You are on page 1of 8

Article

Cite This: Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX pubs.acs.org/IECR

Low-Temperature H2S Removal from Gas Streams over γ‑FeOOH,


γ‑Fe2O3, and α‑Fe2O3: Effects of the Hydroxyl Group, Defect, and
Specific Surface Area
Yanning Cao,† Xiaohai Zheng,† Zhongjie Du,† Lijuan Shen,† Ying Zheng,*,‡ Chaktong Au,†
and Lilong Jiang*,†

National Engineering Research Center of Chemical Fertilizer Catalyst, Fuzhou University, Fuzhou, Fujian 350002, P. R. China

Department of Chemical and Biochemical Engineering, Western University, 1151 Richmond Street, London, Ontario N6A 3K7,
Canada
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Samples of γ-FeOOH, γ-Fe2O3, and α-Fe2O3


were prepared by calcining a γ-FeOOH precursor at selected
Downloaded via LMU MUENCHEN on October 20, 2019 at 07:35:26 (UTC).

temperatures and tested for low-temperature H2S removal.


The prepared materials were characterized by powder X-ray
diffraction, N2 adsorption−desorption measurement, trans-
mission electron microscopy, and in situ Fourier transform
infrared spectroscopy of pyridine adsorption. The results
suggested that the oxygen vacancies, specific surface areas, and
surface hydroxyl groups of the iron oxides have significant
influence on the desulfurization performance at room
temperature. In H2S adsorption experiments, the sample
calcined at 250 °C (FE-250) showed the highest performance, reaching 53% of the maximum sulfur capacity, and
desulfurization precision at 25 °C.

1. INTRODUCTION Gong et al.25 prepared α-FeOOH and γ-FeOOH by adding


Hydrogen sulfide (H2S) is predominantly released from ammonium bicarbonate and sodium carbonates as precipitants,
agricultural sewage, coal gasification, and natural gas refineries. respectively. With larger specific surface area, the former
It is harmful to the environment and is difficult to get rid of showed higher desulfurization activity than the latter. Huang et
because it is poisonous to most catalysts.1−4 It is of great al.26 reported that the presence of oxygen vacancies is an
importance to develop efficient technologies for the removal of important factor in H2S removal by γ-Fe2O3 because the
H2S, not only for environment protection but also for the vacancies facilitate the diffusion of ions in the lattice and thus
efficacy of a large number of catalytic processes. Various promote the sulfur capture ability. In general, the desulfuriza-
methods including condensation, adsorption, and oxidation tion performance of sorbents can be affected by many factors.
have been developed for H2S removal.5−9 The use of sorbents Our earlier works demonstrated that the high desulfurization
is popular because it is reliable, environmentally benign, and activities of FeOOH can be attributed to its large specific
workable in a wide range of temperatures.10−13 Over the past surface area, as well as the ample presence of surface oxygen
decade, metal-oxide-based materials have been investigated for vacancies and OH groups.27−29 The work reported herein aims
H2S adsorption, but the results were not satisfactory.14,15 For to investigate the combined effects of the factors mentioned
example, zinc-based materials were prepared as desulfurization above.
sorbents because of their high sulfur capacity and favorable In the present study, we prepared γ-FeOOH, γ-Fe2O3, and
sulfidation thermodynamics.16−19 However, efficient desulfur- α-Fe2O3 using γ-FeOOH as precursor. Through the control of
ization can only be achieved at high temperatures (∼300 °C). calcination temperature, we regulated the specific surface area,
The previously published works indicated that the sorbents amount of oxygen vacancies, and surface OH groups. The
based on transition-metal oxides became effective for H2S physicochemical properties of the as-prepared sorbent
removal only at high temperatures, none of which can perform materials were characterized using various techniques, and
well at low temperatures due to the limitation of active sites the desulfurization experiments were carried out at room
and low diffusion rates of sulfur species.20−22 Nonetheless, temperature. We investigated the effects of oxygen vacancies,
increasing attention has been devoted to iron-based materials
for low-temperature desulfurization because the materials are Received: June 26, 2019
inexpensive and can easily be regenerated.23,24 Iron-based Revised: August 26, 2019
sorbents mainly include iron oxides (e.g., α-Fe2O3, γ-Fe2O3, Accepted: October 1, 2019
etc.) and iron oxyhydroxides (e.g., α-FeOOH, γ-FeOOH, etc.). Published: October 1, 2019

