You are on page 1of 10

Applied Catalysis B: Environmental xxx (xxxx) xxxx

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


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

Efficient recovery of hydrogen and sulfur resources over non-sulfide based


LaFexAl12-xO19 hexaaluminate catalysts by H2S catalytic decomposition
Guoxia Jianga,b, Xin Zhanga,c,*, Fenglian Zhanga,b, Zizhong Liue, Zhuo Wangf, Zhengping Haoa,b,c,
Caihong Lind
a
National Engineering Laboratory for VOCs Pollution Control Material & Technology, University of Chinese Academy of Sciences, Beijing, 101408, PR China
b
Key Laboratory of Environmental Nanotechnology and Health Effects, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing, 100085,
PR China
c
Research Center for Environmental Material and Pollution Control Technology, University of Chinese Academy of Sciences, Beijing, 101408, PR China
d
Center of Research & Development, Shandong Sunway Petrochemical Engineering Share Co., Ltd, Beijing, 100015, PR China
e
Chemistry and Environment Science College, Inner Mongolia Normal University, Inner Mongolia Key Laboratory of Green Catalysis, Hohhot, Inner Mongolia 010022, PR
China
f
Beijing Institute of Nanoenergy and Nanosystems, Chinese Academy of Sciences, Beijing, 100083, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: Simultaneous recovery of hydrogen and sulfur resources from acidic gas hydrogen sulfide (H2S), whose content
Hydrogen resource may reach 20–100 % after hydrodesulfurization in oil refineries, is increasingly attractive. Nowadays, the
Sulfur recover widespread utilization of H2S decomposition technology is mainly restricted by the low hydrogen and sulfur
H2S yield, i.e. 20 % at 800 °C. In this work, for the first time a novel non-sulfide based LaFexAl12-xO19 hexaaluminate
Active phase
catalysts was used for high (50 %) hydrogen yield at 800 °C. At the same time, the active phases and reaction
Catalytic mechanism
mechanism were confirmed by various characterization techniques. The XRD, Raman, XPS, Mössbauer and XAFS
results verified that the Fe species in the hexaaluminate structure, especially octahedral-coordinated Fe3+,
served as the main active phases. The reaction routes/mechanism for H2S decomposition were also verified by
density functional theory (DFT) method. H2S was firstly adsorbed on the active sites and then completed the
decomposition by direct dehydrogenation mechanism. Briefly, current study has minimized the challenges for
H2S decomposition at industrial scale. It is anticipated that our findings will offer potential application of H2S
decomposing technology.

1. Introduction H2S decomposition [5]:

2H2S ↔ 2H2 + 1/4S8, ΔH =79.5 kJ/mol (1)


Hydrogen sulfide (H2S) is a flammable, corrosive and very poiso-
nous gas, which could be harmful to living organisms even at low In recent years, various methods have been explored and tested on
concentrations. However, except for relatively small amounts, i.e. H2S decomposition, including thermolysis, electrolysis, photolysis,
0.001–16 %, produced by natural sources, H2S is widely generated in plasmolysis and many their variants [6–10]. Among these, the most
chemical industries [1]. From hydrodesulfurization in oil refineries, the promising process is thermal decomposition. However, the H2S de-
content of H2S may reach 20–100 % [2]. The famous Claus process is composition reaction is highly endothermic and the equilibrium con-
extensively applied to H2S treatment, during which H2S is converted version is low even at high temperatures (only 20 % H2 yield at
into sulfur and water [3]. Meanwhile, it is worthy pointing out that 1000 °C) [11]. Therefore, certain amounts of studies have been ongoing
hydrogen (H2), as alternative energy source and important chemical to overcome and tackle the drawback by employing suitable catalysts to
feedstock, is needed due to its increasing demand [4]. Therefore, taking improve the H2S conversion and hydrogen yield.
seriously for both the sustainability of resource and the fulfillment of Various catalysts for H2S decomposition based on metal sulfides,
stringent environmental requirements, an attractive alternative could disulfides and oxides have been investigated. For metal sulfides, mainly
be the recovery of hydrogen and elemental sulfur simultaneously by focused on FeS, CuS and NiS, however, maximum H2 yield attained was


Corresponding author at: National Engineering Laboratory for VOCs Pollution Control Material & Technology, University of Chinese Academy of Sciences, Beijing,
101408, PR China.
E-mail address: xzhang@ucas.ac.cn (X. Zhang).

https://doi.org/10.1016/j.apcatb.2019.118354
Received 25 August 2019; Received in revised form 21 October 2019; Accepted 27 October 2019
Available online 04 November 2019
0926-3373/ © 2019 Elsevier B.V. All rights reserved.

Please cite this article as: Guoxia Jiang, et al., Applied Catalysis B: Environmental, https://doi.org/10.1016/j.apcatb.2019.118354
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

