You are on page 1of 7

Chemical Engineering Journal 287 (2016) 47–53

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Adsorption of hydrogen sulfide (H2S) on zeolite (Z): Retention


mechanism
Léa Sigot, Gaëlle Ducom ⇑, Patrick Germain
Université de Lyon, INSA-Lyon, LGCIE-DEEP, EA4126, 20 Avenue Albert Einstein, F-69621 Villeurbanne Cedex, France

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Physicochemical characterization of Z
was performed before and after H2S
adsorption.
 Water is present in Z pores and pH
promotes H2S dissociation in HS.
 After H2S adsorption, the adsorbate
degraded between 300 and 630 °C in
inert atmosphere.
 After H2S adsorption, the adsorbate
could oxidize around 355 °C in air
atmosphere.
 An adsorption–dissociation–
oxidation mechanism led to the
formation of elemental sulfur.

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

Article history: Zeolites are efficient adsorbents for the removal of hydrogen sulfide (H2S) – a trace compound commonly
Received 30 August 2015 found in biogas. The assessment of zeolite regeneration potential is an important point to consider before
Received in revised form 2 November 2015 using them in adsorption processes for biogas treatment. In this study, it was first shown that after
Accepted 3 November 2015
adsorption of H2S, by heating a 13X zeolite at 350 °C, the desorbed quantities of sulfur compounds were
Available online 10 November 2015
very low. Afterwards, the objective was to understand H2S removal mechanism. Complementary physic-
ochemical characterizations of the studied zeolite before and after H2S adsorption were performed: speci-
Keywords:
fic surface area, porosity, pH, infrared spectroscopy, thermogravimetry and differential scanning
Zeolite
Hydrogen sulfide
calorimetry. The speciation of sulfur in the studied zeolite after H2S adsorption was investigated and a
Biogas retention mechanism was proposed. It was demonstrated that H2S was converted in elemental sulfur
Physicochemical characterization in the zeolite. This mechanism involves H2S adsorption at the zeolite surface, dissolution and dissociation
Adsorption–dissociation–oxidation of H2S in pore water, and oxidation to form elemental sulfur that can polymerize. The major formation of
mechanism elemental sulfur limits the thermal regeneration of the studied zeolite and other regeneration methods
should be considered.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction sulfide (H2S). Biogas desulfurization is commonly achieved by


physical, chemical or biological treatments. Nevertheless, adsorp-
Biogas is a promising source of energy. However, its use in tion is required when a thorough removal of H2S is necessary, for
energy conversion devices is hampered by the presence of trace instance for fueling a fuel cell [1–6]. In a previous study [7], three
compounds such as sulfur compounds and particularly hydrogen adsorbents were compared at lab-scale for the removal of H2S in a
synthetic gas: an activated carbon, a silica gel and a zeolite (Z). Z
proved to be the most efficient adsorbent for H2S removal, with
⇑ Corresponding author.
an adsorption capacity higher than 180 mgH2S/gZ.
E-mail address: gaelle.ducom@insa-lyon.fr (G. Ducom).

http://dx.doi.org/10.1016/j.cej.2015.11.010
1385-8947/Ó 2015 Elsevier B.V. All rights reserved.
48 L. Sigot et al. / Chemical Engineering Journal 287 (2016) 47–53