© XXXX American Chemical Society A DOI: 10.1021/acs.iecr.9b03430


Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

specific surface areas, and surface OH groups on H2S removal as the H2S breakthrough capacity (X), which was calculated by
in detail. The findings of the present study provide new the equation28
insights that are beneficial for the design of iron-based 32 × V × C
adsorbents for efficient H2S removal at low temperatures. X= × 100%
(22.4 × G)(1‐C)
2. EXPERIMENTAL SECTION In the equation, 32 (g mol−1) is the molar mass of sulfur, 22.4
2.1. Preparation of Sorbent Materials. γ-FeOOH was (L mol−1) is the gas molar volume at the standard state, V (L)
prepared by a modified precipitation method. In a typical stands for the stream volume and is measured by a wet-gas
procedure, an aqueous solution of FeSO4·7H2O (0.1 M) was flowmeter located at the exit of the reactor, G (g) is the weight
premade in a flask. Then, an aqueous solution of ethyl- of the adsorbent, and C (%) is the H2S concentration in the
enediaminetetraacetic acid (EDTA) (0.1 mol L−1) at a Fe2+/ gas stream.
EDTA molar ratio of 1:0.03 was added. Subsequently, a NH3· Unlike the test for H2S breakthrough capacity, the evaluation
H2O solution (6.0 M) was added at a constant rate into the of desulfurization precision was conducted using a standard gas
FeSO4 solution with vigorous stirring under the condition of of 0.5% H2S and 99.5% N2. Meanwhile, the H2S concentration
air isolation until pH = 8. After precipitation, the suspension was directly detected with a high-accuracy H2S detection tube
was exposed to air and allowed to crystallize at 30 °C for 3 h at the outlet. The curve of desulfurization was plotted
with stirring at this pH value. The obtained precipitate was according to the detected H2S concentration.
washed with deionized water to remove sulfate and impurity
ions and dried at 90 °C for 24 h. The resulting precursor was 3. RESULTS AND DISCUSSION
divided into four equal portions, which were sequentially 3.1. TG−DSC Analysis. A simultaneous thermal analysis
placed in a vacuum furnace and calcined at 150, 250, 350, and technique was applied to study the phase transformation of the
450 °C for 5 h. The calcined samples are herein denoted as FE- as-prepared γ-FeOOH precursor. As can be seen from Figure
150, FE-250, FE-350, and FE-450, respectively. 1, the precursor exhibits an endothermic peak at 108 °C, which
2.2. Characterization. The X-ray diffraction (XRD)
measurements of the adsorbents were performed on a
PANalytical diffractometer (X’PertPro) with a Co Kα radiation
source (λ = 0.17902 nm) at 40 mA and 40 kV. N2 adsorption−
desorption experiments were carried out on a Micromeritics
ASAP 2020 instrument. The samples were outgassed at 120 °C
for 2 h before N2 adsorption at −196 °C to measure the
Brunauer−Emmett−Teller (BET) surface area and Barett−
Joyner−Halenda pore size distribution. Preliminary thermog-
ravimetry−differential scanning calorimetry (TG−DSC) anal-
ysis of the γ-FeOOH precursor was conducted to follow the
phase transition process. The measurement was carried out
using a simultaneous thermal analyzer (Setsys Evolution). The
sample (20 mg) was first dried at 90 °C in an oven for 2 h and
then heated at a linear rate (10 °C min−1) from 50 to 500 °C
in a N2 flow (20 mL min−1). Transmission electron
Figure 1. TG−DSC curves of the γ-FeOOH precursor.
microscopy (TEM) and high-resolution TEM (HR-TEM)
images were taken using a high-resolution transmission
electron microscope (Tecnai G2 F20, operating at 200 kV). is ascribed to the release of physically adsorbed water. A large
In situ Fourier transform infrared spectroscopy (FT-IR) of endothermic peak with a sharp weight loss is detected at 225
pyridine adsorption were collected at 4 cm−1 resolution and °C, which is associated with dehydration (chemically adsorbed
128 scans over a Nicolet 6700 spectrometer (Thermo Fisher water) and dehydroxylation.30 In addition, there is a weak
Scientific) equipped with an MCT detector. The sample (25 exothermic peak at 470 °C with almost no weight loss
mg) was heated in a He flow at 150, 250, 350, or 450 °C for 5 corresponding to phase transformation of γ-Fe2O3 to α-
h and then cooled to 25 °C. Subsequently, pyridine was Fe2O3.3132
introduced into the system until adsorption saturation. After A previous study32 suggested that for the γ-Fe2O3 derived
purged with He, DRIFTS spectra were acquired at elevated from phase transition of γ-FeOOH there is existence of surface
temperatures. hydroxyl groups. It is noted that there are exothermic and
2.3. Desulfurization Performance Test. The evaluation endothermic processes, though indistinguishable, accompanied
of H2S breakthrough capacity for the samples was performed in by a slight weight loss in the temperature range of 350−450
a fixed-bed reactor at 25 °C. The sorbent sample (0.5 g) was °C. These phenomena may be attributed to factors such as the
secured in a quartz tube of 5 mm internal diameter, and a 4% removal of residual hydroxyl groups, rearrangement of ions
H2S/N2 gas stream (20 mL min−1) was fed into the reactor at a within the structure, crystallization of amorphous γ-Fe2O3 as
weight hourly space velocity (WHSV) of 2400 mL·g−1·h−1. well as phase transition of γ-Fe2O3 to α-Fe2O3.33 Therefore, we
The actual value of gas volume was measured by a wet-gas calcined the divided γ-FeOOH portions separately at different
flowmeter. Simultaneously, the H2S concentration at the outlet temperatures to obtained γ-FeOOH, γ-Fe2O3, and α-Fe2O3
was regularly recorded using a high-precision H2S detection according to the TG−DSC results.
tube. The test was stopped when the H2S concentration 3.2. Desulfurization Performance. Figures 2 and 3
reached the breakthrough point (330 ppm at the outlet). The illustrate the desulfurization precision and H2S breakthrough
amount of sulfur captured per gram of adsorbent was defined capacity of the FE-150, FE-250, FE-350, and FE-450
B DOI: 10.1021/acs.iecr.9b03430
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 2. (A) Desulphurization precision and (B) standard deviation (error bar) indicated for the FE-150, FE-250, and FE-350 samples. All data
are average of at least four measures (feed gas composition: 0.5% H2S and 99.5% N2).