only 20 % at 770 °C [12–14]. Notably, metal disulfides (such as MoS2, 6, 8, 10 and 12) were prepared using the carbonate route. Typically, to
WS2, FeS2, CoS2 and NiS2) and metal oxides (mainly focused on per- prepare LaFe2Al10O19 (LaFe2, x = 2), La(NO3)3, Fe(NO3)3·9H2O, and Al
ovskites [15–17]) presented analogous results, maximally receiving 20 (NO3)3·9H2O were dissolved individually in 60 °C deionized water in
% H2 yield at 800 °C [16,18]. Thus, the decomposition activity of H2S proportion. Except aluminum nitrate, metal nitrates were mixed to-
was still low and unsatisfactory. And the relatively low H2 yield ap- gether firstly to form clear solutions, followed by adjusting the pH = 1
peared to be the mainly restricted parameter for the widespread utili- with nitric acid. Then, the aluminum nitrate solution was added into
zation of the decomposition catalysts and the H2S decomposition pro- the metal nitrates mixture. This mixed solution was poured under
cess. vigorous stirring into a saturated (NH4)2CO3 solution heated at 60 °C to
More importantly, the reaction processes for metal oxides catalysts form precursor precipitate. During the precipitation process, the pH
were partly based on “two-step process” [10]. The key point of this value of the solution was maintained between 7.5 and 8.0. After con-
mechanism was that metal oxides were sulfided to sulfides, which were tinuous stirring at 60 °C for 4 h and aging for 3 h, the precipitate was
the actual catalysts: filtered, washed, and then dried at 120 °C overnight. The obtained
precursor materials were calcined by step-wise in the temperature
MxOy + yH2S → MxSy + yH2O, (2)
range 500–1200 °C for 6 h under air condition. All prepared samples
And then, the “two-step process” associated with metal sulfides were named as LaFex, such as LaFe6 (LaFe6Al6O19).
could happen. A “lowest” metal sulfide MxSy has affinity with sulfur, Catalytic decomposition of H2S was tested in a vertical fixed bed
forming the “highest” sulfide MxSy+z and hydrogen. In their turn, the quartz reactor placed in an external heating furnace at atmospheric
“highest” sulfides can give back the corresponding “lowest” sulfide by pressure. In a typical experimental run, 0.5 g of catalysts (20–40 mesh)
thermal decomposition: was loaded between quartz wool in the center of the quartz tube. The
typical compositions of feed gas were 1000 ppm H2S/N2, and the total
MxSy + zH2S → MxSy+z + zH2, (3) flow rate of the reaction mixture was 200 mL/min with a gas hourly
MxSy+z → MxSy + zS0, (4) space velocity (GHSV) of 24000 h−1. The outlet gas stream from fur-
nace was first passed through a condenser to trap the sulfur gas in ef-
H2S → H2 + S0. (5) fluent stream, and then were analyzed via an online gas chromato-
graphy (GC) equipped with a thermal conductivity detector (TCD, with
While the main drawbacks of the two-step processes were the
Plot/Q columns) and a flame photometric detector (FPD, with GasPro
handling of large mass of sulfides, and the need to cool and re-heat
columns). And the carrier gas of GC was nitrogen (N2). Before entering
them because of the differences in operating temperatures of the two
TCD detector, the outlet gas stream was passed through a gas bubbler
steps. Meanwhile, due to the phases change, the real catalysts were the
containing saturated aqueous Cu (II) acetate solution to remove un-
formed intermediate sulfides but not oxides. Thus, metal oxides could
converted H2S.
not be recovered after reaction.
The conversion of hydrogen sulfide and the hydrogen yield were
Therefore, it is extremely desired to explore new type H2S decom-
calculated in accordance with equation:
position catalysts with new reaction routes, excellent catalytic activity
and high thermal stability by taking into account the fact that the de- (H2 S)in − (H2 S)out
H2 S conversion = × 100%
composition reaction is high temperature favorable thermodynamically (H2 S)in (6)
(almost higher than 800 °C).
Hexaaluminates consisting of alternatively stacked spinel blocks (H2 )out
H2 yield = × 100%
and mirror planes have been regarded probably as the most promising (H2 S)in (7)
catalytic materials for high-temperature applications due to their ex-
ceptional thermal stability in combination with extraordinary sub-
2.2. Catalyst characterization
stituting capability [19,20]. The general formula can be presented as
ABxAl12−xO19, A represents a mono-, di-, or trivalent large cation re-
The structure of the samples were investigated using powder X-ray
siding in the mirror plane, and the component B could be a transition-
metal (Mn, Fe, Co, Cu, Ni, etc.) or a noble metal ion (Ir, Ru, Pd, Rh), diffraction (XRD, PAN analytical X’Pert-Pro, Cu Kα radiation, λ
=1.5418 Å) operated at 40 kV and 40 mA in the 2θ range of 10-90° with
which can partially or even completely substitute Al crystallographic
5°/min scanning rate. The morphology and chemical composition of
sites [21]. Particularly, the stable spinel structure made sulfidation
prepared samples were examined using field emission scanning electron
reactions difficult, just like the results on perovskites [16,22]. There-
microscopy with energy dispersive X-ray spectroscopy (FESEM/EDX,
fore, the new reaction route possibility for H2S catalytic decomposition
JSM-6700 F), transmission electron microscopy and high-resolution
was provided.
transmission electron microscopy (TEM/HRTEM, JEM-2100 F, 200 kV).
In this study, the active Fe species and the remarkably outstanding
Raman spectra were performed at room temperature using a micro-
thermostability of hexaaluminates were combined. And Fe substituted
Raman spectrometer (HORIBA, France) with a 532 nm Ar+ laser as the
hexaaluminates catalysts LaFexAl12-xO19 (LaFex, x = 2, 4, 6, 8, 10 and
excitation source in a backscattering geometry. The laser power in-
12) were synthesized and applied for H2S decomposition reaction. The
cident on the catalysts was less than 10 mW. The time of acquisition
results shown that as high as 50% hydrogen yield was achieved, which
was varied according to the intensity of the Raman scattering.
has never been reported before. Subsequently, the main active phases
Surface compositions of the samples were investigated with X-ray
were considered to be the Fe species presented in the hexaaluminate
photoelectron spectroscopy (XPS, Thermo ESCALAB 250, USA) using a
structure, especially octahedral-coordinated Fe3+. In addition, it was
monochromatic Al Kα exciting radiation (1486.71 eV). The shift of the
verified that, in contrast to traditional two-step process, H2S was firstly
binding energy due to relative surface charging was corrected using the
adsorbed on the active sites and then completed the decomposition by
C1 s level at 284.8 eV as an internal standard. Near-surface atomic
direct dehydrogenation processes.
concentrations of the fresh and spent materials were calculated ac-
cording to peak areas and sensitivity factors.
2. Experimental The 57Fe Mössbauer spectra of the as-prepared materials were re-
corded using a Topologic 500A spectrometer and a proportional
2.1. Catalyst preparation and activity test counter at 80 K with a spectrometer working at constant acceleration
using a Rh57Co source. 57Fe Mössbauer spectral parameters such as the
Samples with nominal composition LaFexAl12-xO19 (LaFex, x = 2, 4, isomer shift (IS), the electric quadrupole splitting (QS) and the relative