From an industrial point of view, the regenerability of the cho- 350 °C for 30 min after H2S adsorption (Z+H2S) with samples of
sen adsorbent is a relevant issue in order to reduce the operating 600–910 mg (flow rate: 27–30 mL min1). The desorbed gas was
costs. Moreover, it is a key concern for waste minimization as collected in a sampling bag. Sulfur compounds were immediately
adsorbents are generally disposed of after use. But incomplete quantified by micro GC coupled with mass spectrometry
regeneration was reported after H2S adsorption on activated car- (lGC-MS).
bons and an iron-based adsorbent [3,4,8,9]. So, in this paper, Z
regenerability after H2S adsorption is investigated. The objective 2.3. Elemental analysis
is to identify the speciation of sulfur in Z and the associated H2S
removal mechanism. A sample of ground Z (<80 lm) was analyzed by ICP-OES
The removal mechanisms of H2S have already been studied in (inductively coupled plasma – optical emission spectrometry) after
activated carbons (impregnated or unimpregnated). In the pres- alkaline fusion and acidic dissolution to determine its elemental
ence of oxygen, several authors revealed an oxidative mechanism content. Major elements were quantified: silicon (Si), aluminum
for H2S removal by activated carbons, leading to the formation of (Al), iron (Fe), manganese (Mn), magnesium (Mg), calcium (Ca),
elemental sulfur and water. As a function of pH and temperature, sodium (Na), potassium (K), titanium (Ti) and phosphorus (P).
H2S can also be oxidized to sulfur oxides or sulfuric acid [9–17].
But the retention mechanism of H2S in zeolites has not been widely 2.4. Total sulfur content
studied. Some authors suggested the formation of elemental sulfur
[10,18], but Micoli et al. proposed an acid-base reaction [5]. The total sulfur content was measured for Z and Z+H2S ground
In this study, thermal desorption of H2S was first studied. Then, samples (<80 lm) using a carbon-sulfur analyzer. The measure-
to understand H2S removal mechanism, complementary physico- ment principle is based on the total combustion of the sample in
chemical characterizations of Z solid matrix before and after H2S oxygen flow in an induction furnace, followed by infrared analysis
adsorption were performed in order to compare the different of the formed SO2.
results and converge to a conclusion. A similar methodology was
used previously to understand the interactions between 2.5. pH of the adsorbent leachate
octamethylcyclotetrasiloxane and a silica gel [19].
The pH of Z and Z+H2S leachates was determined according to
the EN 12457-2 standard for leaching tests of granular waste mate-
2. Material and methods
rial [22]. A liquid to solid ratio of 10 L/kg was used. A volume of
200 mL of demineralized water was added to a sample of 20 g of
2.1. Materials: Z and H2S
adsorbent and the mixture was stirred for 24 h to reach equilib-
rium. The suspension was then filtered (<0.45 lm) under vacuum
Zeolites are crystalline and porous aluminosilicates made up of
a regular 3D framework resulting from the arrangement of SiO4 and the pH of the filtrate was measured.
and AlO4 tetrahedrons [20,21]. A 13X zeolite (1.6–2.5 mm beads
SiliporiteÒ G5 from CECA) used previously in adsorption lab-tests 2.6. Total soluble sulfur and sulfate ions in leachate
showed a very good H2S removal efficiency. These dynamic adsorp-
tion tests were described elsewhere [7]. They were performed with The total soluble sulfur content of the leachates obtained
according to the procedure described in Section 2.5 was measured
a synthetic polluted gas containing 80 ppmv H2S. The diluent gas
was dry N2 containing 1% O2 due to incomplete air separation by ICP-OES following the NF EN ISO 11885 [23]. The total sulfur in
solution includes all soluble sulfur ions such as sulfates, thiosul-
in the N2 generator. N2 was used instead of methane (CH4) and car-
bon dioxide (CO2) which are the main components of biogas. fates and some sulfides, for the most common [24]. The spectral
band at 181.978 nm was used for the quantification. Measure-
In this study, the zeolite was characterized before and after H2S
adsorption. In the following, the notation ‘‘Z” refers to zeolite ments were performed for Z and Z+H2S leachates.
These leachates were also analyzed by ion chromatography.
before adsorption (fresh Z) and the notation ‘‘Z+H2S” refers to zeo-
lite after H2S adsorption (spent Z). Sulfates (SO2
4 ) were quantified according to the NF EN ISO
10304-2 [25].
For practical reasons, the characterizations were performed on
several polluted samples with different H2S uptakes. The average
2.7. Specific surface area and porosity
quantity of H2S adsorbed by each sample (Z+H2S_1 to Z+H2S_3) is
given in Table 1 in mgH2S/gZ and in mgS/gZ+H2S (if expressed as total
adsorbed sulfur relatively to the mass of adsorbent after adsorption). Surface area, micropore volume and pore size measurements
These quantities were determined by gas chromatography (GC) were performed by nitrogen adsorption (at 77 K) in an ASAP
analyses of the gas entering and leaving the adsorption column. 2010 (Micromeritics) instrument. Samples were first outgassed
under vacuum at 300 °C. Specific surface area and porosity were
determined before and after H2S adsorption (Z and Z+H2S). The
2.2. Thermodesorption
micro- and mesopore size distribution was determined by the den-
sity functional theory (DFT).
To assess the thermal regeneration potential of the zeolite, des-
orption tests in inert (He) and oxidant (air) flow were conducted at
2.8. Fourier transform infrared spectroscopy (FTIR)

Table 1 Z and Z+H2S ground samples (<1 mm) were analyzed by FTIR in
Average H2S quantity adsorbed in the different Z+H2S samples. attenuated total reflectance (ATR) mode. FTIR analyses enable the
characterization of the chemical bonds present in the material
Sample H2S adsorbed quantity
thanks to the measurement of their typical vibration frequency.
(mgH2S/gZ) mgS/gZ+H2S
Absorbance spectra were acquired between 500 and 4000 cm1
Z+H2S_1 133 110 with a resolution of 4 cm1. OMNIC software (Nicolet instrument)
Z+H2S_2 195 153
was used for data acquisition and post-treatment. A qualitative and
Z+H2S_3 252 189
comparative approach was favored.
L. Sigot et al. / Chemical Engineering Journal 287 (2016) 47–53 49

