You are on page 1of 9

Applied Energy 115 (2014) 531–539

Contents lists available at ScienceDirect

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

Kinetics and modeling of hydrogen iodide decomposition


for a bench-scale sulfur–iodine cycle
Thanh D.B. Nguyen a,b, Yun-Ki Gho a, Won Chul Cho a, Kyoung Soo Kang a, Seong Uk Jeong a,
Chang Hee Kim a, Chu-Sik Park a, Ki-Kwang Bae a,⇑
a
Department of Hydrogen and Fuel Cell, Korea Institute of Energy Research (KIER), Gajeong-ro 152, Yuseong-gu, Daejeon 305-343, Republic of Korea
b
School of Chemical Engineering, Hanoi University of Science and Technology, No. 1, Dai Co Viet, Hanoi, Viet Nam

h i g h l i g h t s

 Kinetics of HI decomposition over Pt/c-alumina 1.0 wt% in a S–I cycle is studied.


 New kinetic parameters of the reaction are estimated from the measured data.
3
 Modeling of a HI decomposer for hydrogen production rate of 1 Nm /h is carried out.
 Effects of reactor type and HIx composition on the reactor performance are analyzed.

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

Article history: In this work, the decomposition of hydrogen iodide (HI) over platinum catalyst in a frame work of the
Received 14 June 2013 development of a bench-scale Sulfur–Iodine (S–I) cycle is studied. The catalyst Pt/c-alumina 1.0 wt% is
Received in revised form 15 September 2013 prepared by impregnation–calcination method. The experiments of HI decomposition over the as-pre-
Accepted 17 September 2013
pared catalyst are conducted at the temperature range of 350–550 °C and at the atmospheric pressure.
Available online 10 December 2013
The experimental data are then used to estimate new kinetic parameters for HI decomposition on the
basis of Langmuir–Hinshelwood type where the surface reaction is considered as the rate-limiting step.
Keywords:
The kinetics with the estimated parameters shows a reasonable agreement with the experimental data. It
Sulfur–Iodine (S–I) cycle
Hydrogen iodide decomposition
also reflects the fact that, HI conversion is significantly decreased with a small amount of iodine present
Catalyst in the feeding solution.
Surface reaction Thereafter, the kinetic model is applied to the modeling of a HI decomposer for the hydrogen produc-
Langmuir–Hinshelwood model tion rate of 1 Nm3/h in which hot helium gas is used to provide heat for the decomposition. Effects of
Modeling heat-exchanger reactor configuration and composition of the feeding solution on the reactor size and
the heat consumed are examined using the proposed model. Calculation results show that heat con-
sumed for the co-current configuration is less than that for the counter-current configuration of the reac-
tor. I2 impurity and high water content in the feeding solution also result in an increase of reactor size and
the heat required.
Ó 2013 Published by Elsevier Ltd.

1. Introduction fossil fuels, thermo-chemical water-splitting cycles have been


studied [1–3].
Hydrogen has been considered as a clean fuel to substitute the Thermochemical water-splitting processes are conducted by
fossil fuel across the economy including transportation, stationary incorporating cyclic chemical reactions at which water can be
and mobile power generation. Also, it has been known as the made to react with chemical reagents and the products of chemical
energy carrier with the inherently clean nature since the heat re- reactions can be thermally decomposed to yield hydrogen and oxy-
leased from the reaction between hydrogen and oxygen is accom- gen at far lower temperatures compared to the thermal decompo-
panied by water as the only product. Thus, hydrogen energy shows sition of water. Therefore, water is considered as the feedstock for
the unique advantage compared to the common fossil fuels due to this kind of processes and all other reactants can be recovered and
the pollution and carbon dioxide emissions. As reported, in order to reused [4,5]. Literature review carried out by the researchers at the
avoid the pollution and contaminations from the utilizations of General Atomics (GA) reported that 115 different unique cycles of
thermochemical water-splitting processes have been studied [6].
⇑ Corresponding author. Tel.: +82 42 860 3445; fax: +82 42 860 3428. Among these processes, sulfur–iodine (S–I) cycle is expected to
E-mail address: kkbae@kier.re.kr (K.-K. Bae).

0306-2619/$ - see front matter Ó 2013 Published by Elsevier Ltd.


http://dx.doi.org/10.1016/j.apenergy.2013.09.041
532 T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539

Nomenclature

Cp,i molar specific heat capacity of component i, J/mol K pH2 O partial pressure of water, kPa
DR reactor diameter, m pHI partial pressure of hydrogen iodide, kPa
Fi flow rate of component i, mol/s pI2 partial pressure of iodine, kPa
DG0T Gibbs standard free energy change, kJ/mol p0HI initial partial pressure of hydrogen iodide, kPa
DHR heat of reaction, kJ/mol
p0I2 initial partial pressure of iodine, kPa
K H2 adsorption equilibrium constant of hydrogen, kPa1
rHI reaction rate of HI decomposition, mol/s-kg catalyst
K H2 O adsorption equilibrium constant of water, kPa1
R gas constant, J/mol K
KHI adsorption equilibrium constant of hydrogen iodide,
T absolute temperature, K
kPa1
TF absolute temperature of heating medium (helium gas),
K I2 adsorption equilibrium constant of iodine, kPa1
K
Kp reaction equilibrium constant
U overall heat transfer coefficient, J/s m2 K
k surface reaction rate constant, mol/kg catalyst-s
W catalyst weight, kg
kf surface forward reaction rate constant, mol/kg catalyst-s
x conversion of hydrogen iodide
kr surface reverse reaction rate constant, mol/kg catalyst-s
xE equilibrium conversion of hydrogen iodide
L length of the catalyst bed, m
qB bulk density of the catalyst, kg/m3
Mi molar mass of component i, kg/mol
pH2 partial pressure of hydrogen, kPa