2
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

spectral areas (A) of the different components of the absorption patterns durable experiments emerged in the first cycle and then vanished in the
were evaluated. All IS values were reported with respect to metallic second cycle? Therefore, it is strongly desired to elucidate the inter-
iron (α-Fe) at the measurement temperature. Data analysis was per- correlation between the catalysts structure, properties, active phase and
formed using a nonlinear least square fitting routine that models the the extraordinarily excellent catalytic performance as well as the cat-
spectra as a combination of singlets, quadruple doublets and magnetic alytic mechanism.
sextuplets based on a Lorentzian line shape profile.
The powder samples were prepared for X-ray Absorption Fine 3.2. The chemical state of Fe species in LaFex catalysts
Structure (XAFS) specimens with thickness appropriate for Fe K-edge
measurements. The XAFS measurements were carried out at the Beijing The X-ray powder diffraction (XRD) patterns of LaFex catalysts were
Synchrotron Radiation Facility (BSRF). Fe foil was used as a reference provided in Fig. 2a and b. As illustrated in Fig. 2a, typical diffraction
sample, and all spectra were measured in the transmission mode at peaks corresponding to a well crystallized magnetoplumbite (MP,
room temperature. JCPDS 36–1314, marked with β) phase were the dominant features of
LaFe2, LaFe4 and LaFe6 catalysts. In the case of LaFe8 catalysts, several
2.3. Theoretical calculation weak reflections associated with the impurities of α-Fe2O3 (JCPDS 84-
0310) and LaFeO3 (JCPDS 75-0439) were simultaneously detected.
The DFT calculations were conducted using CASTEP with a plane What’s more, the neat patterns were even predominated by the reflec-
wave basis set. The Perdew-Burke-Ernzerhof (PBE) functional with the tions related to α-Fe2O3 and LaFeO3 with Fe substitution increased
generalized gradient approximation (GGA) was applied to calculate the further to 10 and 12. Especially for LaFe12 catalysts, monophasic MP
exchange-correlation energy. The ion core and valence electron inter- peaks were not detected.
action was described by Vanderbilt-type ultrasoft pseudopotential. A The phase evolution of LaFex catalysts due to Fe substitution could
plane wave cut-off energy of 380 eV was used in the calculation. be clearly revealed from the 30–40° (2θ) range of the XRD patterns. As
Brillouin zone sampling was calculated using a Monkhorst-Pack grid shown in Fig. 2b, the peak positions corresponding to MP phase
with respect to the symmetry of the system. The convergence criteria (marked with β) continuously shifted to lower diffraction angles with
for configuration optimization were set to the tolerance for SCF, energy, the x increased from 2 to 10, which mainly resulted from the enlarge-
maximum force, with a maximum displacement of 1 × 10−6 eV/atom, ment of the cell parameters of MP hexaaluminate. In other words, the
2 × 10−5 eV/atom, 0.05 eV/Å and 2 × 10−3Å, respectively. iron content incorporated into the hexaaluminate lattice increased with
the Fe substitution. Furthermore, the intensity of peaks related to MP
3. Results and discussion phase gradually weakened as Fe substitution increased, indicating that
excessive substitution would essentially hindered the formation of
3.1. Catalytic performance of LaFex catalysts monophasic MP phase [23].
Raman spectroscopy is a powerful means for the conformation of
The catalytic performance of LaFex catalysts for H2S decomposition structural phase transitions, moreover, to determine phases and struc-
were displayed in Fig. 1a and b. As shown, the LaFe6 samples were the ture of multioxide systems. As shown in Fig. 2c, the LaFex catalysts with
most effective catalysts, which could present as high as 50 % H2 yield at x below 8 presented similar features. In the case of LaFe2 catalysts, the
800 °C. To the best of our knowledge, it was remarkably higher than the dominated broad band around 800 cm−1 was mainly originated from
previous results reported in literatures (Table S1). Especially, as an the motions of the tetrahedral and bipyramidal group in hexaaluminate
oxide catalyst, Fe-substituted hexaaluminate seems to be more active crystal structure [24]. However, the characteristic band gradually
than perovskites with Cr and/or Co occupying B-site, which has the best shifted to lower wavenumbers with the rise of Fe substitution to 8. The
known catalystic activity (20 % H2S conversion at 800 ℃) [16,17]. shifted modes in the 600–700 cm−1 region could ascribe to the
Besides, the two-cycle durable experiments (Fig. 1c) with 10 h for stretching modes of octahedral units in hexaaluminate structure, which
per cycle were performed over LaFe6 catalysts at 800 °C. And the results became rather obvious along with the substitution increased [24].
revealed that the LaFe6 catalysts exhibited excellent stability, which Hence, it indicated that the hexaaluminate structure with the Fe sub-
could proceed consecutively for two cycles (20 h) without notable de- stitution up to 8 could be maintained. Furthermore, it was evident that
activation. It seemed not appear to be a situation where the resulting the LaFe10 and LaFe12 catalysts exhibited distinctly different char-
produced sulfur covered the active sites and reduced activity. acteristics with hexaaluminate structure. The band at 217, 241, 286,
Particularly, some interesting phenomena were also observed. As 398 and 489 cm−1 were related to the presence of α-Fe2O3 [25].
shown in Fig. 1a, the H2 yield increased gradually with the rise of Fe Moreover, the emerging band at 432 and 608 cm−1 for LaFe12 catalysts
substitution from 2 to 6. Subsequently, however, the H2 yield drasti- belonged to the orthorhombic structure of LaFeO3 [26]. These results
cally decreased upon the growth of Fe substitution from 8 to 12. At the were well aligned with the above XRD analysis.
same time, unexpected behaviors were obtained in the case of H2S Fig. 3 displayed the transmission electron microscope (TEM) images
conversion in Fig. 1b. For LaFe2, LaFe4 and LaFe6 catalysts, the H2S of the LaFe6 samples. And the figures (Fig. 3a) verified that the LaFe6
conversion gradually increased with the rise of Fe substitution and samples had a typical hexxalluminate sheet structure, which was more
temperature. Whereas, the H2S conversion for LaFe8, LaFe10 and LaFe12 evident in high-resolution image (HRTEM) and selected-area electron
catalysts increased greatly (almost higher than 50 %), expressing an diffraction (SAED) characterizations. The lattice fringes in the HRTEM
opposite trend compared with H2 yield. Apart from these, in Fig. 1c, an image (Fig. 3b) and discrete electron diffraction points (insert image in
initially increased stage of H2 yield (almost 4 h) was observed for the Fig. 3c) clearly proved that the samples were highly crystallized. The
first cycle. However, the increasing stage vanished in the second cycle inter-planar spacing derived from SAED spectra was calculated to be
and the H2 yield was maintained at 50% throughout the second cycle. 0.56 nm, corresponding to the crystal face of (004) (JCPDS 36–1314).
57
Overall, according to the catalytic performance, several critical is- Fe Mössbauer spectroscopy was employed at 80 K to characterize
sues need to be explored. Such as why the LaFex catalysts with lower Fe the iron state in the as-prepared materials. The fitted hyperfine para-
substitution (x = 2, 4 and 6) could achieve higher H2 yield than those meters, namely, isomer shift (IS), quadrupole splitting (QS) and mag-
with higher Fe substitution (x = 8, 10 and 12)? Why the H2S conversion netic hyperfine field (H) were collected in Table 1. And the spectra
of LaFex catalysts were always higher than the corresponded H2 yield at fitting results were displayed in Fig. 4.
each specific temperature, especially for the LaFe8, LaFe10 and LaFe12 For LaFe2, LaFe4 and LaFe6 catalysts, it was clear that the H values
catalysts (almost lower than 30 % H2 yield with respect to higher than in Table 1 were all bellowing 520 KOe, testifying that monophasic MP
50% H2S conversion)? Why an initially increased stage of H2 yield on structure was presented. Namely, Fe species exist in the MP crystal