2.9. Thermogravimetry and differential scanning calorimetry (TG-DSC) Sulfite


Tetrathionate SO32-
S4O62-
Using a Labsys 1600 Setaram thermal analyzer, thermogravime- Elemental Sulfur
try (TG) and differential scanning calorimetry (DSC) were per- Sulfide sulfur Thiosulfate dioxide Sulfate
S2- S0 S2O32- SO2 SO42- Sulfur oxidaon
formed in replicates on Z and Z+H2S samples in either nitrogen
-II 0 +II +IV +VI degree
or air atmosphere. This apparatus, equipped with a thermobalance,
+V/2
recorded, as a function of temperature, both the weight sample
evolution (TG) and the heat flow variation between the sample Fig. 1. Main oxidation states of sulfur.
and a reference material (DSC). Samples of approximately 20 mg
were introduced in an alumina crucible and heated from 25 °C to
850 °C at a rate of 10 °C/min. Table 3
Z elemental analysis.

3. Results Composition (% weight)


Si Al Fe Mn Mg Ca Na K Ti P
3.1. Thermodesorption 21.32 13.62 0.72 0.01 0.76 0.46 10.14 0.26 0.08 0.03

Thermodesorption was performed at 350 °C in helium and air


after H2S adsorption. Results of the desorbed sulfur compound The Si/Al ratio is lower than 8–10, indicating that Z is hydrophilic
analysis are given in Table 2. [21]. The presence of metals (aluminum and sodium) can catalyze
In helium, the desorbed quantities of sulfur compounds were chemical reactions. The catalytic potential of traces of iron (0.7% in
significantly lower than the quantity of H2S adsorbed in the sample weight), magnesium (0.8% in weight) and calcium (0.5% in weight)
(Z+H2S_3). Less than 6% of the adsorbed sulfur was desorbed. cannot be excluded either. Nguyen-Thanh and Bandosz [26]
Among the quantified sulfur species, H2S was actually measured showed that the impregnation of an activated carbon with iron
in the desorption gas, but SO2 was also analyzed, which seems to salts (1.3% Fe) double its H2S adsorption capacity.
indicate that H2S has been partly oxidized during the adsorption.
In air, about 19% of the adsorbed sulfur was desorbed and the 3.3. Total sulfur content
main analyzed specie was SO2. The amount of desorbed SO2 was
multiplied by 4 compared to desorption in He flow. This could be The total sulfur content in Z was below the quantification limit
due to the presence of some little desorbable sulfur compounds (<0.01% in weight). After H2S adsorption, the sample Z+H2S_1 con-
that would be partially oxidized during the desorption process at tained 12.6% sulfur in weight, that is to say an adsorbed quantity of
high temperature and oxidizing conditions. about 126 mgS/gZ+H2S. This result is consistent with the data pre-
Thus, poor thermal desorption was achieved by heating Z+H2S sented in Table 1 (110 mgS/gZ+H2S) from GC analyses. This good
at 350 °C in helium or air flow, demonstrating that H2S is not phys- cross checking of the two mass balances suggests the formation
ically adsorbed. Consequently, the thermal regeneration of Z+H2S of strongly bonded species during the adsorption process because
seems limited. As a result, one objective of this study was to iden- no sulfur compound desorption occurred during the grinding pre-
tify the chemical nature (sulfur speciation) of the adsorbate in Z vious to the total sulfur analysis (see Section 2.4).
+H2S and to understand H2S removal mechanism.
Sulfur chemistry is complex. This element can be found in sev- 3.4. pH of the adsorbent leachate
eral oxidation states, the most commons being summed-up in
Fig. 1. In Z+H2S, sulfur can a priori be speciated into hydrogen sul- The pH of Z leachate was 10.0. After H2S adsorption, the lea-
fide – the compound initially removed from the gas – or, in case of chate became significantly more acidic: leachate pH was equal to
chemisorption, as other sulfide or hydrosulfide compounds, ele- 7.3, 7.2 and 6.6 for samples Z+H2S_1, Z+H2S_2 and Z+H2S_3
mental sulfur, thiosulfates, tetrathionates, sulfur dioxide or other respectively.
sulfur oxides, sulfites or sulfates. In the presence of water, H2S is a weak diacid with two equilib-
In the next sections, complementary physicochemical charac- rium constants [17]: pKa1 = 7.2 and pKa2 = 13.9 (Eqs. (1) and (2)).
terizations of Z and Z+H2S are reported. By gathering all the results,
a removal mechanism could be proposed and is described in the H2 SðaqÞ þ H2 OðlÞ ¡ HSðaqÞ þ H3 OþðaqÞ pKa1 ¼ 7:2 ð1Þ
Discussion part (Section 4).
HSðaqÞ þ H2 OðlÞ ¡ S2 þ
ðaqÞ þ H3 OðaqÞ pKa2 ¼ 13:9 ð2Þ
3.2. Elemental analysis
Subscripts (aq) and (l) respectively stand for aqueous and
The studied zeolite is composed – in weight – of about 21% sil- liquid.
icon, 14% aluminum, 10% sodium and oxygen (Table 3). The molar Considering the pH value of Z leachate, H2S is a priori mainly
Si/Al ratio calculated from this analysis is 1.4, which is typical of dissociated into HS (at least at the beginning of the adsorption
zeolites X (within the range 1–1.5) [21]. This Z is a Faujasite NaX. test). The occurrence of H2S (physical adsorption) is possible at

Table 2
Sulfur compound quantities thermodesorbed from Z+H2S_3 at 350 °C in helium and air – comparison with the quantity of H2S adsorbed in the sample.

