy 33 (2008) 4974e4980.
[2] J.E. Funk, R.M. Reinstrom, Energy requirements in production
As shown in Fig. 6AeC, the static pressure and velocity
of hydrogen from water, Ind. Eng. Chem. Process. Des. Dev. 5
magnitude were independent of the heater temperature and (1966) 336e342.
STV ratio, and this could be attributed to the fact that the fluid [3] A. Giaconia, G. Caputo, A. Ceroli, M. Diamanti, V. Barbarossa,
porosity was 0.4 in all cases. Therefore, the pressure drop of the P. Tarquini, S. Sau, Experimental study of two phase separation
flowing mixture gas and velocity drop were affected not by the in the Bunsen section of the sulfureiodine thermochemical
thermal heat and surface area involved in the catalyst reaction, cycle, Int. J. Hydrogen Energy 32 (2007) 531e536.
[4] D.M. Ginosar, L.M. Petkovic, A.W. Glenn, K.C. Burch, Stability
but were rather affected by fluid porosity (solid factor).
of supported platinum sulfuric acid decomposition catalysts
In conclusion, the CFD analysis for HI decomposition
for use in thermochemical water splitting cycles, Int. J.
simulation was performed by applying the actual operation Hydrogen Energy 32 (2007) 482e488.
conditions and HI decomposer design. Several factors, such as [5] D. O'Keefe, C. Allen, G. Besenbruch, L. Brown, J. Norman,
hydrogen production, reaction heat, inner temperature, R. Sharp, K. McCorkle, Preliminary results from bench-scale
pressure, velocity, and mass distribution of the HI decom- testing of a sulfureiodine thermochemical water-splitting
poser, were investigated. cycle, Int. J. Hydrogen Energy 7 (1982) 381e392.
[6] Z. Wang, L. Wang, S. Chen, P. Zhang, J. Xu, J. Chen,
Decomposition of hydrogen iodide over PteIr/C bimetallic
 The hydrogen production depended on the STV ratio. In catalyst, Int. J. Hydrogen Energy 35 (2010) 8862e8867.
this study, the predicted hydrogen production was [7] D.R. O'Keefe, J.H. Norman, D.G. Williamson, Catalysis
1.12 mol/h, 1.99 mol/h, and 3.10 mol/h for STV ratios of 0.3, research in thermochemical water-splitting processes, Catal.
0.4, and 0.5, respectively. The hydrogen production at an Rev. 22 (1980) 325e369.
STV ratio of 0.5 was higher than that at 0.3 by 2 mol/h. [8] I. Ilda, The kinetic behaviour of the decomposition of
 For the three different STV ratios, the heat required for the HI hydrogen iodide on the surface of platinum,, Z. Phys. Chem.
109 (1978) 221e232.
decomposition reaction was investigated. With an increase
[9] Y. Shindo, N. Ito, K. Haraya, T. Hakuta, H. Yoshitome,
in the amount of hydrogen production, the reaction heat Kinetics of the catalytic decomposition of hydrogen iodide in
increased, and its value was 0.1117 kW/h, 0.1604 kW/h, and the thermochemical hydrogen production, Int. J. Hydrogen
0.2147 kW/h at STV ratios of 0.3, 0.4, and 0.5, respectively. Energy 9 (1984) 695e700.
 The inner temperature of the HI decomposer was studied [10] Y. Zhang, Z. Wang, J. Zhou, J. Liu, K. Cen, Effect of
in the form of contour images at ITC 1e5. A temperature preparation method on platinum-ceria catalysts for
hydrogen iodide decomposition in sulfureiodine cycle, Int. J.
higher than 750 K was required to generate the reaction
Hydrogen Energy 33 (2008) 602e607.
stably. In addition, at ITC 2 and ITC 3 in the center of the
[11] L. Wang, D. Li, P. Zhang, S. Chen, J. Xu, The HI catalytic
catalyst zone, the temperature was 773e801 K, which was decomposition for the lab-scale H2 producing apparatus of
lower than that at the preheating region. the iodineesulfur thermochemical cycle, Int. J. Hydrogen
 The inner pressure and velocity of the HI decomposer were Energy 37 (2012) 6415e6421.