3
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

Fig. 1. Catalytic performance of LaFex catalysts. a H2 yield and b H2S conversion of the LaFex catalysts as a function of temperature in the feed gas of 1000 ppm H2S/
N2 (24000 h−1). c Durability experiment results with two cycles on LaFe6 catalysts at 800 °C in the feed gas of 1000 ppm H2S/N2 (24000 h−1).

lattice by replacing the Al sites [27–29]. In addition, the IS values for rise of Fe substitution to 6 (4.2 wt%, 9.7 wt% and 14.9 wt% for LaFe2,
LaFe2, LaFe4 and LaFe6 catalysts were 0.18–0.5 mm/s, indicating the LaFe4 and LaFe6 catalysts, respectively. Calculated based on the relative
unique presence of Fe3+ species [30–33]. It was verified by the Fe 2p X- area and actual Fe content listed in Table 2), demonstrating that more
ray photoelectron spectroscopy (XPS) analysis (Fig. S1a) as well. Fe ions entered into the hexaaluminate lattice by occupying the octa-
On the other hand, it has been widely accepted that there are three hedral sites of MP structure. It was extremely consistent with the
different kinds of crystallographic sites in the MP-type La-hex- Raman results that the band related to stretching modes of octahedral
aaluminate structure, which are tetrahedral (Al2), octahedral (Al1, Al4, units was gradually enhanced with the rise of Fe substitution.
Al3) and bipyramidal (Al5) sites [34]. As for the fitting results in Table 1 In the LaFe8 and LaFe10 catalysts cases, two new sextets with H
related to LaFe2, LaFe4 and LaFe6 catalysts, the IS < 0.3 mm/s illu- values higher than 520 KOe emerged, which were corresponded to
strated the existence of Fe3+ ions in tetrahedral (Th) coordination Fe3+ in α-Fe2O3 and LaFeO3 structure, respectively. The higher hy-
[35,36]. More detailed, when IS < 0.3 mm/s, the highly distorted tri- perfine field values (533–536 KOe) was regarded as the presence of
gonal bipyramidal (Tr) Al5 sites were expected to show higher QS value Fe3+ in α-Fe2O3, while Fe3+ in LaFeO3 resulted in H values locating at
than that observed for the symmetric tetrahedral Al2 sites in the spinel 524–525 KOe, taking into account the fact that the H values of Fe3+ in
block. Subsequently, the 0.3 mm/s < IS < 0.5 mm/s revealed the α-Fe2O3 were greater than that of Fe3+ in LaFeO3 [27–29]. However,
presence of Fe3+ ions in octahedral (Oh) coordination [37–39]. In ad- the results suggested that the MP structure vanished in LaFe12 catalysts,
dition, the remaining sublattices of LaFe6 catalysts associated with Fe3+ which was well aligned with the XRD and Raman analysis. Importantly,
ions in octahedral sites (Al1, Al3, and Al4) can be further separated by states with IS > 0.8 mm/s are assigned to Fe2+ ions (0.7 wt%),which
the value of H. The H with the largest value (515.6 kOe) was compar- also proved by the Fe 2p XPS analysis in Fig. S1a.
able to that of the octahedral Al3 sites in MP-hexaaluminate. Mean-
while, the sextets with the lowest hyperfine fields (377.1 kOe) were
3.3. Active phases of the reaction
attributed to the octahedral Al4 sites [28]. Moreover, it is significant to
emphasize that the Fe3+ (Oh) content increased progressively with the
Although the above analysis revealed the phase compositions of the

4
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

Fig. 2. Structural characterization of LaFex catalysts. a X-ray diffraction patterns of LaFex catalysts. b The 30–40 degree (2 theta) range of the X-ray diffraction
patterns LaFex catalysts. c Raman spectra of LaFex catalysts.

catalysts, however, the actual active phases and reaction mechanism during the first cycle. Consequently, the catalytic performance should
were still uncertain. be improved significantly for the second cycle at each or specific re-
Therefore, in order to get insight to the feasibility of traditional two- action temperature. However, in terms of H2S conversion and H2 yields,
step process over the LaFex catalysts, two cycle H2 yield experiments the used catalysts exhibited very similar results compared with those of
versus reaction temperature were carried out over the LaFe6 catalysts fresh catalysts (Fig. 5a). Importantly, the used catalysts under-
under the employed reaction conditions. And the catalytic perfor- performed the fresh catalysts, which was contrary to the expectation. It
mances were exhibited in Fig. 5a. was suspected that different reaction mechanisms occurred among the
From the standpoint of traditional two-step process, it seemed lo- LaFex catalysts.
gical to expect that the sulfurized LaFe6 active phases should be formed To clarify these phenomena, various characterizations were

Fig. 3. Morphology characterization of LaFe6 catalysts. a TEM images, b and c HRTEM images of LaFe6 catalysts. The insert image in c was SAED pattern of LaFe6
catalysts.