Compound Adsorbed quantity Thermodesorbed quantity


In helium In air
mgS/gZ+H2S mgcompound/gZ+H2S mgS/gZ+H2S mgcompound/gZ+H2S mgS/gZ+H2S
SO2 0 16.9 8.4 67.9 34.0
H2S 189 2.2 2.1 1.5 1.4
Sum 189 10.5 35.4
50 L. Sigot et al. / Chemical Engineering Journal 287 (2016) 47–53

the end of the dynamic adsorption test, when the pH significantly Table 4
acidifies. The formation of sulfide ions (S2) is unlikely and their Specific surface area and porosity of Z before and after H2S adsorption.

reaction with Z metal cations (e.g. Na+, Ca2+ or Fe2+ – cf. Table 3) Sample Specific surface Micropore Main pore
to produce metal sulfides is not realistic. area (m2/g) volume (cm3/g) diameter (Å)
The presence of water in Z pores – necessary for the dissociation Z 700 0.26 9.5
– is proved in Sections 3.7 and 3.8. The local pH in the porous net- Z+H2S_3 440 0.17 8.8
work has a significant effect on the speciation of the final products.
It seems to control the retention mechanism involved in the pores:
a pH variation can induce a diversity of products [15–17]. Note 0.35
that, in the presence of water, biogas is an acidic gas because it
is composed of an important part of CO2. Consequently, the varia- 0.3 Z

Differenal pore volume (cm3.g -1.Å-1)


tion of pH conditions in Z bed might be faster in biogas than in N2.
Z+H2S_3
0.25

3.5. Total soluble sulfur and sulfate ions in leachate 0.2

The main objective of this analysis was to evaluate the solubil- 0.15
ity of the sulfur specie(s) formed in Z+H2S.
Measurements revealed a quantity of total sulfur in solution of 0.1
about 0.07 mgS/gZ for Z sample and 15.1 mgS/gZ+H2S for Z+H2S_2
sample. The content of total soluble sulfur is naturally higher after 0.05

H2S adsorption. However, the quantity of total soluble sulfur mea-


0
sured in Z+H2S_2 leachate is only 10% of the quantity adsorbed, 5 7 9 11 13 15
with one fifth as SO2 4 (Fig. 2). Sulfur compounds present in Z Pore diameter (Å)
+H2S are rather insoluble: used leaching conditions conducted only
Fig. 3. Pore size distribution of Z and Z+H2S_3.
to a partial leaching. A replicate realized with Z+H2S_1 sample gave
similar results.
Insoluble sulfur species include H2S, other sulfides, tetrathion-
ates or elemental sulfur. However, the first two propositions are Furthermore, this adsorbate is well bonded to the adsorbent since
not in accordance with pH considerations (§ 3.4). it was not removed during the outgassing under vacuum at 300 °C
before the analysis (Section 2.7). In the case of activated carbon,
Adib et al. [15] consider that a desorption at 120 °C under vacuum
3.6. Specific surface area and porosity
(105 Torr) totally removes physically adsorbed compounds such
as H2S. They also point out that elemental sulfur (if produced)
The specific surface area of Z was 700 m2/g (Table 4), which is in
should be deposited on pore walls and/or block some pore
the range of literature data [20]. Its micropore volume was
entrances. Our observations are similar and indicate the possible
0.26 cm3/g. The measured pore size distribution was very narrow
formation of elemental sulfur in Z+H2S, reducing the micropore
– which is typical of zeolites – with a peak centered at 9.5 Å
volume and the pore diameter.
(Fig. 3). Theoretically, the pore diameter is 7.4 Å for 13X zeolites
resulting from the assembly of tetrahedra (SiO4 and AlO4) in cages
[21]. The kinetic diameter of H2S is 3.6 Å [27]. The diameter of Z 3.7. FTIR
pores is larger than the size of H2S molecules; consequently, H2S
can enter the pores. In the case of H2S adsorption in activated car- Z and Z+H2S_3 FTIR spectra are presented in Fig. 4. The global
bon, conclusions from the literature converge toward a major shape of Z spectrum is quite similar to the spectra presented by
activity in the pores of small diameters (<10 Å) [14,15,17,26]. Thus, Montanari et al. [28] for a 13X zeolite and by Zhdanov et al. [29]
Z porous structure seems appropriate. for several types of synthetic faujasites with different Si/Al ratios.
The specific surface area and the micropore volume measured The bands visible between 400 and 1400 cm1 (Fig. 4) are char-
after H2S adsorption were divided by 1.5 (Table 4). A part of the acteristic of bond vibrations between Si, O and Al [29,30]. It is a sig-
porous network was no longer accessible. The peak of porosity nature of Z skeleton.
was less intense and slightly shifted toward smaller diameters
(Fig. 3). This is due to the deposit of the adsorbate in Z+H2S pores.
0.7 Z+H2S_3 Z 445
957-959
180 0.6
0.04 2700-3700
160 S adsorbed 1654
Quanty of sulfur compound