was performed 4 times and the standard deviation presented in


Figure 2B shows that the tests are well reproducible. Overall,
the samples show a descending order of FE-250 > FE-150 >
FE-350 > FE-450 both for desulfurization precision (Figure 2)
and H 2 S breakthrough capacity (Figure 3). Table 1
summarizes the performance of various adsorbents for the
removal hydrogen sulfide at room temperature. It can be found
that the adsorption capacity of γ-FeOOH and γ-Fe2O3 is
higher than most of the listed adsorbents. The H2S
breakthrough capacity of FE-250 is 53%, while that of FE-
450 is close to 0%. It is clear that the calcination temperature
adopted for the preparation of sorbents has a significant effect
on desulfurization performance. The reason may be related to
certain key factors of sorbent generated at different calcination
temperatures, such as crystal phase, specific surface area, lattice
Figure 3. H2S breakthrough capacity of the FE-150, FE-250, FE-350, defect, and surface hydroxyl group, which will be discussed in
and FE-450 samples (feed gas: 4% H2S/N2). detail in the later section.
3.3. Structural Studies. The structures of FE-150, FE-250,
adsorbents, respectively. As shown in Figure 2, the samples FE-350, and FE-450 samples were characterized by XRD, and
calcined at different temperatures are different in desulfuriza- the results are presented in Figure 4A. The diffraction patterns
tion precision. The desulfurization performance of FE-450 is of the samples are obviously different. The sample calcined at
the worst, showing a H2S concentration of 3580 ppm in the 150 °C shows peaks that can be ascribed to γ-FeOOH (PDF
outlet at the start of recording. The FE-250 sample displays the no. 00-044-1415). After calcination at 250 and 350 °C, the
best desulfurization precision, showing an outlet H2S diffraction peaks of FE-250 and FE-350 match those of γ-
concentration of only 130 ppm after a reaction of 400 min. Fe2O3 (PDF no. 01-089-5892), while weak signals of γ-
In addition, the test of desulfurization precision for all samples FeOOH are still recognizable over the FE-250 sample. It is

Table 1. Summary of Metal Oxide and Hydroxide for H2S Removal at Room Temperature

support material
metal material formula adsorption capacity (%) SBET (m2/g) pore volume (cm3/g) average pore size (nm) refs
iron none γ-FeOOH (FE-150) 46.9 79 0.19 9.80 this work
γ-Fe2O3 (FE-250) 53.0 137 0.23 6.80
γ-Fe2O3 (FE-350) 39.6 68 0.18 11.0
α-Fe2O3 (FE-450) 0 17 0.10 22.6
iron none γ-Fe2O3/SiO2 40.61 113 8.21 8.1 26
γ-Fe2O3 10.58 82 9.75 11.6
α-Fe2O3 0 36 2.44 54.2
iron none Fe2O3 10 110 NA NA 34
iron none α-FeOOH 6.5 188 0.579 12.3 35
activated carbon α-FeOOH 17.1 500 0.449 3.6
iron none Fe3O4 1.54 73 0.313 17 36
graphene oxide Fe3O4 1.37 176 0.253 5.7
zinc none ZnO 0.89 NA NA NA 17
carbon nanotube ZnO 2.99 NA NA NA
cobalt none CoOOH 6.91 203 0.679 NA 37
graphene oxide CoOOH 6.62 191 0.417 NA

C DOI: 10.1021/acs.iecr.9b03430
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 4. (A) XRD pattern of the samples calcined at different temperatures and (B) crystal structure of γ-FeOOH.