5
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

Table 1
57
Fe Mössbauer spectra results of LaFex catalysts at 80 K.
Catalyst IS (mm/s)a QS (mm/s)b H (KOe)c A (%)d Mass ratio (%)e Site assignment

LaFe2 0.215 0.677 50.0 6.9 doublet MP, Fe3+(Thf), Al2


0.271 2.328 20.1 2.8 doublet MP, Fe3+(Trg), Al5
0.490 0.640 30.0 4.2 doublet MP, Fe3+(Ohh)
LaFe4 0.303 0.390 doublet PARA
0.220 0.636 426.6 37.2 8.0 sextet MP, Fe3+(Th), Al2
0.270 2.430 324.0 17.7 3.8 sextet MP, Fe3+(Tr), Al5
0.429 0.620 505.0 45.1 9.8 sextet MP, Fe3+(Oh)
LaFe6 0.360 0 doublet PARA
0.214 0.634 463.2 27.0 6.5 sextet MP, Fe3+(Th), Al2
0.280 2.480 396.0 11.2 2.7 sextet MP, Fe3+(Tr), Al5
0.473 0.372 377.1 6.4 1.5 sextet MP, Fe3+(Oh), Al4
0.482 0.410 496.1 25.0 6.0 sextet MP, Fe3+(Oh), Al1
0.505 0.212 515.6 30.4 7.4 sextet MP, Fe3+(Oh), Al3
LaFe8 0.300 −0.100 536.0 6.3 1.7 sextet α-Fe2O3
0.390 0.180 525.0 12.8 3.4 sextet LaFeO3
0.210 0.630 470.0 29.0 7.8 sextet MP, Fe3+(Th), Al2
0.270 2.660 405.0 42.0 11.3 sextet MP, Fe3+(Tr), Al5
0.465 0.494 490.6 9.6 2.6 sextet MP, Fe3+(Oh)
LaFe10 0.369 0.180 533.0 16.0 4.8 sextet α-Fe2O3
0.352 0.148 524.0 30.0 8.9 sextet LaFeO3
0.261 0.324 517.0 29.0 8.7 sextet MP, Fe3+(Th), Al2
0.280 2.364 512.1 18.0 5.4 sextet MP, Fe3+(Tr), Al5
0.460 0.510 444.0 7.0 2.1 sextet MP, Fe3+(Oh)
LaFe12 0.304 0.155 534.0 38.0 13.6 sextet α-Fe2O3
0.347 0.123 524.1 60.0 21.4 sextet LaFeO3
1.020 −0.260 530.8 2.0 0.7 sextet Fe2+

a
Isomer shift.
b
Quadrupole splitting.
c
Hyperfine field.
d
Relative area.
e
The actual content of each specific Fe species.
f
Tetrahedral.
g
Trigonal bipyramidal.
h
Octahedral.

conducted on fresh, used and reused LaFe6 catalysts. The XRD analysis scattering of second-rate nearby metal ions [45]. As known, the radial
(Fig. S2a) revealed that the MP structure of LaFe6 catalysts were well distribution function may indicate the distance between atom level and
maintained even after two-cycle reactions. More importantly, no peaks central absorbing atom. The first coordinate level in catalyst samples
related to sulfides and sulfates appeared. Whereas, the sulfur species implied that the coordination of Fe-O was changed after reaction. It was
were detected by XPS and energy dispersive X-ray spectroscopy (EDX), shown that the distance between the scattering center (Fe ions) and the
and the actual surface elemental contents were listed in Table 2. The O2− ions was shorten after used, due mainly to the reduction of a few
EDX mapping (Fig. S3) shown that sulfur species distributed uniformly Fe3+ to Fe2+. Besides, as stated in literature, the Fe-S bond distance
in all the catalysts. Besides, it was considered that the peaks located at was generally considered to be located at 2.23 Å [46]. While, there was
160.88 eV, 162.24 eV and 168 eV in S 2p XPS (Fig. S2d) were ascribed no discernible peak in this position.
to S2-x2− species, S2- species and SO42− species, respectively [40–42]. Hence, the traditional two-step mechanism may not favorable to
Moreover, further reaction could only cause a slight increase in sulfur explain the results in this work by considering that the MP structure of
content (Table 2), which increased from 2.5 wt% for the used catalyst to LaFe6 catalysts were well maintained even proceeded two cycles
4.3 wt% for the reused catalyst. (Figs. 5d and S2a) and only a tiny proportion of Fe was sulfurized to
X-ray absorption fine structure spectroscopy (XAFS) was further sulfide and sulfate species (Table 2). As from the perspective of tradi-
applied to explore the phase compositions of the fresh used and reused tional two-step mechanism, the catalysts should be sulfurized easily and
catalysts. The normalized Fe K-edge X-ray absorption spectra, X-ray abundantly, and more sulfur species were beneficial to the reaction.
absorption near edge structure (XANES) region and radial distribution Therefore, the sulfides in traditional two-step mechanism seemed not
functions were displayed in Fig. 5b–d. The Fe-K absorbing edge shifted play the main role of actual active phases.
to a lower energy position (Fig. 5c), indicating the decrease of valent of Consequently, it was likely acceptable to conclude that the Fe spe-
Fe ions [43,44]. These results were in line with Fe 2p XPS analysis in cies presented in the stable hexaaluminate structure rather than sulfides
Fig. S2b, in which the relative amount of Fe2+ species increased slightly played the role of actual active phases in the reaction. Particularly, it
after two cycle reaction (area percentage 15% in used catalyst and 20 % was extremely important to observe that the H2 yield in Fig. 1a over
in reused catalyst). fresh LaFex catalysts at 800 °C were positively correlated with the
It was astonishing to be observed that the spectra profiles in Fig. 5d content variation trend (Table 1) of Fe3+ ions in octahedral (Oh) co-
for all the catalysts were almost consistent with each other, revealing ordination within hexaaluminate structure. Hence, the Fe3+ (Oh)
that the structure were stable during the reaction and no more phases maybe among the major active site species contributing to the excellent
emerged. There were two strong amplitude peaks between 1.0 Å and catalytic performance.
4.0 Å. The first peak at 1.50 Å was related to the first coordinate level of
ferrite, which was corresponding to Fe-O coordinate peak caused by the
scattering of nearby O2− ions [45]. The second peak appeared at 2.6 Å 3.4. Reaction mechanism over hexaaluminate catalysts
were ascribed to second coordinate level, which was caused by the
Based on theradical reaction mechanism [47] and the analysis of

6
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

Fig. 4. 57Fe Mössbauer spectra fitting results of LaFex catalysts. The solid lines show the curve fits to the raw data. Average uncertainties in Mössbauer para-
meters: ± 0.03, Fe fraction; ± 2 KOe, hyperfine magnetic field; ± 0.03 mm/s, isomer shift; ± 0.05 mm/s quadrupole splitting.