Sulfate 0.5 0.03


140
Other soluble sulfur species 0.02
Absorbance

120 0.4
+ = total S in soluon
(mgs/gZ+H2S)

100 0.01
0.3 561
80 0.00
672
3400 2400 1400 742
60 0.2
40
0.1 2700-3700
20 1654

0 0.0
Solid matrix Leachate 3900 3400 2900 2400 1900 1400 900 400
Wavenumber (cm-1)
Fig. 2. Comparison of the sulfur quantity adsorbed in Z+H2S_2 sample with the
quantity of total soluble sulfur (including SO2
4 ) measured in the leachate. Fig. 4. FTIR spectra of Z and Z+H2S_3.
L. Sigot et al. / Chemical Engineering Journal 287 (2016) 47–53 51

The comparison of Z and Z+H2S spectra shows very few differ- Temperature (°C) Exothermic
ences. In particular, no S-H stretching vibration characteristic of 0 100 200 300 400 500 600
H2S is observed between 2500 and 2625 cm1 on Z+H2S spectrum
0 3,500
3,000
(Fig. 4) [18,31]. This vibration is not visible either because its inten- -5
m1_Z+H2S 2,500
sity is too low in comparison to the signal baseline (weak stretch)

DSC : Heat flux (mW/g)


-10

TG : Weight loss (%)


or because there is no (or few) S–H bonds, which means that sulfur 2,000
is not mainly speciated as H2S. Note that if elemental sulfur was -15 1,500
formed as suggested in Section 3.6, it could not be highlighted -20 m2_Z+H2S 1,000
because it does not have a recognizable signature in FTIR. 500
The large band at 2700–3700 cm1 and the peak centered at
-25
0
1654 cm1 (Fig. 4) are representative of OH bond vibrations -30
-500
(stretching and bending respectively). These bonds can be associ-
-35 -1,00
ated with water signature. After H2S adsorption, these vibrations 0.0
are more intense, which suggests the formation of water. A chem-

dTG : (%/min)
ical reaction producing water can be supposed. -0.5

-1.0

3.8. TG-DSC -1.5

Samples of Z and Z+H2S were analyzed in replicates by TG-DSC -2.0

in nitrogen and air. Profiles are presented in Figs. 5 and 6. Fig. 6. TG-DSC profiles and dTG for Z+H2S_3 in air.