be the most promising process that can be driven by nuclear en- and improve the reaction rate at different operating conditions
ergy for the hydrogen production [6,7]. including reaction temperature and composition of the reactants
In principle, the thermochemical S–I cycle of water-splitting [4,9–16]. Most of the experiments were conducted in a lab-scale
consists of the following reactions: fixed-bed reactor where the transportation of the reactants was as-
sumed to be the plug flow within the catalyst bed. The preferred
– Bunsen reaction (at 20–120 °C; DH ¼ 75  15 kJ/mol) catalysts used in those studies were Platinum (Pt) supported on
different substrates such as, activated carbon, alumina, ceria, and
2H2 O þ I2 þ SO2 ¼ H2 SO4 þ 2HI ðR:1Þ silica. Nickel and activated carbon have also been studied instead
of Pt to reduce the catalyst cost [17]. On the other hand, the effects
– Sulfuric acid decomposition of supporting materials, metal loading, and catalyst fabrication
(at 600–900 °C; DH ¼ 186  3 kJ/mol) methods on the performance of the hydrogen iodide decomposi-
H2 SO4 ¼ H2 O þ SO2 þ 1=2O2 ðR:2Þ tion have been studied [11,14,16] since the adsorption of iodine
and the unexpected distribution of the catalytic metals on the sub-
– Hydrogen iodide decomposition strates may hinder and diminish the conversion of hydrogen
(at 300–500 °C; DH  12 kJ/mol) iodide.
Several studies have proposed the reaction kinetics of HI cata-
2HI ¼ H2 þ I2 ðR:3Þ
lytic decomposition for the thermochemical water-splitting cycles
The Bunsen reaction is exothermic, sulfuric acid decomposi- [4,9,10]. The kinetic models were derived from experimental data
tion is strongly endothermic and hydrogen iodide is slightly based on specific assumptions of reaction mechanism using Lang-
endothermic. The net chemical change resulting from the chemi- muir–Hinshelwood rate expression since Langmuir–Hinshelwood
cal reaction cycle of the three reactions (Reactions (R.1), (R.2), model is often used to describe the kinetics of heterogeneously cat-
(R.3)) is water decomposition or water-splitting. The fundamental alytic reactions [18,19]. Oosawa et al. [9] proposed a kinetic model
description of the whole S–I process can be found elsewhere [6,8] for the catalytic decomposition of hydrogen iodide in the magne-
at which the cycle can be divided into three sections: (1) Bunsen sium–iodine thermo-chemical cycle using Platinum/Activated car-
reaction section, (2) sulfuric acid concentration and decomposi- bon (Pt/C) 1.05 wt% and active carbon catalysts. The model was
tion section, and (3) hydrogen iodide concentration and decom- valid at 1 atm and 500–700 K (227–427 °C). Shindo et al. [10] also
position. Although the S–I cycle is expected to be a promising developed a mathematic model for hydrogen iodide decomposition
process for hydrogen production, there are several critical fea- in a magnesium–iodine process where the commercial Platinum/
tures that should be highly concerned for the stable operations Alumina (Pt/c-alumina) 0.5 wt% was used. The model was devel-
and the thermal efficiency. One of these points is hydrogen iodide oped based on a simplification of the work proposed by Oosawa
decomposition. et al. [9]. Recently, a kinetic model of the hydrogen iodide decom-
HI decomposition (Reaction (R.3)) is only slightly endothermic. position over the commercial activated carbon was proposed by
However, it shows low conversion efficiency due to the thermody- Favuzza et al. [4] in which iodine atom was expected to be an auto
namic limitation. In order to increase the reaction rate, the thermal catalytic agent and the Langmuir–Hinshelwood model type was
decomposition of hydrogen iodide can be performed either in gas also utilized.
phase or liquid phase by using suitable catalysts [4,8]. However, In this work platinum catalyst supported by alumina
researches have been recently focused on the homogeneous gas- Pt/c-alumina 1.0 wt% is fabricated. Experiments of HI decompo-
phase reaction in the presence of solid catalysts. The decomposi- sition over the as-prepared catalyst are conducted at the temper-
tion is often operated at the temperature range of 300–500 °C ature range of 350–550 °C, 1 atm. New parameters for the
based on the standpoints of the construction materials and the reaction activation energy and heat of I2 adsorption are esti-
thermal efficiency of the whole cycle [9–11]. mated from the experimental data based on a reported reaction
Many experimental tests on hydrogen iodide decomposition mechanism [10]. Next, the kinetic model with the estimated
have been conducted over various kinds of solid catalysts to verify parameters is applied to the modeling of the HI decomposer in
T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539 533