noted that compared with FE-350, FE-250 is poorer in larger pores are formed due to the merging of smaller pores
crystallinity and smaller in grain size as indicated by the caused by structural rearrangement and phase transformation
relatively weak signals and large full width at half-maximum to α-Fe2O3. Therefore, the BET surface area and pore volume
values.38 As for the sample calcined at 450 °C, the crystal of the samples calcined at 350 and 450 °C decrease, while the
phase is α-Fe2O3 (PDF no. 01-086-2368) and the strong and pore diameter increases.41,42 In general, being larger in the
sharp peaks suggest high crystallinity and large particle size.39 BET surface area and pore volume is beneficial for
γ-FeOOH (lepidocrocite) is an isomorphic boehmite with a desulfurization. However, the FE-250 and FE-150 samples
layered structure, which consists of arrays of cubic close- are somewhat similar in desulphurization precision and H2S
packed (ccp) anions (O2−/OH−) stacked along the [150] breakthrough capacity despite the specific surface area of the
direction with Fe3+ occupying the octahedral interstices. This former being nearly twice that of the latter. Therefore, in
structure consists of double chains of Fe(O,OH)6 octahedral addition to the specific surface area, there may be other factors
running parallel to the c-axis. The double chains share edges affecting the desulfurization performance of the samples.
with adjacent double chains, forming corrugated sheets of 3.5. TEM Images. The TEM images of the FE-150, FE-
octahedral, which are held together by hydrogen bonds (Figure 250, and FE-450 samples are shown in Figure 5. As can be seen
4B). from Figure 5A, FE-150 is composed of platelets that are 200−
According to the results of TG−DSC and XRD analyses, the 400 nm in length and 40−60 nm in width. The HR-TEM
processes of hydroxylation and phase transition of γ-FeOOH image (Figure 5B) shows that FE-250 is dominated by (210)
could be illustrated as follows: At around 190 °C, there is the crystal facets of γ-FeOOH, showing a lattice spacing of 0.329
desorption of weakly adsorbed water as well as partial removal nm. The FE-250 sample retains the nanosheet structure of FE-
of hydroxyl groups. At 250 °C, the sample would have lost 150 (Figure 5C,D). There are slit pores on the lamellar
most of the hydroxyl groups and there is phase transformation structure, which may be due to the cleavage of nanosheet
from γ-FeOOH to γ-Fe2O3. At 350 °C, there is structural (about 3.2 nm in thickness) in γ-FeOOH caused by the
rearrangement and grain growth as well as formation of defect removal of hydroxyl and water during thermal treatment. The
sites. In the process of phase transformation, the collapse of γ- drastic dehydroxylation and dehydration cause the collapse of
FeOOH structure is relatively rapid at the beginning and there the γ-FeOOH structure and result in the formation of γ-Fe2O3.
is ample presence of γ-Fe2O3 crystal nuclei. However, in the However, we cannot obtain clear HR-TEM images of FE-250
later stage, the rearrangement process gradually slows down due to the strong magnetism of γ-Fe2O3. As for the FE-450
and the growth of γ-Fe2O3 grains becomes lagging,40 as sample, the nanoparticles are dense and irregular in size
reflected in the weak and broad XRD peaks of the FE-250 and (Figure 5E). The HR-TEM image of FE-450 in Figure 5F
FE-350 samples. At 450 °C, there is the complete conversion reveals that the nanoparticles are enclosed by (104) facets with
of γ-Fe2O3 to α-Fe2O3. an interplanar spacing of 0.270 nm, which confirms that the
3.4. Texture Analysis. Nitrogen adsorption−desorption crystalline phase of FE-450 is α-Fe2O3.43
measurements were conducted to investigate the texture of the The above TEM images intuitively show the change of
samples (Table 1). After calcination at 150 °C, the γ-FeOOH morphology and crystal phase with the calcination temper-
has low BET surface area, pore volume, and average pore size. ature, which is consistent with the XRD and BET results.
When the calcination temperature was 250 °C, the resulted Based on this information, it is deduced that there is variation
FE-250 sample, which is mainly γ-Fe2O3, has significantly in the number of hydroxyl groups across the sorbents.
larger BET surface area and pore volume but smaller average Meanwhile, with the rearrangement of iron and oxygen ions,
pore size. A further rise of calcination temperature to 350 and there is generation of defects in the iron−oxygen matrices.44
450 °C would result in a gradual decrease of BET surface area With the idea that hydroxyl groups and defects could get
and pore volume but in an increase of average pore size. This involved in the low-temperature removal of hydrogen sulfide
may be because the removal of hydroxyl groups at 150 °C by ferric oxide, we proceeded to characterize these species by
would result in pores of a small size.30 Previous studies by FT-IR and in situ DRIFTS of pyridine adsorption.
Sudakar and Giovanoli showed that the discharge of adsorbed 3.6. In Situ DRIFTS of Pyridine Adsorption. Due to
water during the initial stage of γ-FeOOH dehydroxylation dehydroxylation during heat treatment, there is ready
could create a large number of micropores.30,40 Therefore, the formation of a large amount of oxygen vacancies on the
BET surface area of FE-250 (137 m2 g−1) is higher than that of FeOOH surface. Meanwhile, corresponding to the defect sites,
FE-150 (79 m2 g−1). When calcined at 450 °C, a number of there is the presence of electron-receptive Fe3+ ions of
D DOI: 10.1021/acs.iecr.9b03430
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 6. In situ DRIFTS spectra of pyridine on γ-FeOOH recorded


at different temperatures. (RT-Py represents pyridine adsorption at
room temperature.)

is a basic hydroxyl group and can directly adsorb the H2S


molecules.
To avoid the interference of physically adsorbed and
hydrogen-bonded pyridine, we treated the FE-150, FE-250,
FE-350, and FE-450 samples at 150 °C under vacuum before
pyridine adsorption. As shown in Figure 7, the peak intensity

Figure 5. TEM images (left) and corresponding HR-TEM images


(right) of samples: (A, B) FE-150, (C, D) FE-250, and (E, F) FE-450.

unsaturated coordination. In the reaction of H2S removal, both


H2S and SH− groups readily bond to the electrophilic oxygen
vacancies on the surface of ferric oxide. Therefore, the
concentration of surface oxygen vacancies on ferric oxide has
a significant impact on its desulfurization performance.26 In
view of the electrophilic properties of oxygen vacancies on the Figure 7. Pyridine absorption curves of FE-150, FE-250, FE-350, and
surface of ferric oxide, we characterized and measured the FE-450 samples with pyridine adsorbed at 150 °C.
Lewis acid sites of the catalyst.45
In view that pyridine is the most used basic probe to study of pyridine increases with the rise of calcination temperature,
surface acidity, we used in situ DRIFTS of pyridine adsorption which indicates that the number of Lewis acid sites increases
to characterize the surface acid sites of the γ-FeOOH sample with the rise of calcination temperature. In this temperature
(Figure 6). The sample exposed to pyridine at room range, most of the OH groups of the samples have been
temperature shows absorption bands at 1400−1700 cm−1. removed and there is a change of Fe3+ chemical environment
The bands at ca. 1448 and 1489 cm−1 can be ascribed to the and the generation of a large number of Lewis acid sites. As for
ν19b and ν19a modes of pyridine adsorbed on Lewis acid sites, FE-450, the intensity of the bands suffers a decrease because of
respectively.46 The bands at 1576 cm−1 are assigned to the the formation of α-Fe2O3 (Figure 4A) as well as the reduction
physically adsorbed pyridine.47 The bands at 1593 cm−1 are of the BET surface area (Table 1). Consequently, there is a
related to hydrogen-bonded pyridine, having the nitrogen atom significant decrease in the number of Lewis acid sites. Taking
weakly interacting with surface hydroxyl groups through that oxygen vacancies act as Lewis acid sites on the iron-based
hydrogen bonding.48 When the temperature increases to 150 materials, the number of oxygen vacancies across the samples
°C, the absorption bands assigned to hydrogen-bonded and decreases in the order of FE-350 > FE-450 > FE-250 > FE-150.
physically adsorbed pyridine disappear completely, indicating As can be seen, there is a certain correspondence between the
that both of them are weakly adsorbed and relatively easy to oxygen vacancies and desulfurization performance such as in
remove. As for the absorption bands at ca. 1448, 1489 and the cases of FE-250 and FE-150 (Figure 3). However, although
1605 cm−1, they can still be detected when the temperature is FE-350 has the highest number of oxygen vacancies, it is not
raised to 250 °C, suggesting the presence of Lewis acid sites. In the best in desulfurization performance, while having the least
addition, there is no detection of bands that could be related to number of oxygen vacancies, FE-150 shows desulfurization
Brønsted acid sites, which indicates that the −OH in FeOOH performance better than that of FE-350. The results indicate
E DOI: 10.1021/acs.iecr.9b03430
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