investigated through pressure drop and reducing velocity [12] L.M. Petkovic, D.M. Ginosar, H.W. Rollins, K.C. Burch,
starting from the preheating zone. At this point, the pres- C. Deiana, H.S. Silva, M.F. Sardella, D. Granados, Activated
carbon catalysts for the production of hydrogen via the
sure drop was approximately 24 Pa and the reducing ve-
sulfureiodine thermochemical water splitting cycle, Int. J.
locity was 0.024 m/s. Hydrogen Energy 34 (2009) 4057e4064.
[13] Y. Oosawa, T. Kumagai, S. Mizuta, W. Kondo, Y. Takemori,
For future research, the effect of fluid porosity and STV K. Fujii, Kinetics of the catalytic decomposition of hydrogen
ratio on hydrogen production/reaction heat/hydraulic ther- iodide in the magnesiumeiodine thermochemical cycle, Bull.
mal analysis of the HI decomposer will be studied. The find- Chem. Soc. Jpn. 54 (1981) 742e748.
ings of this research can be used as a basis for the optimal [14] Y. Zhang, Z. Wang, J. Zhou, J. Liu, K. Cen, Catalytic
decomposition of hydrogen iodide over pre-treated Ni/CeO2
design of the HI decomposer.
catalysts for hydrogen production in the sulfureiodine cycle,
Int. J. Hydrogen Energy 34 (2009) 8792e8798.
[15] D. Neumann, Phasengleichgewichte von HJ/H2O/J2-
Conflicts of interest Lo sungen, PhD Thesis, Lehrstuhl fur Thermodynamik,
RWTH Aachen University, Aachen, Germany, 1987 [In
All contributing authors declare no conflicts of interest. German].
[16] H. Engels, K.F. Knoche, Vapor pressures of the system HI/
H2O/I2 and H2, Int. J. Hydrogen Energy 11 (1986) 703e707.
Acknowledgments nnissen, Direct dissociation of
[17] K.F. Knoche, H. Engels, J. Tho
hydrogen iodide into hydrogen and iodine from HI/H2O/Iz-
This work was supported by the National Research Founda- solution, in: Proc. 5th World Hydrogen Energy Conf., 1984.
tion of Korea (NRF) grant funded by the Korean government [18] C. Berndha user, K.F. Knoche, Experimental investigations of
(MSIP; Grant No. 53154-14). thermal HI decomposition from H2O-HI-I2 solutions, Int. J.
Hydrogen Energy 19 (1994) 239e244.
N u c l E n g T e c h n o l 4 7 ( 2 0 1 5 ) 4 2 4 e4 3 3 433

[19] M. Lanchi, A. Ceroli, R. Liberatore, L. Marrelli, M. Maschietti, iodineesulfur hydrogen production cycle, Int. J. Hydrogen
A. Spadoni, P. Tarquini, SeI thermochemical cycle: a Energy 37 (2012) 12967e12972.
thermodynamic analysis of the HIeH2OeI2 system and [29] L. Wang, Y. Imai, N. Tanaka, S. Kasahara, S. Kubo, K. Onuki,
design of the HIx decomposition section, Int. J. Hydrogen Thermodynamic considerations on the purification of H2SO4
Energy 34 (2009) 2121e2132. and HIx phases in the iodine-sulfur hydrogen production
[20] J.E. Murphy, J.P. O'Connell, A properties model of the process, Chem. Eng. Commun. 199 (2012) 165e177.
HIeI2eH2OeH2 system in the sulfureiodine cycle for [30] C.E. Bamberger, D.M. Richardson, Hydrogen production from
hydrogen manufacture, Fluid Phase Equilib. 288 (2010) water by thermochemical cycles, Cryogenics 16 (1976)
99e110. 197e208.