a bench-scale S–I cycle for the hydrogen production rate of to 0.3 g of the loaded solid catalyst. The ratio of the thickness of the
1 Nm3/h using hot helium gas as heating medium. Effects of catalyst layer to the average diameter of the particles is therefore
the reactor configuration and composition of the feeding solution in the range of 26–37. Thus, the reaction rate is not effected by
on the performance of HI decomposition and on the reactor size the mass transfer process since it was reported that the back-
are analyzed. mixing effect is neglected when the ratio of the bed height to the
mean size of the catalyst particles is higher than 20 [9]. The
back-mixing of a continuous flow reactor is the propensity of the
2. Experiments products of a chemical reaction to intermingle with the feeding
of the reactants. Back-mixing is a detrimental feature of the flow
Pt/c-Al2O3 1.0 wt% is prepared by the impregnation–calcination reactors because this will influence negatively on the reaction
method. Alumina (c-Al2O3, Alfa–Aesar) is added into an aqueous kinetics. The ideal condition where no back-mixing takes place is
solution of Tetra amine platinum(II) chloride hydrate 98% called plug flow [20].
(Pt(NH3)4Cl2xH2O, Sigma–Aldrich) with Pt loading of 1.0% and All the outlet gases except H2 are first condensed to separate I2
the mixture is stirred in a round-bottomed flask. This step is fol- from the products. H2O and unreacted HI are then absorbed in KI
lowed by the impregnation process which is conducted in a rotary solution and finally the trace of H2O is trapped by silica gel. Con-
vacuum evaporator (R-215, BÜCHI) under a vacuum condition at centrations of H2 and Argon in the outlet gas flow are detected
60 °C for 3 h and then the catalyst is dried in an oven (OF-22GW, by an online GC (7890A GC by Agilent Technologies). The outlet
JEIO Tech.) at 100 °C for 24 h. Finally, the catalyst is reduced by flow rate of Argon gas is measured by a film flow meter (VP-3U
hydrogen (50 vol.% and Argon balanced) at 700 °C for 6 h to remove by HORIBA). The measurement of Argon gas flow rate at the outlet
residuals. The particle sizes are in the range of 300–425 lm. The of the system is used to check the leakage from the entire system
commercial HI solution 56–58 wt% by KANTO Chemical Co., Inc. and to calculate the flow rate of hydrogen gas produced from HI
is used as the reactant. The HIx (HI + H2O + I2) solution is also used decomposition. The conversion of HI is then calculated from the
to study the effect of the I2 impurity on HI conversion since a cer- measurement of H2 concentration and the calculated H2 flow rate.
tain amount of I2 may be remained in the feeding solution of the HI The experimental data is presented and analyzed along with the
decomposer in practical systems. kinetic rate equation in the next section.
The experimental system is depicted in Fig. 1. A tubular fixed-
bed reactor with the inner diameter of 9 mm is used as the decom-
poser. The reactant is heated up and vaporized at around 250 °C 3. Kinetics of HI decomposition
and is fed to the reactor along with the carrier gas (Argon). The
flow rate of the feeding HIx solution is varied from 0.1 to 0.9 ml/ As mentioned, the Langmuir–Hinshelwood model was utilized
min and that of Argon gas is in the range of 100–900 ml/min. for the derivation of the kinetics of hydrogen iodide decomposition
The decomposition is conducted at atmospheric pressure and at a since the reaction was practically conducted in the presence of an
temperature range of 350–550 °C. The reaction temperature is con- appropriate solid catalyst. The reaction mechanism proposed by
trolled and regulated by an electrical heater. The reaction time is Shindo et al. [10] is adopted for the kinetic study in the present
varied by changing the flow rate of the feeding solution at a fixed work. The decomposition scheme is illustrated in Fig. 2 and the
amount of catalyst. The height of the bed is 11 mm corresponding reaction mechanism can be written as follows.

TC
PG Quenching

On-line GC

Oil Drain
Reactor
Circulator
Peristaltic
pump

PG
Evaporizer Vent
Ar

MFC
HI
TC solution
Silica gel
Peristaltic KI
pump solution

Fig. 1. Scheme of the experimental system for the decomposition of HI and HIx solutions.
534 T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539

H2O It was suggested by Oosawa et al. and Shindo et al. [9,10] that
HI H2 I2
the adsorption of HI, H2 and H2O on the surface of the catalyst is
phase
Bulk
K HI very small compared to that of I2. Therefore, the adsorption terms
K H2 K I2 KH 2O
of these three components are eliminated from the denominator of
kf Eq. (7) and the final rate equation of the hydrogen iodide decompo-
Catalyst
surface

HI-σ + σ H-σ + I-σ H2O- σ sition can be expressed as


kr
0
dx kð1  xÞ  k 2x
rHI ¼ ¼ 2 ð8Þ
Fig. 2. The reaction scheme of hydrogen iodide decomposition on the surface of
dt pffiffiffiffiffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p0 x
1 þ K I2 p0I2 þ HI2
Pt/c-alumina catalyst.

4. Results and discussion


HI þ r ¼ HI  r ðR:4Þ
HI  r þ r ¼ H  r þ I  r ðR:5Þ 4.1. Kinetic parameter estimation
2H  r ¼ H2 þ 2r ðR:6Þ
2I  r ¼ I2 þ 2r ðR:7Þ In order to estimate the kinetic parameters the differential
method is used and the procedure of the estimation is described
H2 O þ r ¼ H2 O  r ðR:8Þ
as follows.
where r represents an adsorption site on the catalyst surface. At equilibrium state where x = xE and dx
dt
¼ 0, Eq. (8) yields
The mechanism is developed based on the assumptions that the 2kð1  xE Þ
0
adsorptions of hydrogen iodide and water vapor are molecular k ¼ ð9Þ
xE
while those of iodine and hydrogen are dissociative [4,10,21,22].
The rate-determining step is expected to be the decomposition of Here, the equilibrium conversion xE is calculated from the thermo-
hydrogen iodide on the catalyst surface (Reaction (R.6)). Therefore dynamic equilibrium constant of the reaction HI M 1/2 H2 + 1/2 I2,
this reaction is used for the derivation of the kinetics of HI catalytic !
xE DG 0
decomposition. According to Langmuir–Hinshelwood model, the Kp ¼ ¼ exp  T ð10Þ
kinetic equation describing the rate of reaction (R.6) becomes: 2ð1  xE Þ RT
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffi
kf  K HI pHI  kr  K H2 pH2  K I2 pI2 using the thermodynamic data source from the National Aeronau-
dp tics and Space Administration (NASA) at Glenn research center [23].
r HI ¼  HI ¼ h qffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi i2 ð1Þ
dt By substituting Eq. (9) into Eq. (8), the following expression is
1 þ K I2 pI2 þ K H2 pH2 þ K HI pHI þ K H2 O pH2 O
obtained:
in which rHI is reaction rate; pHI ; pI2 and pH2 O are the partial pressure h i
of hydrogen iodide, iodine and water vapor respectively; dx k ð1  xÞ  ð1x xE
E Þx