that the number of oxygen vacancies is not a dominating factor number of Lewis acid sites, the effect of surface hydroxyl
that governs desulfurization performance. groups and defects should not be ignored.
3.7. FT-IR Analysis. We hence turned our attention to the 3.8. Normalization Analysis of the Three Factors.
role of surface hydroxyl groups in the iron oxides. The FT-IR Previous studies52,53 have demonstrated that the interaction
spectra of FE-150, FE-250, FE-350, and FE-450 samples were between adsorbed H2S molecules and metal oxides involves
recorded to investigate the relationship between surface lattice diffusion, which is the migration or diffusion of ions in
hydroxyl groups and desulfurization performance. As shown the oxide lattice. It is well conceived that the presence of defect
in Figure 8, the IR spectrum of sample FE-150 shows typical γ- sites (e.g., oxygen vacancies) not only promotes lattice
diffusion but also reduces the activation energy for ion
exchange, enhancing the interaction of ions with surface iron
atoms as a consequence.54 Nonetheless, it is envisaged that
lattice diffusion at low temperatures is difficult and
desulfurization at room temperature is not controlled by the
rate of lattice diffusion. Although the mechanism of ion
diffusion in the lattice at low temperatures is not well
understood, the conducive effect of surface hydroxyl groups
on the dissociation of H2S molecules into S2− and SH− is
known.55
Based on the above analyses, it is deduced that there is a
correlation between the three factors (viz., the number of
hydroxyl groups, oxygen vacancies, and specific surface area)
and desulfurization performance. We conducted a normal-
ization analysis to explore the degree of influence of the three
factors on H2S breakthrough capacity and desulfurization
precision (Figure 9). FE-250 and FE-150, which have a high
Figure 8. FT-IR spectra of FE-150, FE-250, FE-350, and FE-450
samples.

FeOOH bands at 1153 and 1022 cm−1, which can be assigned


to in-plane Fe−OH bending vibrations.49,50 In addition, there
is no detection of bands related to H2O bending vibration near
1636 cm−1, indicating the complete removal of physically
adsorbed H2O. The sample has a strong and wide band at
2990−3700 cm−1, which can be assigned to the OH stretching
vibration of Fe−OH groups and chemisorbed H2O.51 During
dehydroxylation, the bands at 1550, 1434, and 1336 cm−1 can
be attributed to a small amount of residual carboxylate ligand,
carbonate, and/or NH4+ species. The band of OH stretching
vibration is still detectable over FE-250, indicating that there is
still ample presence of hydroxyl groups. The OH bands Figure 9. Normalization analysis of the three factors (i.e., the number
become extremely weak over FE-350 and almost undetectable of hydroxyl groups, density of oxygen vacancies, and specific surface
over FE-450, indicating the gradual and final removal of area) on (a) desulfurization precision and (b) H2S breakthrough
hydroxyl groups. Overall, the number of hydroxyl groups capacity.
decrease in the order of FE-150 > FE-250 > FE-350 > FE-450.
It is worth pointing out that the bands at 1153 and 1022 cm−1 specific surface area and a large number of hydroxyl groups,
of FE-150 become broadened and weakened with the rise of respectively, have high H2S breakthrough capacity and
calcination temperature, indicating the occurrence of phase desulfurization precision. Although FE-350 has the highest
transition and atom rearrangement. number of oxygen vacancies, its performance is below that of
Generally, with adsorbed H2O on the FeOOH surface, H2S FE-250 and FE-150 due to the decrease of the specific surface
molecules diffuse into the surface basic water film formed by area and surface hydroxyl groups. Despite medium in the
hydroxyl groups and dissociate into S2− and SH− species that number of oxygen vacancies, FE-450 has the lowest
displace the OH− and O2− on the FeOOH surface, leading to desulfurization performance because it has the lowest specific
the formation of FeSx species. To the H2S molecules, the surface area and number of hydroxyl groups. It is hence
surface hydroxyl groups play a “capture” role, blocking the H2S deduced that the specific surface area and the number of
molecules from penetrating into the bulk of adsorbent and hydroxyl groups have stronger influence than the number of
increasing the contact time between H2S molecules and iron oxygen vacancies on desulfurization. To optimize the perform-
oxides. Therefore, the existence of surface hydroxyl groups ance of iron oxides for H2S removal at room temperature, the
improves the sulfur capacity and desulfurization precision. FeOOH precursor has to be treated in a way to achieve high
Perhaps this might explain why in test of desulfurization specific surface area and abundant presence of surface hydroxyl
performance over FE-350 at H2S concentration of 0.5% there groups. At low temperatures, the effect of Lewis acid sites
is high desulfurization precision but low H2S breakthrough (oxygen vacancies) is less significant due to the poor diffusion
capacity. Thus, in addition to the specific surface area and the of ions in the lattice. Such an understanding can serve as a
F DOI: 10.1021/acs.iecr.9b03430
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