3.8.1. In nitrogen This weight loss (Dm2_Z+H2S) can be attributed to the presence
TG curve of Z shows a weight loss (Dm1_Z) associated with a of elemental sulfur [8,16]. As it is quite spread around the sulfur
broad endothermic peak between 100 °C and 440 °C (Fig. 5). This boiling point (444.6 °C), it is possible that sulfur is present as sev-
observation can be related to the vaporization of bonded water eral chemical forms: radicals or different polymers [16]. During
contained in Z pores. Yang [20] stated that a temperature of adsorption tests, the accumulation of sulfur atoms may conduct
350 °C is needed to desorb water from zeolites. The same observa- to the formation of more stable linear or cyclic sulfur molecules
tion is noticed for Z+H2S_3 sample but at lower temperatures, such as S8. The desorption peak is then larger because different
between 100 and 300 °C (Dm1_Z+H2S). This shift toward lower tem- kinds of elemental sulfur based compounds are present [17]. This
peratures could be explained by the beginning of a second phe- second weight loss cannot be assigned to H2S because if it was
nomenon hiding the end of the first one. Another possibility is physically adsorbed, it would have desorbed at temperatures lower
that water may be ‘‘more accessible” after adsorption (lower tem- than 120 °C [8,16]. Likewise, the presence of SO2 would be charac-
perature required for the vaporization). The quantity of water after terized by a weight loss centered at 250 °C. No such observation is
H2S adsorption is higher than the one before adsorption (Fig. 5). As highlighted in Fig. 5; consequently, neither H2S nor SO2 seems sig-
suggested by FTIR analyses (§ 3.7), water formation during adsorp- nificantly retained in Z+H2S. The weight loss cannot either be
tion seems confirmed. The analysis of another sample (Z+H2S_2) assigned to sulfides and sulfates which degrade at temperatures
showed that the quantity of additional water seems related to higher than those studied here.
the H2S uptake in the sample. It is also possible that this increase The weight loss (Dm2_Z+H2S) for Z+H2S_3 is 176–197 mg/gZ+H2S
in weight loss is partly due to some molecular H2S or to SO2 (Fig. 5) depending on the replicate (5 replicates). The amount of
desorbed from the Z+H2S sample, but H2S and SO2 quantities are H2S adsorbed in this sample is about 189 mgS/gZ+H2S. Considering
probably negligible, as shown in Section 3.1. the formation of elemental sulfur only, the mass balance is satis-
A second weight loss (Dm2_Z+H2S) associated with an endother- factory. H2S removal by Z seems actually based on adsorption–
mic peak between 300 and 630 °C is observed for Z+H2S sample in oxidation mechanisms. Involved reactions do not form oxidation
nitrogen (Fig. 5). It is correlated to H2S adsorption in Z+H2S as it is products easy to desorb but elemental sulfur strongly bonded to Z.
absent on Z sample. The presence of a unique weight loss indicates Note that a small endothermic peak is visible around 115 °C
that the system is relatively homogenous with the presence of a corresponding to the sulfur fusion point (113–119 °C).
single major compound [16]. Given the range of temperature, this
compound is difficult to desorb. 3.8.2. In air
Z+H2S_3 sample was also analyzed in air (Fig. 6). Derivative
thermogravimetric curve (dTG) is also given in Fig. 6.
Temperature (°C) Exothermic
0 100 200 300 400 500 600 700 800 TG curve shape is similar to the one obtained in nitrogen with
0 1200 two successive weight losses. The first weight loss (Dm1_Z+H2S)
m1_Z
assigned to bonded water is about 107–132 mgwater/gZ+H2S (2 repli-
-5 TG_Z+H2S 900 cates) which is in the range of measurements performed in N2
m1_Z+H2S
(102–149 mgwater/gZ+H2S). The second weight loss (Dm2_Z+H2S) is
DSC : Heat flux (mW/g)
TG : Weight loss (%)

TG_Z
-10 600
DSC_Z+H2S shifted toward lower temperatures (between 290 °C and 400 °C).
-15 DSC_Z 300 The weight loss in air (Dm2_Z+H2S = 99–111 mg/gZ+H2S) is signifi-
m2_Z+H2S
cantly lower than the one in N2 (176–197 mg/gZ+H2S), but this
-20 0 remains unexplained.
DSC profile in air is significantly different: an intense exothermic
-25 -300
peak (centered at 355 °C) is associated with the second weight loss.
This heat release is probably related to the oxidation (in the
-30 -600
calorimeter) of the elemental sulfur formed in Z+H2S. Adib et al.
Fig. 5. Comparison of TG-DSC profiles for Z and Z+H2S_3 in nitrogen. [15] observed a similar peak at 347 °C and drew the same
52 L. Sigot et al. / Chemical Engineering Journal 287 (2016) 47–53