Table 2 well known that the crystal structure of hexaaluminates is very com-
The chemical composition of surface elements in LaFex catalysts. plicated, consisting of alternatively stacked spinel blocks and mirror
Catalyst Atomic ratio (%) Oβ/Oα Mass ratio (%)
planes. The analysis in Table 1 proved that Fe mainly entered the spinel
blocks substituting the lattice sites of Al in Fe-substituted hex-
La Fe O Al S Fe S aaluminate samples. Thus, the structure of FeAl2O4 spinel was adopted
to simulate hexaaluminates and simplify the calculation. Because hy-
LaFe2 3.1 6.4 57.1 33.2 0.51 13.9
drogen generation was reduced with LaFe8, LaFe10 and LaFe12 catalysts,
LaFe4 3.9 10.9 57.6 27.4 0.56 21.6
LaFe6 4.7 12.7 59.5 22.9 0.61 24.2 related calculation on α-Fe2O3 and LaFeO3 were also conducted for
LaFe8 6.1 15.2 60.7 17.8 0.77 26.9 reference and comparison. The related slab models were displayed in
LaFe10 7.5 17.9 64.4 10.0 1.46 29.9 Fig. S4.
LaFe12 10.3 24.6 65.1 1.61 35.7
To explore the reaction pathways, the stable co-adsorption config-
LaFe6 used 5.2 10.4 55.1 26.8 2.4 0.56 19.4 2.5
LaFe6 reused 5.0 8.1 53.6 29.2 4.0 0.54 15.4 4.3 urations of SH and H or S and 2H served as the reactants and inter-
mediates. All structure and bond length information for the transition
states (TS) and products were displayed in Figs. S5 and S6. And sche-
active phases, a probable direct dehydrogenation reaction pathway was matic potential energy profiles were presented in Fig. 6 to illustrate the
tentatively proposed as following: energy changes for the whole process of H2S interacting with the sur-
face. As the research about H2S interacting with the α-Fe2O3 (001)
-Fe- + H-SH → -Fe-SH + H, (8) surface has been already clear, the results of potential energy profiles
-Fe-SH → -Fe-S + H, (9) were used for comparison [48,49].
In Figs. S5 and S6, H2S firstly approached and adsorbed on the Fe
-Fe-S → -Fe- + S, (10) top site, parallel to the surface of LaFeO3 (010) and FeAl2O4 (100),
H + H → H2. (11) eventually, analogous to the direct dehydrogenation process on the α-
Fe2O3 (001) surface [49]. After the fracture of H–S bond in absorbed
H2S firstly adsorbed on the catalysts surface of Fe (Oh) sites, and
3+
H2S and OeH bond formation, the first dehydrogenation process (H2S
then complete the decomposition by direct dehydrogenation processes. → SH + H) occurred, with the energy barriers of 25.9 and 7.7 kJ/mol
The density functional theory (DFT) method within a periodic slab on p (2 × 2) super-cells of LaFeO3 (010) and FeAl2O4 (100) surface,
model was employed to validate the new proposed mechanism. It is respectively. These values were much smaller than the corresponding

7
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

Fig. 5. Catalytic performance and characterization of LaFe6 catalysts. a H2 yield (line and symbol) and H2S conversion (column) of the fresh and used LaFe6 catalysts
as a function of temperature in the feed gas of 1000 ppm H2S/N2 (24000 h−1). b Normalized Fe K-edge X-ray absorption spectra of the fresh and used LaFe6 catalysts.
c XANES region of the normalized Fe K-edge X-ray absorption spectra of the fresh and used LaFe6 catalysts. d Radial distribution functions around Fe atom of samples
of the fresh and used LaFe6 catalysts.

hypothetical route was feasible and the reaction could occur on the
FeAl2O4 (100) surface easiest.
Differences of the catalytic activities were perhaps mainly derived
from the electronic and structural properties of the catalysts. The Fe-O
atoms on the catalysts surface were used to cleave H2S electronically,
and the O atom acted as a proton acceptor, which meant that the more
negative charge the O atom had, the higher ability to attract protons
[50]. Calculation of Mulliken charge (Table S2) confirmed that the
absolute charge number of O on the three surfaces followed the order:
FeAl2O4 > LaFeO3 > Fe2O3, consistent with the yield performance.
Structurally, differences between the H–S-H angle in H2S (the angle of
H-S-H in the free H2S is 92.1°) and the O-Fe-O angle on the surface for
adsorption and cleavage significantly have an effect on the energy
barrier [51]. Greater differences in angle lead to higher energy barrier
to be overcome [52]. The angle of O-Fe-O on the FeAl2O4 surface had a
moderate difference with the angle of H-S-H in the free H2S, resulting in
the decomposition happen possibly.
Fig. 6. Energy profile of the H2S decomposition process. Finally, the catalytic performance can be explained reasonably by
the direct dehydrogenation mechanism. The catalysts with lower Fe
substitution (LaFe2, LaFe4 and LaFe6) could produce higher H2 yield
α-Fe2O3 (001) (72.0 kJ/mol) [49]. Another dehydrogenation process
(Fig. 1a) were due mainly to the higher content of Fe3+ (Oh) sites. And
(SH → S + H) proceeded, with the energy barriers of 52.8 and 17.3 kJ/
the slightly decreased catalytic performance in Fig. 5a for the second
mol of LaFeO3 (010) and FeAl2O4 (100) surface, respectively. These
cycle might resulted from the formation of little amount sulfur species.
values were also much lower than the corresponding α-Fe2O3 (001)
Besides, it was worth mentioning that the O 1s XPS analysis re-
(130 kJ/mol) [49]. Thus, the FeAl2O4 (100) surface presented the
vealed that oxygen species were presented on the catalyst surface and
lowest energy barriers in the whole process, indicating that the
consumed during the reaction (Fig. S2c). As known, the peak at ∼531.5