guideline for the design of solid adsorbents for desulfurization (9) Zheng, X. X.; Shen, L. J.; Chen, X. P.; Zheng, X. H.; Au, C. T.;
at low temperatures. Jiang, L. L. Amino-modified Fe-Terephthalate metal-organic frame-
work as an efficient catalyst for the selective oxidation of H2S. Inorg.
4. CONCLUSIONS Chem. 2018, 57, 10081−10089.
(10) Fauteux-Lefebvre, C.; Abatzoglou, N.; Blais, S.; Braidy, N.; Hu,
Samples of γ-FeOOH, γ-Fe2O3, and α-Fe2O3 were fabricated Y. Iron oxide-functionalized carbon nanofilaments for hydrogen
by calcining γ-FeOOH at selected temperatures and evaluated sulfide adsorption: The multiple roles of carbon. Carbon 2015, 95,
for H2S removal. Both specific surface area and surface 794−801.
hydroxyl groups have great influence on desulfurization (11) Habibi, R.; Rashidi, A. M.; Daryan, J. T.; Zadeh, A. M. A. Study
performance. Among the prepared sorbents, the FE-250 of the rod-like and spherical nano-ZnO morphology on H2S removal
calcined at 250 °C exhibited the highest H2S breakthrough from natural gas. Appl. Surf. Sci. 2010, 257, 434−439.
capacity (53%) and desulfurization precision at 25 °C. The (12) Liu, D.; Zhou, W.; Wu, J. CeO2−MnOx/ZSM-5 sorbents for
results of the present work elucidate the factors that govern the H2S removal at high temperature. Chem. Eng. J. 2016, 284, 862−871.
desulfurization performance of iron oxides at room temper- (13) Li, L.; Sun, T.; Shu, C.; Zhang, H. Low temperature H2S
ature, and such an understanding is beneficial for the design of removal with 3-D structural mesoporous molecular sieves supported
desulfurizers. ZnO from gas stream. J. Hazard. Mater. 2016, 311, 142−150.


(14) Liu, B.; Wan, Z.; Wang, F.; Zhan, Y.; Tian, M.; Cheung, A. 18O2
AUTHOR INFORMATION label mechanism of sulfur generation and characterization in
properties over mesoporous Sm-based sorbents for hot coal gas
Corresponding Authors desulfurization. J. Hazard. Mater. 2014, 267, 229−237.
*E-mail: ying.zheng@uwo.ca (Y.Z.). (15) Liu, B.; Wei, X.; Zhan, Y.; Chang, R.; Subhan, F.; Au, C.
*E-mail: jll@fzu.edu.cn (L.J.). Preparation and desulfurization performance of LaMeOx/SBA-15 for
ORCID hot coal gas. Appl. Catal., B 2011, 102, 27−36.
Xiaohai Zheng: 0000-0002-3652-2236 (16) Garces, H. F.; Galindo, H. M.; Garces, L. J.; Hunt, J.; Morey,
Lilong Jiang: 0000-0002-0081-0367 A.; Suib, S. L. Low temperature H2S dry-desulfurization with zinc
oxide. Microporous Mesoporous Mater. 2010, 127, 190−197.
Notes (17) Song, H. S.; Park, M. G.; Kwon, S. J.; Yi, K. B.; Croiset, E.;
The authors declare no competing financial interest.


Chen, Z.; Nam, S. C. Hydrogen sulfide adsorption on nano-sized zinc
oxide/reduced graphite oxide composite at ambient condition. Appl.
ACKNOWLEDGMENTS Surf. Sci. 2013, 276, 646−652.
This work was financially supported by the National Natural (18) Cheah, S.; Carpenter, D. L.; Magrini-Bair, K. A. Review of mid-
Science Foundation of China (21878052), Natural Science to high-temperature sulfur sorbents for desulfurization of biomass-and
Foundation of Fujian Province (2018J01693), and Young and coal-derived syngas. Energy Fuels 2009, 23, 5291−5307.
Middle-aged Teacher Education and Science Research (19) Lee, S.; Kim, D. Enhanced adsorptive removal of hydrogen
Foundation of Fujian Province, China (JAT160047). sulfide from gas stream with zinc-iron hydroxide at room temperature.