conclusion. In air, elemental sulfur seems oxidized (in the calorime- shown by TG-DSC. Water molecules association via hydrogen
ter) before reaching its boiling point (444.6 °C). This hypothesis was bonds conducts to the formation of a water film in the pores.
confirmed by the TG-DSC analysis of a sample of pure sulfur in air The step (c) raises once again the importance of water, but also
that also showed an exothermic peak centered at 360 °C associated the role of pH. Water is indispensable for H2S dissociation provided
with a total weight loss. pH allows it. A pH value included between the two H2S acidity con-
Considering that H2S is retained as elemental sulfur in Z+H2S, stants (pKa1 = 7.2 and pKa2 = 13.9) enables the formation of HS.
the oxidation reaction is given by Eq. (3). This is the case for the studied Z, which leachate pH is 10.0. Note
that after adsorption, the environment becomes more acidic: the
SðsÞ þ O2ðgÞ ! SO2ðgÞ ð3Þ pH of the leachate (6.6–7.3) tends toward the inferior limit of
HS predominance zone. pH has a significant role in the final spe-
Subscripts (s) and (g) correspond respectively to solid and gas.
ciation of sulfur. pH is related to the adsorbent composition. The
The associated standard enthalpy of reaction DHreaction is equal
presence of calcium and magnesium oxides can contribute to the
to the SO2 standard enthalpy of formation, that is DHreaction =
alkalinity of Z.
296.8 kJ/mol at 25 °C [32]. As a first approximation, the influence
The step (d) underlines the role of adsorbed oxygen in the adsor-
of the temperature is neglected. This enthalpy of reaction
bent. Nguyen-Thanh and Bandosz [26] stated that the lack of avail-
corresponds to the oxidation of one mole of elemental sulfur i.e.
able oxygen can stop the oxidation well before all the pores are
DHreaction = 9.3 kJ/gS of elemental sulfur.
filled with sulfur. In our study, the main source of oxygen is the
The integration of the exothermic peak (Fig. 6) indicates a heat
gas matrix used for the adsorption tests that is composed of approx-
release of 1.06 J/gZ+H2S, that is to say, by dividing by the associated
imately 1% oxygen in volume (see Section 2.1). In addition, the pres-
weight loss Dm2_Z+H2S, an experimental enthalpy of 9.5 kJ/glost.
ence of metals in Z is able to catalyze H2S oxidation reaction.
The energy balance (comparison of DSC result with DHreaction) is
The final speciation of sulfur depends on all these conditions.
convincing: elemental sulfur is then identified as the major specie
TG-DSC analyses confirmed that elemental sulfur is the main spe-
formed in Z+H2S. The presence of the small endothermic peak asso-
cie formed in Z. Note that sulfur itself is also a catalyst of H2S oxi-
ciated with sulfur fusion can be seen again on Fig. 6. Note that the
dation [10] that is to say the more sulfur is adsorbed, the more H2S
standard enthalpy of reaction associated with H2S oxidation
oxidation is promoted.
(Eq. (4)) is equal to 15.2 kJ/gH2S at 25 °C. This confirms that the
Finally, elemental sulfur can arrange in more stable linear or
presence of molecular H2S is not likely in Z+H2S.
cyclic polymers (step (e)) [14,17]. In standard temperature and
H2 SðgÞ þ 3=2O2ðgÞ ! SO2ðgÞ þ H2 OðgÞ ð4Þ pressure conditions, the most stable allotropic sulfur specie is
cyclo-octasulfur (S8 – practically insoluble in water).
The formation of a minor part of SO2 cannot be excluded but the
quantity of formed SO2 is negligible (see Section 3.1). Indeed, TG-
4. Discussion: H2S removal mechanism by Z DSC results show that S mainly oxidizes in SO2 at temperature
higher than ambient temperature.
The global mechanism of H2S retention in the studied Z is a In the context of biogas purification, this mechanism may be
combination of ‘‘adsorption–dissociation–oxidation”. Thanks to affected by the presence of other kinds of compounds (especially
the different characterizations performed, the following successive CO2) and humidity. CO2 (about 40% v/v in biogas) could be partly
steps are proposed to describe the mechanism: adsorbed and may affect the adsorption capacity of H2S. Indeed, zeo-
lites can be used for biogas upgrading (CO2 removal) [34,35]. A con-
a. H2S adsorption at Z surface: H2S(g) ? H2S(ads) sequence of this carbonation would be a rapid pH decrease and H2S
b. H2S dissolution in water contained in Z pores: might be retained as molecular H2S (pKa1 = 7.2, see Section 3.4). Yet
H2S(ads) ? H2S(aq) some metals from Z structure might solubilize, which would
c. H2S dissociation in the water film: increase their reactivity with H2S. A deficit of O2 could negatively
H2S(aq) + H2O(l) ? HS +
(aq) + H3O(aq) affect the removal of H2S. As concerns humidity, some tests were
d. Oxidation of HS with adsorbed oxygen: performed at laboratory with a humid N2 matrix and the adsorption
HS 
(aq) + O(ads) ? S(ads) + OH(aq) capacity of H2S was reduced because of Z affinity for water.
e. Formation of linear or cyclic sulfur polymers: xS(ads) ? Sx(ads)

Subscript (ads) corresponds to adsorbed compounds. O(ads) rep- 5. Conclusion


resents dissociatively adsorbed oxygen. S(ads) and Sx(ads) symbolize
elemental sulfur in simple or polymeric form. Characterizations of Z before and after H2S adsorption led to
The global reaction leads to the formation of elemental sulfur several key observations:
and water as proposed in Eq. (5). – poor thermal desorption at 350 °C indicating there is no physi-
cal adsorption of H2S,
H2 S þ 0:5O2 ! S þ H2 O ð5Þ
– presence of water and pH promoting H2S dissociation in HS,
The first step (a) underlines the importance of Z physical char- – adsorbate almost insoluble,
acteristics (specific surface area, pore volume, pore size) and gas – adsorbate deposit in the pores reducing the specific surface
dynamics in the adsorption capacity and probably the adsorption area, pore volume and pore diameter,
kinetics. – formation of water during adsorption,
The second step (b) highlights the role of water contained in Z – no S–H bonds,
pores, as also found with other adsorbents [13,15,33]. The amount – degradation of the adsorbate between 300 and 630 °C in inert
of water must be sufficient but not too important to prevent pores atmosphere,
from filling with condensed water. The dissolution step is also – oxidation of the adsorbate around 355 °C in air atmosphere.
affected by the adsorbent morphology. Indeed, pores must be suf-
ficiently small to allow the formation of a water film even at low The results of the different characterizations converged to pro-
humidity [15], but sufficiently large to contain the adsorbate and pose a removal mechanism of H2S in Z based on adsorption–oxida-
oxygen molecules. The presence of bonded water in Z pores was tion and formation of elemental sulfur.
L. Sigot et al. / Chemical Engineering Journal 287 (2016) 47–53 53