8
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

eV was related to surface adsorbed oxygen species (donated as Oα), and References
the band at ∼529.5 eV was ascribed to lattice oxygen species (donated
as Oβ) [53]. And the Oβ/Oα in Fig. S2c decreased from 0.61 to 0.54 after [1] M.S. Shah, M. Tsapatsis, J.I. Siepmann, Chem. Rev. 117 (2017) 9755–9803.
two cycle reaction, indicating that the lattice oxygen was involved in [2] A.K. Gupta, S. Ibrahim, A. Al Shoaibi, Prog. Energy Combust. Sci. 54 (2016) 65–92.
[3] A. Shaver, M. ElKhateeb, A.M. Lebuis, Angew. Chemie Int. Ed. English 35 (1996)
the reaction. 2362–2363.
The O 1s XPS spectra for all the LaFex catalysts were also provided [4] C. Duan, R.J. Kee, H. Zhu, C. Karakaya, Y. Chen, S. Ricote, A. Jarry, E.J. Crumlin,
in Fig. S1b. It indicated that lattice oxygen species went up with the D. Hook, R. Braun, Nature 557 (2018) 217.
[5] X. Zong, J.F. Han, B. Seger, H.J. Chen, G.Q. Lu, C. Li, L.Z. Wang, Angew. Chemie Int.
increase of Fe substitution, and the Oβ/Oα rise sharply especially when Ed. English 53 (2014) 4399–4403.
iron oxides appeared (LaFe8, LaFe10 and LaFe12 catalysts). As known, [6] L. Zhao, Y. Wang, L. Jin, M.L. Qin, X. Li, A.J. Wang, C.S. Song, Y.K. Hu, Green
lattice oxygen were nucleophilic reagents and were usually responsible Chem. 15 (2013) 1509–1513.
[7] K.M. Kwok, S.W.D. Ong, L. Chen, H.C. Zeng, ACS Catal. 8 (2017) 714–724.
for selective oxidation reactions. In conjunction with the phenomenon [8] D. Ipsakis, T. Kraia, G.E. Marnellos, M. Ouzounidou, S. Voutetakis, R. Dittmeyer,
that Fe2+ emerged after the reaction, H2S selective catalytic oxidation A. Dubbe, K. Haas-Santo, M. Konsolakis, H.E. Figen, N.O. Güldal, S.Z. Baykara, Int.
perhaps occurred simultaneously [54], as a Fe3+ redox cycle was in- J. Hydrogen Energ. 40 (2015) 7530–7538.
[9] T.Y. Cong, A. Raj, J. Chanaphet, S. Mohammed, S. Ibrahim, A. Al Shoaibi, Int. J.
volved in the H2S selective oxidation process. Moreover, the sequence
Hydrogen Energ. 41 (2016) 6662–6675.
of H2S conversion were extremely consistent with that of Oβ/Oα. Thus, [10] A.P. Reverberi, J.J. Klemeš, P.S. Varbanov, B. Fabiano, J. Clean. Prod. 136 (2016)
the occurrence of H2S selective catalytic oxidation associated with the 72–80.
sulfidation are responsible for the higher H2S conversion for LaFe8, [11] F. Faraji, I. Safarik, O.P. Strausz, E. Yildirim, M.E. Torres, Int. J. Hydrogen Energ. 23
(1998) 451–456.
LaFe10 and LaFe12 catalysts (Fig. 1b) and the initially increased stage of [12] A. Bishara, O.A. Salman, N. Khraishi, A. Marafi, Int. J. Hydrogen Energ. 12 (1987)
H2 yield in the first time experiment of catalysts durability (Fig. 1c). 679–685.
[13] O.A. Salman, A. Bishara, A. Marafi, Energy 12 (1987) 1227–1232.
[14] V.E. Kaloidas, N.G. Papayannakos, Int. J. Hydrogen Energ. 12 (1987) 403–409.
[15] N.O. Guldal, H.E. Figen, S.Z. Baykara, Int. J. Hydrogen Energ. 40 (2015)
4. Conclusion 7452–7458.
[16] N.O. Guldal, H.E. Figen, S.Z. Baykara, Chem. Eng. J. 313 (2017) 1354–1363.
Synergistic recovery of hydrogen and sulfur resources were con- [17] N.O. Guldal, H.E. Figen, S.Z. Baykara, Int. J. Hydrogen Energ. 43 (2018)
1038–1046.
ducted over catalytic decomposition of acidic gas H2S on non-sulfide [18] T. Chivers, J.B. Hyne, C. Lau, Int. J. Hydrogen Energ. 5 (1980) 499–506.
based LaFex hexaaluminate catalysts. The LaFe6 catalysts could present [19] A.J. Zarur, J.Y. Ying, Nature 403 (2000) 65–67.
more than 50% H2 yield and proceed consecutively for two cycles [20] J.G. McCarty, Nature 403 (2000) 35–36.
[21] M. Tian, X.D. Wang, T. Zhang, Catal. Sci. Technol. 6 (2016) 1984–2004.
(20 h) without notable deactivation. The outstanding catalytic perfor-
[22] L.Z. Zhang, J.-M.M. Millet, U.S. Ozkan, J. Mol. Catal. A-Chem. 309 (2009) 63–70.
mance in this work provided possibility for the technological innova- [23] G. Groppi, C. Cristiani, P. Forzatti, J. Catal. 168 (1997) 95–103.
tion in chemical industries involved acidic gas H2S produced. [24] J. Kreisel, G. Lucazeau, H. Vincent, J. Solid State Chem. 137 (1998) 127–137.
[25] D.L.A. De Faria, S. Venâncio Silva, M.T. De Oliveira, J. Raman Spectrosc. 28 (1997)
Compared with conventional Claus process, hydrogen in H2S is no
873–878.
longer wasted as H2O, but can be converted into valuable hydrogen [26] E. Traversa, P. Nunziante, L. Sangaletti, B. Allieri, L.E. Depero, H. Aono, Y. Sadaoka,
resource and energy. J. Am. Ceram. Soc. 83 (2000) 1087–1092.
Additionally, it was considered that the Fe species presented in the [27] Y. Zhang, X.D. Wang, Y.Y. Zhu, X. Liu, T. Zhang, J. Phys. Chem. C 118 (2014)
10792–10804.