Chem. Eng. J. 2019, 363, 43−48.
(20) Bakker, W. J.; Kapteijn, F.; Moulijn, J. A. A high capacity
REFERENCES
manganese-based sorbent for regenerative high temperature desulfur-
(1) Shah, M. S.; Tsapatsis, M.; Siepmann, J. I. Hydrogen sulfide ization with direct sulfur production: Conceptual process application
capture: From sbsorption in polar liquids to oxide, zeolite, and metal- to coal gas cleaning. Chem. Eng. J. 2003, 96, 223−235.
organic framework adsorbents and membranes. Chem. Rev. 2017, 117, (21) Zhang, R.; Huang, J.; Zhao, J.; Sun, Z.; Wang, Y. Sol−gel auto-
9755−9803. combustion synthesis of zinc ferrite for moderate temperature
(2) Lei, G.; Dai, Z.; Fan, Z.; Zheng, X.; Cao, Y.; Shen, L.; Xiao, Y.; desulfurization. Energy Fuels 2007, 21, 2682−2687.
Au, C.; Jiang, L. Porous nanosheets of carbon-conjugated graphitic (22) Slimane, R. B.; Abbasian, J. Utilization of metal oxide-
carbon nitride for the oxidation of H2S to elemental sulfur. Carbon containing waste materials for hot coal gas desulfurization. Fuel
2019, 155, 204−214.
Process. Technol. 2001, 70, 97−113.
(3) Kamali, F.; Eskandari, M. M.; Rashidi, A.; Baghalha, M.;
(23) Baird, T.; Campbell, K. C.; Holliman, P. J.; Hoyle, R.; Noble,
Hassanisadi, M.; Hamzehlouyan, T. Nanorod carbon nitride as a
G.; Stirling, D.; Williams, B. P. Mixed cobalt-iron oxide absorbents for
carbo catalyst for selective oxidation of hydrogen sulfide to sulfur. J.
low-temperature gas desulfurisation. J. Mater. Chem. 2003, 13, 2341−
Hazard. Mater. 2019, 364, 218−226.
(4) Lei, G.; Cao, Y.; Zhao, W.; Dai, Z.; Shen, L.; Xiao, Y.; Jiang, L. 2347.
Exfoliation of graphitic carbon nitride for enhanced oxidative (24) Dhage, P.; Samokhvalov, A.; Repala, D.; Duin, E. C.;
desulfurization: A facile and general strategy. ACS Sustainable Chem. Tatarchuk, B. J. Regenerable Fe-Mn-ZnO/SiO2 sorbents for room
Eng. 2019, 7, 4941−4950. temperature removal of H2S from fuel reformates: performance, active
(5) Wu, M.; Shi, L.; Lim, T.-T.; Veksha, A.; Yu, F.; Fan, H.; Mi, J. sites, Operando studies. Phys. Chem. Chem. Phys. 2011, 13, 2179−
Ordered mesoporous Zn-based supported sorbent synthesized by a 2187.
new method for high-efficiency desulfurization of hot coal gas. Chem. (25) Gong, Z.; Tian, Y.; Li, W. Research on the crystalline phase and
Eng. J. 2018, 353, 273−287. desulfurization activity of FeOOH prepared by different raw material.
(6) Zheng, X.; Li, Y.; Zhang, L.; Shen, L.; Xiao, Y.; Zhang, Y.; Au, C.; J. China Coal Soc. 2006, 31, 790−793.
Jiang, L. Insight into the effect of morphology on catalytic (26) Huang, G.; He, E.; Wang, Z.; Fan, H.; Shangguan, J.; Croiset,
performance of porous CeO2 nanocrystals for H2S selective oxidation. E.; Chen, Z. Synthesis and characterization of γ-Fe2O3 for H2S
Appl. Catal., B 2019, 252, 98−110. removal at low temperature. Ind. Eng. Chem. Res. 2015, 54, 8469−
(7) Zhang, F.; Zhang, X.; Jiang, G.; Li, N.; Hao, Z.; Qu, S. H2S 8478.
selective catalytic oxidation over Ce substituted La1−xCexFeO3 (27) Cao, Y.; Shen, L.; Hu, X.; Du, Z.; Jiang, L. Low temperature
perovskite oxides catalyst. Chem. Eng. J. 2018, 348, 831−839. desulfurization on Co-doped α-FeOOH: Tailoring the phase
(8) Lei, C.; Zhou, W.; Shen, L.; Zheng, X.; Feng, Q.; Liu, Y.; Lei, Y.; composition and creating the defects. Chem. Eng. J. 2016, 306,
Liang, S.; Zhang, D.; Jiang, L.; Zhou, K. Enhanced selective H2S 124−130.
oxidation performance on Mo2C modified g-C3N4. ACS Sustainable (28) Cao, Y.; Hu, X.; Lin, X.; Lin, Y.; Huang, R.; Jiang, L.; Wei, K.
Chem. Eng. 2019, 7, 16257−16263. Low-temperature desulfurization on iron oxide hydroxides: influence

G DOI: 10.1021/acs.iecr.9b03430
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