rHI ¼ ¼ 2 ð11Þ
K HI ; K I2 ; K H2 ; K H2 O are adsorption equilibrium constants of the corre- dt pffiffiffiffiffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p0 x
sponding components. The reaction time (contact time), t (s), is de- 1 þ K I2 p0I2 þ HI2
fined as
or
W vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t¼P ð2Þ uh iffi
F i  Mi u ð1  xÞ  ð1xE Þx rffiffiffiffiffiffiffirffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t xE 1 K I2 p0 x
where W (kg) is weight of the catalyst, Fi (mol/s) and Mi (kg/mol) are ¼ pffiffiffi þ p0I2 þ HI ð12Þ
r HI k k 2
the feed rate and molar mass of each component feeding to the
reactor. Eq. (12) can be expressed as a linear function (p1and p2 are indepen-
0 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Let k = kfKHI and k ¼ kr K I2 K H2 , Eq. (1) becomes dent variables)
0 pffiffiffiffiffiffiffiffiffiffiffiffiffi
dpHI kpHI  k pI2 pH2 y ¼ p1 x1 þ p2 x2 ð13Þ
r HI ¼  ¼h qffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi i2 ð3Þ
dt
1 þ K I2 pI2 þ K H2 pH2 þ K HI pHI þ K H2 O pH2 O in which
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
iffi
In terms of HI conversion, x, the partial pressures of the components uh rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u ð1  xÞ  ð1xE Þx
are expressed as t xE p0 x
y¼ ; x1 ¼ 1; x2 ¼ p0I2 þ HI and
r HI 2
pHI ¼ p0HI ð1  xÞ ð4Þ rffiffiffiffiffiffiffi
x 1 K I2
pH2 ¼ p0HI ð5Þ p1 ¼ pffiffiffi ; p2 ¼
2 k k
x
pI2 ¼ p0I2 þ p0HI ð6Þ The value of x2 is calculated from the measured data at each
2
experiment and the parameters p1 and p2 are obtained from the
where p0HI
and p0I2
are the initial pressures of hydrogen iodide and linear regression method. The surface reaction rate constant k
iodine in the feeding solution. Hence, Eq. (3) can be written as: and I2 adsorption constant K I2 are then estimated based on the val-
dx ues of p1 and p2 and expressed as the following equations as the
r HI ¼ Arrhenius function of temperature.
dt
0  
kð1  xÞ  k 2x 37:1
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffi 2 k ¼ 3:13  101 exp  ð14Þ
pffiffiffiffiffiffiffi p0 x pffiffiffiffiffiffiffiffi p0 x RT
1 þ K I2 p0I2 þ HI2 þ K H2 HI2 þ K HI p0HI ð1  xÞ þ K H2 O pH2 O  
75:1
K I2 ¼ 1:80  109 exp ð15Þ
ð7Þ RT
T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539 535

Table 1
Kinetic parameters of HI decomposition on different catalyst types.

Studies Catalyst type Activation energy (kJ/mol) Adsorption heat of I2 (kJ/mol)


This study Pt/c-alumina 1.0 wt% 37.1 75.1
Shindo et al. [10] Pt/c-alumina 0.5 wt% 27.6 95.8
Oosawa et al. [9] Pt/Activated carbon 1.05 wt% 53.2 86.7
Oosawa et al. [9] Activated carbon 34.4 86.7
Favuzza et al. [4] Activated carbon 55.5 86.6

As given in Eqs. (14) and (15), the apparent activation energy of 0.25
hydrogen iodide composition over the Pt/c-alumina 1.0 wt% is
37.1 kJ/mol and the adsorption heat of iodine on the catalyst is
0.2
75.1 kJ/mol. The kinetic parameters of HI decomposition over dif-