[21] M.K. Hadj-Kali, V. Gerbaud, J.-M. Borgard, O. Baudouin, [31] J.E. Funk, Thermochemical hydrogen production: past and
P. Floquet, X. Joulia, P. Carles, HIx system thermodynamic present, Int. J. Hydrogen Energy 26 (2001) 185e190.
model for hydrogen production by the sulfureiodine cycle, [32] G. de Beni, C. Marchetti, Hydrogen, key to the energy market,
Int. J. Hydrogen Energy 34 (2009) 1696e1709. Eurospectra 9 (1970) 46e50.
[22] J. Hur, J.P. O'Connell, K.-K. Bae, K.-S. Kang, J.W. Kang, [33] S. Kubo, H. Nakajima, S. Kasahara, S. Higashi, T. Masaki,
Measurements and correlation of solideliquid equilibria of H. Abe, K. Onuki, A demonstration study on a closed-cycle
the HI I2 H2O system, Int. J. Hydrogen Energy 36 (2011) hydrogen production by the thermochemical water-splitting
8187e8191. iodineesulfur process, Nucl. Eng. Design 233 (2004) 347e354.
[23] H. Guo, P. Zhang, S. Chen, L. Wang, J. Xu, Review of [34] S. Kubo, S. Kasahara, H. Okuda, A. Terada, N. Tanaka,
thermodynamic properties of the components in HI Y. Inaba, H. Ohashi, Y. Inagaki, K. Onuki, R. Hino, A pilot test
decomposition section of the iodineesulfur process, Int. J. plan of the thermochemical water-splitting iodineesulfur
Hydrogen Energy 36 (2011) 9505e9513. process, Nucl. Eng. Design 233 (2004) 355e362.
[24] C. Kane, S.T. Revankar, Sulfureiodine thermochemical cycle: [35] S. Kubo, H. Nakajima, S. Higashi, T. Masaki, S. Kasahara, M.
HI decomposition flow sheet analysis, Int. J. Hydrogen Energy Nomuara, R&D Program on Thermochemical Water-splitting
33 (2008) 5996e6005. Iodine-sulfur Process at JAERI, Rep. No. GENES4/ANP2003,
[25] L.C. Brown, G.E. Besenbruch, R.D. Lentsch, K.R. Schultz, JAERI, Kyoto, Japan, 1e7.
J.F. Funk, P.S. Pickard, A.C. Marshall, S.K. Showalter, High [36] H. Nakajima, M. Sakurai, K. Ikenoya, G.J. Hwang, K. Onuki,
Efficiency Generation of Hydrogen Fuels Using Nuclear S. Shimizu, in: A Study on a Closed-cycle Hydrogen
Power, Final Technical Report No. GA-A24285 Rev.1, General Production by Thermochemical Water-splitting IS Process,
Atomics, San Diego, CA, 2003. ICONE-7, Tokyo, Japan, 1999. ICONE-7104.
[26] M. Roth, K.F. Knoche, Thermochemical water splitting [37] A. Roine, HSC Chemistry for Windows, Chemical Reaction
through direct HI-decomposition from H2O/HI/I2/H2 and Equilibrium Software with Extensive Thermochemical
solutions, Int. J. Hydrogen Energy 14 (1989) 545e549. Database, Version 5.1, Outokumpu Research Oy, Pori,
[27] P. Wang, A. Anderko, R.D. Young, A speciation-based model Finland, 2002, ISBN 952-9507-08-9.
for mixed-solvent electrolyte systems,, Fluid Phase Equilib. [38] L.C. Brown, R.D. Lentsch, G.E. Besenbruch, K.R. Schultz,
203 (2002) 141e176. J.E. Funk, Alternative Flow Sheets for the Sulfureiodine
[28] L. Wang, Y. Imai, N. Tanaka, S. Kasahara, S. Kubo, K. Onuki, Thermochemical Hydrogen Cycle, Rep. No. GA-A24266,
S. Chen, P. Zhang, J. Xu, Simulation study about the effect of General Atomics, San Diego, CA, 2003.
pressure on purification of H2SO4 and HIx phases in the

Вам также может понравиться