of precipitation pH on structure and performance. Ind. Eng. Chem. amorphous silica prepared from iron oxide of bacterial origin. ACS
Res. 2015, 54, 2419−2424. Appl. Mater. Interfaces 2013, 5, 518−523.
(29) Shen, L.; Cao, Y.; Du, Z.; Zhao, W.; Lin, K.; Jiang, L. Illuminate (49) Ristić, M.; Musić, S.; Godec, M. Properties of γ-FeOOH, α-
the active sites of γ-FeOOH for low-temperature desulfurization. Appl. FeOOH and α-Fe2O3 particles precipitated by hydrolysis of Fe3+ ions
Surf. Sci. 2017, 425, 212−219. in perchlorate containing aqueous solutions. J. Alloys Compd. 2006,
(30) Sudakar, C.; Subbanna, G.; Kutty, T. Effect of anions on the 417, 292−299.
phase stability of γ-FeOOH nanoparticles and the magnetic properties (50) Nasrazadani, S. The application of infrared spectroscopy to a
of gamma-ferric oxide derived from lepidocrocite. J. Phys. Chem. Solids study of phosphoric and tannic acids interactions with magnetite
2003, 64, 2337−2349. (Fe3O4), goethite (α-FeOOH) and lepidocrocite (γ-FeOOH). Corros.
(31) Dinesen, A.; Pedersen, C.; Koch, C. B. The thermal conversion Sci. 1997, 39, 1845−1859.
of lepidocrocite (γ-FeOOH) revisited. J. Therm. Anal. Calorim. 2001, (51) Ning, P.; Li, K.; Yi, H.; Tang, X.; Peng, J.; He, D.; Wang, H.;
64, 1303−1310. Zhao, S. Simultaneous catalytic hydrolysis of carbonyl sulfide and
(32) Cornell, R. M.; Schwertmann, U. The Iron Oxides: Structure, carbon disulfide over modified microwave coal-based active carbon
Properties, Reactions, Occurrences and Uses; John Wiley & Sons, 2003. catalysts at low temperature. J. Phys. Chem. C 2012, 116, 17055−
(33) Jing, Z.; Wu, S. Synthesis, characterization and magnetic 17062.
properties of γ-Fe2O3 nanoparticles via a non-aqueous medium. J. (52) Polychronopoulou, K.; Fierro, J.; Efstathiou, A. Novel Zn-Ti-
Solid State Chem. 2004, 177, 1213−1218. based mixed metal oxides for low-temperature adsorption of H2S from
(34) Hussain, M.; Abbas, N.; Fino, D.; Russo, N. Novel mesoporous industrial gas streams. Appl. Catal., B 2005, 57, 125−137.
silica supported ZnO adsorbents for the desulphurization of biogas at (53) Carnes, C. L.; Klabunde, K. J. Unique chemical reactivities of
low temperatures. Chem. Eng. J. 2012, 188, 222−232. nanocrystalline metal oxides toward hydrogen sulfide. Chem. Mater.
(35) Lee, S.; Lee, T.; Kim, D. Adsorption of hydrogen sulfide from 2002, 14, 1806−1811.
gas streams using the amorphous composite of α-FeOOH and (54) Kim, G.; Wang, S.; Jacobson, A. J.; Reimus, L.; Brodersen, P.;
activated carbon powder. Ind. Eng. Chem. Res. 2017, 56, 3116−3122. Mims, C. A. Rapid oxygen ion diffusion and surface exchange kinetics
(36) Arcibar-Orozco, J. A.; Wallace, R.; Mitchell, J. K.; Bandosz, T. J. in PrBaCo2O5+x with a perovskite related structure and ordered A
Role of surface chemistry and morphology in the reactive adsorption cations. J. Mater. Chem. 2007, 17, 2500−2505.
of H2S on iron (hydr)oxide/graphite oxide composites. Langmuir (55) Seredych, M.; Bandosz, T. J. Reactive adsorption of hydrogen
2015, 31, 2730−2742. sulfide on graphite oxide/Zr(OH)4 composites. Chem. Eng. J. 2011,
(37) Florent, M.; Bandosz, T. J. Effects of surface heterogeneity of 166, 1032−1038.
cobalt oxyhydroxide/graphite oxide composites on reactive adsorp-
tion of hydrogen sulfide. Microporous Mesoporous Mater. 2015, 204,
8−14.
(38) Schimanke, G.; Martin, M. In situ XRD study of the phase
transition of nanocrystalline maghemite (γ-Fe2O3) to hematite (α-
Fe2O3). Solid State Ionics 2000, 136−137, 1235−1240.
(39) Xiao, Y.; Zheng, X.; Chen, X.; Jiang, L.; Zheng, Y. Synthesis of
Mg-doped ordered mesoporous Pd-Al2O3 with different basicity for
CO, NO, and HC elimination. Ind. Eng. Chem. Res. 2017, 56, 1687−
1695.
(40) Giovanoli, R.; Brütsch, R. Kinetics and mechanism of the
dehydration of γ-FeOOH. Thermochim. Acta 1975, 13, 15−36.
(41) Zheng, X.; Chen, X.; Chen, J.; Zheng, Y.; Jiang, L. Synthesis
and application of highly dispersed ordered mesoporous silicon-doped
Pd-alumina catalyst with high thermal stability. Chem. Eng. J. 2016,
297, 148−157.
(42) Ruan, H.; Gilkes, R. Kinetics of phosphate sorption and
desorption by synthetic aluminous goethite before and after thermal
transformation to hematite. Clay Miner. 1996, 31, 63−74.
(43) Zhao, W.; Zheng, X.; Liang, S.; Zheng, X.; Shen, L.; Liu, F.;
Cao, Y.; Wei, Z.; Jiang, L. Fe-doped γ-Al2O3 porous hollow
microspheres for enhanced oxidative desulfurization: Facile fabrica-
tion and reaction mechanism. Green Chem. 2018, 20, 4645−4654.
(44) Pinney, N.; Morgan, D. Ab initio study of structurally bound
water at cation vacancy sites in Fe-and Al-oxyhydroxide materials.
Geochim. Cosmochim. Acta 2013, 114, 94−111.
(45) Thavornprasert, K.-A.; Capron, M.; Jalowiecki-Duhamel, L.;
Gardoll, O.; Trentesaux, M.; Mamede, A.-S.; Fang, G.; Faye, J.;
Touati, N.; Vezin, H.; et al. Highly productive iron molybdate mixed
oxides and their relevant catalytic properties for direct synthesis of 1,
1-dimethoxymethane from methanol. Appl. Catal., B 2014, 145, 126−
135.
(46) Shen, L.; Zheng, X.; Lei, G.; Li, X.; Cao, Y.; Jiang, L.
Hierarchically porous γ-Al2O3 nanosheets: Facile template-free
preparation and reaction mechanism for H2S selective oxidation.
Chem. Eng. J. 2018, 346, 238−248.
(47) Akçay, M. The surface acidity and characterization of Fe-
montmorillonite probed by in situ FT-IR spectroscopy of adsorbed
pyridine. Appl. Catal., A 2005, 294, 156−160.
(48) Hashimoto, H.; Itadani, A.; Kudoh, T.; Kuroda, Y.; Seno, M.;
Kusano, Y.; Ikeda, Y.; Nakanishi, M.; Fujii, T.; Takada, J. Acidic

H DOI: 10.1021/acs.iecr.9b03430
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

You might also like