HI conversion, X [-]
ferent types of catalyst obtained from the literature [4,9,10] are
also given in Table 1 in comparison with that proposed in this 0.15
study. Regarding to the values reported in the literature, the activa-
tion energy and the adsorption heat of I2 estimated in the present
study are reasonable and the Pt/c-alumina catalyst is effective 0.1
with HI decomposition. Model-I2/HI = 0.00
Model-I2/HI = 0.05
The integration form of the rate equation (Eq. (11)) shows the 0.05 Model-I2/HI = 0.10
variation of hydrogen iodide conversion as a function of contact Data-I2/HI = 0.00
Data-I2/HI = 0.05
time at a specific reaction temperature. Calculated results obtained Data-I2/HI = 0.10
from the kinetic model are presented in Fig. 3 in comparison with 0
0 1 2 3 4 5 6
the experimental data. It is observed that at a given temperature
hydrogen iodide conversion increases as the contact time in- Contact time t [s]
creases. Slope of the conversion is high at the initial stage of the
Fig. 4. Effects of excess I2 in the feeding solution on the conversion of HI at 450 °C.
reaction and then it tends to slow down when the contact time
is increased since at the initial period the reactant is readily
adsorbed on the catalyst with less adsorptive resistance. As an obtained from the proposed kinetic model shows a reasonable
endothermic reaction, it is also observed that the conversion of agreement with the present experimental data. However, the
hydrogen iodide increase with an increase in the decomposition model does not capture well some data points (see Fig. 3) because
temperature. Pt/c-alumina catalyst exhibits a high and unstable activity at the
In a practical process there is a certain amount of iodine re- initial stages of reaction [10].
mained in the solution feeding to the HI decomposer. The presence
of iodine in the feeding solution induces a decrease in the conver- 4.2. Modeling of a HI decomposer for 1 Nm3/h of hydrogen
sion of hydrogen iodide as iodine is one of the products of the
decomposition. The effect of molar ratio of iodine and hydrogen io- Based on the proposed kinetics, the modeling of a HI decom-
dide (I2/HI) in the feeding solution on hydrogen iodide conversion poser for 1 Nm3/h of hydrogen in a closed S–I cycle with hot he-
obtained from the experimental data and the numerical results at lium gas loop is presented in this section. The sensitivity study
450 °C is presented in Fig. 4. As observed, hydrogen iodide conver- for the design purpose of the catalyst bed is also addressed.
sion decreases significantly with the presence of a small amount of
iodine in the feeding solution. It is attributed to the shift of the 4.2.1. Model equations
backward reaction (R.3) in the presence of excess iodine. This The governing equations consisting of mass and heat balances
phenomenon was also reported by other research works [4,9,10]. are presented below under the assumptions: (1) the reactor is
It is also observed from Figs. 3 and 4 that, the numerical results operated in the steady state and the transport of the reactant is
plug-flow; (2) the catalyst bed is uniformly distributed; (3) axial
and radial dispersion of mass and energy are neglected; and (4)
the feeding mixture is an ideal gas mixture. The reactor model of
the HI catalytic decomposer is depicted in Fig. 5.
0.25
The material balance for the reactant hydrogen iodide over an
0.2
elementary slide of a solid catalyzed plug-flow reactor gives
HI conversion, X [-]

F HI;0  F HI;0 xin ¼ F HI;0  F HI;0 xout þ r HI DW ð16Þ


0.15
Model-623K
Model-673K
0.1 Model-723K
Model-773K FF,in, TF,in
Model-823K dW dL
DR

0.05 Data-623K
Data-673K Fi,in Fi,out
0 Data-723K
Data-773K Tin Tout
Data-823K
-0.05
0 1 2 3 4 5 6
L
Contact time t [s]
FF,out, TF,out
Fig. 3. Effects of temperature and reaction time on the conversion of hydrogen
iodide. Fig. 5. Schematic diagram of the HI catalytic decomposer used in this study.
536 T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539

In the differential form, Eq. (16) can be written as HIx outlet


dx rHI Comp. mol/h
¼ ð17Þ H2O 1,733.9
dW F HI;0
I2 44.6
For a given bulk density of the catalyst bed, qB (the bulk density HI 357.1
of the catalyst used in this work qB ¼ 625 kg/m3), the weight of the 44.6
H2 mHe=10,901.5 mol/h
catalyst in a differential volume of the reactor, dW, is expressed in
Temp. ~450 oC THe=700 oC
terms of the reactor diameter, DR and the differential length, dL as
Press. 7 bar-g
pD2R ml/min 307,958
dW ¼ qB dL ð18Þ
4
By substituting Eq. (18) into Eq. (17), the conversion of hydro- HIx inlet
gen iodide along the reactor can be obtained as follows mHe=10,901.5 mol/h Comp. mol/h
THe=610 oC
H2O 1,733.9
dx p D2R
¼ q rHI ð19Þ I2 0.0
dL 4F HI;0 B
HI 446.4
In a closed system, hot helium gas is selected as the heating Total 2,180.3
medium for the endothermic reaction of HI decomposition. Thus, Temp. 195 oC
a tubular reactor in the form of a heat exchanger is often used. Press. 7 bar-g
At the industrial scale, the reactor in the form of shell-and-tube ml/min 199,342
heat exchanger is usually adopted in which the reactants flow in
the tube side where the catalyst is loaded while the heating fluid Fig. 6. Operating condition of the HI decomposer for 1 Nm3/h of hydrogen.
is transported in the shell side. However, for a bench scale
with the hydrogen production rate of 1 Nm3/h, a double-pipe
heat-exchanged reactor is used (see Fig. 5). the helium gas at the outlet of the decomposer. As obtained from
The energy balance in terms of the temperature of the process the process simulation, the outlet temperature of the helium gas
stream and the heating stream along the reactor is in turn ex- is 610 °C (see Fig. 6).
pressed as follows. The temperature gradient of the main stream Because the catalyst bed diameter and the bed length are not
flowing along the catalyst bed in the inner pipe is provided in the process simulation, in this work the diameter of
the catalyst bed is supposed to be in the range of 2 in.
dT U pDR ðT F  TÞ pD2
¼ P  P R q rHI DHR ðTÞ ð20Þ (0.0254 m) to 3 in. (0.0762 m). The length of the catalyst bed is
dL F i C p;i ðTÞ 4 F i C p;i ðTÞ B determined from the calculation result of the present model.
The temperature gradient of the heating fluid along the reactor at
the outer pipe can be written as
4.2.3. Numerical results
dT F U pDR ðT F  TÞ In this section, calculation results obtained from the modeling
¼ ð21Þ
dL F F C p;F ðT F Þ of the HI decomposer in the form of double-pipe heat exchanger
type for 1 Nm3/h of hydrogen are presented by solving the model
In Eq. (21), the plus sign ‘‘+’’ indicates that the flow of the heat-
equations (Eqs. (19)–(21)). The effects of flow direction of the heat-
ing fluid is counter-currently to that of reactants and the minus
ing medium (hot helium gas flow) on the reactor size, catalyst
sign ‘‘’’ implies that the flows of the heating medium and the
weight and on the required heat load are also examined.
reactants are co-current. Details of the derivation of Eqs. (20) and
Fig. 7 shows the equilibrium and calculated HI conversion and
(21) can be found elsewhere [24,25]. The overall heat transfer coef-
both the HIx gas and helium temperature profiles along the
ficient U in the packed beds for the heat-exchanger reactor type
catalyst bed where the diameter of the catalyst bed is 3 in.
was reported to be in the range of 152.35 J/s m2 K to 187.24 J/
(0.0762 m) and the heating medium flows co-currently with the
s m2 K [26]. In the present work the average value of 170 J/s m2 K
is used.
0.25
4.2.2. Boundary and operating conditions 700
The diagram of the HI decomposer using helium loop for the 650
0.20
hydrogen production of 1 Nm3/h obtained from the process simu- 600
lation is shown in Fig. 6. The conditions of the process and heating
HI conversion (-)