stable hexaaluminate structure, especially the Fe3+ (Oh), rather than [28] Y.Y. Zhu, X.D. Wang, A.Q. Wang, G.T. Wu, J.H. Wang, T. Zhang, J. Catal. 283
the sulfides were the actual active phases in the reaction. And the tra- (2011) 149–160.
ditional two-step mechanism was not involved in the reaction. [29] Y.Y. Zhu, X.D. Wang, G.T. Wu, Y.Q. Huang, Y. Zhang, J.H. Wang, T. Zhang, J. Phys.
Chem. C 116 (2011) 671–680.
Significantly, a probable reaction pathway was tentatively proposed [30] P. Sazama, B. Wichterlová, E. Tábor, P. Šťastný, N.K. Sathu, Z. Sobalík, J. Dědeček,
and verified by DFT method. H2S firstly adsorbed on the catalysts Š. Sklenák, P. Klein, A. Vondrová, J. Catal. 312 (2014) 123–138.
surface of Fe3+ (Oh) sites, and then completed the decomposition by [31] G. Fierro, G. Moretti, G. Ferraris, G.B. Andreozzi, Appl. Catal. B: Environ. 102
(2011) 215–223.
direct dehydrogenation processes. It was revealed that the synthesized [32] P. Boroń, L. Chmielarz, J. Gurgul, K. Łątka, T. Shishido, J.-M. Krafft, S. Dzwigaj,
hexaaluminate catalysts presented the lowest energy barriers in the Appl. Catal. B: Environ. 138–139 (2013) 434–445.
whole process. Moreover, the H2S selective catalytic oxidation reaction [33] C.H. Choi, W.S. Choi, O. Kasian, A.K. Mechler, M.T. Sougrati, S. Bruller,
K. Strickland, Q.Y. Jia, S. Mukerjee, K.J.J. Mayrhofer, F. Jaouen, Angew. Chemie
could occur simultaneously, and the lattice oxygen participated in the
Int. Ed. English 56 (2017) 8809–8812.
reaction. The generated little sulfide and sulfate species were re- [34] E. Tronc, F. Laville, M. Gasperin, A.M. Lejus, D. Vivien, J. Solid State Chem. 81
sponsible for the decrease of catalytic activities. (1989) 192–202.
[35] G. Ung, J. Rittle, M. Soleilhavoup, G. Bertrand, J.C. Peters, Angew. Chemie Int. Ed.
English 53 (2014) 8427–8431.
[36] M.M. Rodriguez, B.D. Stubbert, C.C. Scarborough, W.W. Brennessel, E. Bill,
Declaration of Competing Interest P.L. Holland, Angew. Chemie Int. Ed. English 51 (2012) 8247–8250.
[37] P. Sazama, N.K. Sathu, E. Tabor, B. Wichterlová, Š. Sklenák, Z. Sobalík, J. Catal. 299
The authors declare that they have no known competing financial (2013) 188–203.
[38] M.M. Khusniyarov, T. Weyhermuller, E. Bill, K. Wieghardt, Angew. Chemie Int. Ed.
interests or personal relationships that could have appeared to influ- English 47 (2008) 1228–1231.
ence the work reported in this paper. [39] I. Yamada, K. Tsuchida, K. Ohgushi, N. Hayashi, J. Kim, N. Tsuji, R. Takahashi,
M. Matsushita, N. Nishiyama, T. Inoue, T. Irifune, K. Kato, M. Takata, M. Takano,
Angew. Chemie Int. Ed. English 50 (2011) 6579–6582.
[40] Z. Guo, X.W. Wang, Angew. Chemie Int. Ed. English 57 (2018) 5898–5902.
Acknowledgements [41] T. Jurca, M.J. Moody, A. Henning, J.D. Emery, B.H. Wang, J.M. Tan, T.L. Lohr,
L.J. Lauhon, T.J. Marks, Angew. Chemie Int. Ed. English 56 (2017) 4991–4995.
This work was financially supported by the National Natural Science [42] Y.D. Shao, Z. Guo, H. Li, Y.T. Su, X.W. Wang, Angew. Chemie Int. Ed. English 56
(2017) 3226–3231.
Foundation (21976177, 21976176, 21577158, 21507148), the [43] W.C.M. Gomes, D.M.A. Melo, P.M. Pimentel, E.P. Marinho, M.A.F. Melo, R.S. Nasar,
National Key R&D Program of China (2018YFA0209302) and the Chem. Phys. 490 (2017) 67–74.
University of Chinese Academy of Sciences. [44] M. Soloviov, A.K. Das, M. Meuwly, Angew. Chemie Int. Ed. English 55 (2016)
10126–10130.
[45] L.X. Wang, J. Zhang, Q.T. Zhang, N.C. Xu, J. Song, J. Magn. Magn. Mater. 377
(2015) 362–367.
Appendix A. Supplementary data [46] A. Matamoros-Veloza, O. Cespedes, B.R.G. Johnson, T.M. Stawski, U. Terranova,
N.H. de Leeuw, L.G. Benning, Nat. Commun. 9 (2018).
Supplementary material related to this article can be found, in the [47] A.M. El-Melih, L. Iovine, A. Al Shoaibi, A.K. Gupta, Int. J. Hydrogen Energ. 42
(2017) 4764–4773.
online version, at doi:https://doi.org/10.1016/j.apcatb.2019.118354.

9
G. Jiang, et al. Applied Catalysis B: Environmental xxx (xxxx) xxxx

[48] J.J. Song, X.Q. Niu, L.X. Ling, B.J. Wang, Fuel Process. Technol. 115 (2013) 26–33. [51] D.C. Look, I.J. Lowe, J.A. Northby, J. Chem. Phys. 44 (1966) 3441–3452.
[49] L.X. Ling, J.J. Song, S.P. Zhao, R.G. Zhang, B.J. Wang, RSC Adv. 4 (2014) [52] R.E. Johnson Jr., R.H. Dettre, J. Phys. Chem. 68 (1964) 1744–1750.
22411–22418. [53] C.N.R. Rao, V. Vijayakrishnan, A.K. Santra, M.W.J. Prins, Angew. Chemie Int. Ed.
[50] M. Haranczyk, R. Bachorz, J. Rak, M. Gutowski, D. Radisic, S.T. Stokes, J.M. Nilles, English 31 (1992) 1062–1064.
K.H. Bowen, J. Phys. Chem. B 107 (2003) 7889–7895. [54] X.R. Ren, L.P. Chang, F. Li, K.C. Xie, Fuel 89 (2010) 883–887.

10

You might also like