550
Temperature ( C)

streams are also given. In this bench scale, the preheating treat- 0.15
o

500
ment of the HIx solution before feeding to the decomposer is sup-
posed not to use due to high capital cost. Therefore, the HIx 450
0.10
solution obtained from the top product of the HI distillation col- Dreactor = 3.0 in. = 0.076 m
400
umn is directly fed to the decomposer at 195 °C. In order to pro- o
THIx,in = 195 C 350
duce 1 Nm3/h of hydrogen, the total mole flow rate of the 0.05 THe,in = 700 C
o

300
feeding HIx solution is 2180.3 mol/h under the assumption that HI calculated conversion
20% of the feeding HI is converted. For the heating medium, helium HI equilibrium conversion 250
0.00 Temperature of reactant flow
gas at about 700 °C is fed to the outer pipe at the mole flow rate of Temperature of He flow 200
10901.5 mol/h to provide the heat for the decomposition. The inlet
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
temperature of helium gas (700 °C) is selected based on the sug-
Length of catalyst bed (m)
gestions reported in literature [8,27]. The flow rate of the heating
stream is selected based on the calculation of the heat demand Fig. 7. HI conversion and temperature profiles along the catalyst bed for co-current
of the whole S–I cycle and based on the desired temperature of double-pipe heat-exchanged configuration.
T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539 537

process stream. As the desired goal of 1 Nm3/h of hydrogen, the 0.25 550
length of the bed is 0.613 m and the required weight of the Dreactor = 3.0 in. = 0.076 m
Pt/c-alumina 1.0 wt% catalyst is 1.747 kg. The outlet temperatures THIx,in = 195 C
o 500
0.20 o
of the product gases and helium gas in this case are 483 and 598 °C, THe,in = 700 C
2
U = 170 J/s-m -K 450
respectively.

HI conversion (-)

Temperature ( C)
In case of using the counter-current configuration, HI conver- 0.15

o
400
sion and temperature profiles along the length of the catalyst
bed with the diameter of 3 in. (0.0762 m) are presented in Fig. 8. 350
0.10
The calculation results show that for the expected production rate
of hydrogen, the bed length of 0.603 m and the catalyst weight of QHe,co = 6.46 kW 300
1.719 kg are required. The temperature of helium gas flow is re- 0.05 QHe,counter = 6.89 kW
duced to 591 °C and that of the process stream is increased to HI convers.-cocurrent
250
502 °C at the outlets. 0.00
HI convers.-countercurrent
Temp.-cocurrent
It is observed that by changing the flow direction of the heating Temp.-countercurrent
200
medium from counter-current regime to co-current regime, the
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
outlet temperature of the product gases is increased while that
of helium gas is decreased. In order to analyze the effect of the Length of catalyst bed (m)
heating flow direction, the temperature of the reacting gases and Fig. 9. Effect of co-current and counter-current flow configuration on HI conversion
HI conversion along the bed for the two regimes are depicted in and temperature of the reacting gases.
Fig. 9. Since the temperature difference between the two flows in
the co-current regime is higher than that in the counter-current
configuration for the major part of the bed (see Figs. 7 and 8), 1.8
the reacting gas temperature and HI conversion within the bed is
higher in the case of co-current flow regime (see Fig. 9). As calcu- 7.5
1.7
lated, for the bed diameter of 3 in. (0.0762 m) the heat duty pro- Weight of catalyst (kg)

Required heat load (kW)


vided for the co-current and counter-current configurations is
1.6
6.46 and 6.89 kW, respectively. Thus, the heat consumed for the
decomposition is lower when the co-current flow regime is used. 7.0
However, it is worthily noted that heat transfer characteristics 1.5
discussed above might be different from other bench-scale S–I
cycles where the preheating treatment of the feeding solution is 1.4
6.5
included or where another reactor configuration is used for HI
decomposition. 1.3 Weight of catalyst (counter-current)
Weight of catalyst (co-current)
A number of calculations are also carried out based on the pres- Required heat load (counter-current)
ent model to examine the effect of the bed diameter on the catalyst Required heat load (co-current)
1.2 6.0
weight and the required heat load for the decomposition. As shown
2.0 2.5 3.0
in Fig. 10, when the bed diameter increases, the catalyst weight in-
Reactor diameter (in)
creases while the heat required is decreased for both co-current
and counter-current flow regimes of the two fluids. It is found that Fig. 10. Effects of the reactor diameter on the weight of catalyst and the required
by increasing the bed diameter the heat transfer area between the heat load.
heating medium and the reacting gases is decreased which results
in a decrease in the heat consumed. However, the catalyst weight
is increased to meet the required production rate. 0.25
As mentioned, there is a certain amount of I2 present in the gas
mixture feeding to the decomposer in practical operations which is
0.20
HI conversion (-)

0.25
700 0.15
Dreactor = 3.0 in. = 0.076 m
o
650 THIx,in = 195 C
o
0.20 THe,in = 700 C
600 0.10 U = 170 J/s-m -K
2
HI conversion (-)

550
Temperature ( C)

I2/HI = 0.000
0.15
o

500 0.05 I2/HI = 0.025


450 I2/HI = 0.050
0.10 I2/HI = 0.075
Dreactor = 3.0 in. = 0.076 m 400
o
0.00
THIx,in = 195 C 350
o
0.0 0.5 1.0 1.5 2.0
0.05 THe,in = 700 C
300 Length of catalyst bed (m)
HI calculated conversion
HI equilibrium conversion 250
0.00 Temperature of reactant flow Fig. 11. Effect of excess I2 in the feeding solution on HI conversion long the catalyst
Temperature of He flow 200 bed for the co-current configuration of the decomposer.

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7


Length of catalyst bed (m)
caused by the HIx distillation process. For the co-current configu-
Fig. 8. HI conversion and temperature profiles along the catalyst bed for the ration of the decomposer, calculation results given in Fig. 11 show
counter-current double-pipe heat-exchanged configuration. that the size of the reactor increases considerably with only a small
538 T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539

7.0 the rate-limiting step. The kinetics with the proposed parameters
1.8 Catalyst weight
6.5 not only shows a reasonable agreement with the experimental
Required heat load
6.0 data but also reflects the negative effect of I2 impurity in the feed-
1.7 5.5 ing solution.

Required heat load (kW)


5.0 For the design purpose, the modeling of a HI decomposer for the
Catalyst weight (kg)

1.6
4.5 hydrogen production rate of 1 Nm3/h with the heat provided from
4.0 hot helium gas is carried out. Numerical results show that without
1.5
3.5 the preheating step of the feeding solution, the heat consumed in
1.4 3.0 the co-current double-pipe heat-exchanged reactor is less than
2.5 that in the counter-current configuration. Both iodine impurity
1.3 Dreactor = 3.0 in. = 0.076 m 2.0 and high water content in the initial solution cause an increase
o
THIx,in = 195 C 1.5 of reactor size and the heat required. It is suggested that the re-
o
1.2 THe,in = 700 C 1.0
U = 170 J/s-m -K
2 moval of I2 from the feeding solution is necessary. However, based
0.5
on the practical point of view, the water content should be kept at
1.1 0.0
0.2 0.3 0.4 0.5 0.6 0.7
a certain high level with in the common range (HI/H2O = 1/4 to 1/
1.5) in order to wash out the adsorbed iodine from the catalyst bed
HI/H2O (-) in the continuous operation.
Fig. 12. Effect of H2O content in the feeding solution on the catalyst weight and the
required heat load. Acknowledgements

This work has been performed by Nuclear Hydrogen Production


amount of excess iodine due to the negative effect of iodine impu-
Technology Development and Demonstration (NHDD) Project and
rity on the reaction rate as discussed above. Therefore, the removal
conducted under the framework of Research and Development
of the impure I2 in the top product of the HIx distillation should be
Program of the Korea Institute of Energy Research (KIER)
considered or the feeding reactants should be preheated up to the
(B3-2413-03). The authors appreciate the financial support of
desired temperature to reduce the size of the decomposer.
Ministry of Science, ICT & Future Planning of Korea.
Another concern experienced by the authors that the mole ratio
of HI and H2O (HI/H2O) in the feeding solution also affects the per-
formance of the HI decomposition process. The ratio HI/H2O is re- References
sulted from the top product of the HIx distillation and it is varied in
a relatively high range from 1/4 to 1/1.5. Thus, the effect of HI/H2O [1] Brown LC, Funk JF and Showalter SK. High efficiency generation of hydrogen
fuels using nuclear power. Nuclear Energy Research Initiative (NERI): General
mole ratio in the feeding solution of the decomposition on the cat-
Atomics; 2000 July [report no.: GAA23451].
alyst weight and the heat required is studied in the present work. [2] Kasahara S, Kubo S, Hino R, Onuki K, Nomura M, Nakao S-I. Flowsheet study of
Fig. 12 shows the required catalyst weight and heat consumed as a the thermochemical water-splitting iodine–sulfur process for effective
function of HI/H2O mole ratio for the production rate of 1 Nm3/h of hydrogen production. Int J Hydrogen Energy 2007;32:489–96.
[3] Shin Y, Lee K, Kim Y, Chang J, Cho W, Bae K. A sulfur–iodine flowsheet using
hydrogen. It is observed that the catalyst weight and the heat re- precipitation, electrodialysis, and membrane separation to produce hydrogen.
quired for the decomposition are both increased when the H2O Int J Hydrogen Energy 2012;37:16604–14.
content in the feeding solution increases since the specific heat [4] Favuzza P, Felici C, Nardi L, Tarquini P, Tito A. Kinetics of hydrogen iodide
decomposition over activated carbon catalysts in pellets. Appl Catal B: Environ
capacity of H2O is the highest compared to that of the other com- 2011;105:30–40.
ponents HI, I2, and H2 which are present in the reacting gaseous [5] Hino R, Matsui K, Yan XL. The role of hydrogen in the world economy. In: Yan
flow. Furthermore, when the water content in the HIx solution in- XL, Hino R, editors. Nuclear hydrogen production handbook. New York: CRC
press; 2011. p. 3–46.
creases the total mole flow rate of the reactants also increases to [6] Brown LC, Lentsch RD, Besenbruch GE, Schultz KR, Funk JF. Alternative
meet the required production rate. Therefore, an increase in H2O flowsheets for the sulfur-iodine thermochemical hydrogen cycle. General
content in the feeding solution leads to an increase of the heat con- Atomics; 2003 February [report no.: GAA24266].
[7] Li H, Tan G, Zhang W, Suppiah S. Development of direct resistive heating
sumed for the decomposition. An increase in H2O content in the method for SO3 decomposition in the S–I cycle for hydrogen production. Appl
feeding solution also leads to a decrease in the partial pressure of Energy 2012;93:59–64.
HI which results in a decrease of the reaction rate. Thus, the cata- [8] Onuki K, Kubo S, Tanaka N, Kasahara S. Thermochemical iodine–sulfur process.
In: Yan XL, Hino R, editors. Nuclear hydrogen production handbook. New
lyst required is increased with the increase of H2O content. How-
York: CRC press; 2011. p. 461–98.
ever, as encountered by the authors, it is suggested that H2O [9] Oosawa Y, Kumagai T, Mizuta S, Kondo W, Takemori Y, Fujii K. Kinetics of the
content should be kept at a high level within the observed range catalytic decomposition of hydrogen iodide in the magnesium–iodine
(I2/H2O = 1/4 to 1/1.5) so that I2 can be easily washed out from thermochemical cycle. Bull Chem Soc Jpn 1981;54:742–8.
[10] Shindo Y, Ito N, Haraya K, Hakuta T, Yoshitome H. Kinetics of the catalytic
the catalyst bed during the continuous operation. decomposition of hydrogen iodide in the thermochemical hydrogen
production. Int J Hydrogen Energy 1984;9:695–700.
[11] Kim J-M, Park J-E, Kim Y-H, Kang K-S, Kim C-H, Park C-S, et al. Decomposition
of hydrogen iodide on Pt/C-based catalysts for hydrogen production. Int J
5. Conclusions Hydrogen Energy 2008;33:4974–80.
[12] Chen Y, Wang Z, Zhang Y, Zhou J, Cen K. Platinum–ceria–zirconia catalysts for
The platinum catalyst Pt/c-alumina 1.0 wt% is prepared for the hydrogen production in sulfur–iodine cycle. Int J Hydrogen Energy
2010;35:445–51.
decomposition of hydrogen iodide in a sulfur–iodine thermody- [13] Favuzza P, Felici C, Lanchi M, Liberatore R, Mazzocchia CV, Spadoni A, et al.
namic water-splitting cycle. Experiments on the decomposition Decomposition of hydrogen iodide in the S–I thermochemical cycle over Ni
over the as-prepared catalyst are conducted under various condi- catalyst systems. Int J Hydrogen Energy 2009;34:4049–56.
[14] Wang Z, Wang L, Chen S, Zhang P, Xu J, Chen J. Decomposition of hydrogen
tions showing that Pt/c-alumina is effective with the decomposi-
iodide over Pt–Ir/C bimetallic catalyst. Int J Hydrogen Energy 2010;35:8862–7.
tion of hydrogen iodide. However, the excess of I2 in the feeding [15] Zhang Y, Wang Z, Zhou J, Liu J, Cen K. Effect of preparation method on
solution causes a significant reduction of hydrogen iodide conver- platinum–ceria catalysts for hydrogen iodide decomposition in sulfur–iodine
sion. New kinetic parameters for the decomposition are also cycle. Int J Hydrogen Energy 2008;33:602–7.
[16] Zhang Y, Zhou J, Wang Z, Liu J, Cen K. Catalytic thermal decomposition of
estimated from the experimental data based on Langmuir– hydrogen iodide in sulfur–iodine cycle for hydrogen production. Energy Fuel
Hinshelwood model where the surface reaction is considered as 2008;22:1227–32.
T.D.B. Nguyen et al. / Applied Energy 115 (2014) 531–539 539

[17] Zhang Y, Zhu Q, Lin X, Xu Z, Liu J, Wang Z, et al. A novel thermochemical cycle [23] McBride BJ, Zehe MJ, Gordon S. NASA Glenn coefficients of calculating
for the dissociation of CO2 and H2O using sustainable energy sources. Appl thermodynamic properties of individual species (NASA/TP-2002-
Energy 2013;108:1–7. 211556). Cleveland, Ohio: NASA Glenn Research Center; 2002.
[18] Kaliaguine S, Kaliaguine N, Vanderschuren J. A powerful method for Langmuir– [24] Levenspiel O. Chemical reaction engineering. 3rd ed. New York: John Wiley &
Hinshelwood models discrimination. Chem Eng Sci 1971;26:1169–78. Sons, Inc.; 1999.
[19] Krishna R, Baur R. On the Langmuir–Hinshelwood formulation for zeolite [25] Mann U. Principles of chemical reactor analysis and design: new tools for
catalysed reactions. Chem Eng Sci 2005;60:1155–66. industrial chemical reactor operations. 2nd ed. New Jersey, Canada: John
[20] Liguori L, Bjørsvik H-R. Multijet oscillating disc millireactor: a novel approach Wiley & Sons, Inc.; 2009.
for continuous flow organic synthesis. Org Process Res Dev 2011;15: [26] Lin SS, Flaherty R. Design studies of the sulfur trioxide decomposition reactor
997–1009. for the sulfur cycle hydrogen production process. Int J Hydrogen Energy
[21] Ferng S-S, Lin D-S. Iodine adsorption on arrays, clusters, and pairs of reactive 1983;8:589–96.
sites on the Si(1 0 0) surface. J Phys Chem C 2011;116:3091–6. [27] Shin Y, Chang J, Lee T, Lee K, Kim Y. A cooling system for the secondary helium
[22] Desai R, Hussain M, Ruthven DM. Adsorption on activated alumina. II  kinetic loop in VHTR-based SI hydrogen production facilities. Int J Hydrogen Energy
behaviour. Can J Chem Eng 1992;70:707–15. 2013;38:6182–9.

You might also like