From an operational point of view, the formation of stable sulfur [14] F. Adib, A. Bagreev, T.J. Bandosz, Effect of surface characteristics of wood-based
activated carbons on adsorption of hydrogen sulfide, J. Colloid Interface Sci.
polymers compromise the thermal regeneration of Z but the
214 (1999) 407–415.
understanding of this mechanism should facilitate the develop- [15] F. Adib, A. Bagreev, T.J. Bandosz, Effect of pH and surface chemistry on the
ment of efficient regeneration processes. For instance, zeolite mechanism of H2S removal by activated carbons, J. Colloid Interface Sci. 216
regeneration by high temperature desorption at 500–600 °C in (1999) 360–369.
[16] A. Bagreev, T.J. Bandosz, H2S adsorption/oxidation on unmodified activated
inert atmosphere (but this process is very energy consuming and carbons: importance of prehumidification, Carbon 39 (2001) 2303–2311.
high temperature might damage Z crystalline pore structure) or [17] R. Yan, T. Chin, Y.L. Ng, H. Duan, D.T. Liang, J.H. Tay, Influence of surface
by desorption–oxidation in air at a temperature <400 °C could be properties on the mechanism of H2S removal by alkaline activated carbons,
Environ. Sci. Technol. 38 (2004) 316–323.
investigated. Desorption under reduced pressure to sublime the [18] C.L. Garcia, J.A. Lercher, Adsorption of H2S on ZSM5 zeolites, J. Phys. Chem. 96
formed elemental sulfur is another option. Chemical dissolution (1992) 2230–2235.
with organic solvents is also possible but these solvents are not [19] L. Sigot, G. Ducom, P. Germain, Adsorption of octamethylcyclotetrasiloxane
(D4) on silica gel (SG): retention mechanism, Microporous Mesoporous Mater.
environmentally friendly. 213 (2015) 118–124.
[20] R.T. Yang, Adsorbents Fundamentals and Applications, John Wiley & Sons,
Hoboken, New Jersey, 2003.
Acknowledgements [21] D.M. Ruthven, Principles of Adsorption and Adsorption Processes, John Wiley &
Sons, New York, 1984.
This study, part of the PILE-EAU-BIOGAZ project, was supported [22] EN 12457-2:2002, Characterisation of waste – Leaching – Compliance test for
leaching of granular waste materials and sludges – Part 2: One stage batch test
by the French Agence Nationale de la Recherche (ANR), program H- at a liquid to solid ratio of 10 l/kg for materials with particle size below 4 mm
PAC 2010. (without or with size reduction), 2002.
[23] NF EN ISO 11885:2009, Water quality – Determination of selected elements by
inductively coupled plasma optical emission spectrometry (ICP–OES), 2009.
References [24] B. Keller-Lehmann, S. Corrie, R. Ravn, Z. Yuan, J. Keller, Preservation and
simultaneous analysis of relevant soluble sulfur species in sewage samples, in:
[1] R.J. Spiegel, J.L. Preston, J.C. Trocciola, Fuel cell operation on landfill gas at Proceedings of the Second International IWA Conference on Sewer Operation
Penrose Power Station, Energy 24 (1999) 723–742. and Maintenance, Vienna, Austria, October 2006.
[2] S.P. Hernández, F. Scarpa, D. Fino, R. Conti, Biogas purification for MCFC [25] NF EN ISO 10304-2:1996, Water quality – Determination of dissolved anions
application, Int. J. Hydrogen Energy 36 (2011) 8112–8118. by liquid chromatography of ions – Part 2: Determination of bromide, chloride,
[3] G. Monteleone, M. De Francesco, S. Galli, M. Marchetti, V. Naticchioni, Deep nitrate, nitrite, orthophosphate and sulfate in waste water, 1996.
H2S removal from biogas for molten carbonate fuel cell (MCFC) systems, Chem. [26] D. Nguyen-Thanh, T.J. Bandosz, Activated carbons with metal containing
Eng. J. 173 (2011) 407–414. bentonite binders as adsorbents of hydrogen sulfide, Carbon 43 (2005) 359–
[4] N. de Arespacochaga, C. Valderrama, C. Mesa, L. Bouchy, J.L. Cortina, Biogas 367.
deep clean-up based on adsorption technologies for Solid Oxide Fuel Cell [27] Y. Sun, S. Han, Diffusion of N2, O2, H2S and SO2 in MFI and 4A zeolites by
applications, Chem. Eng. J. 255 (2014) 593–603. molecular dynamics simulations, Mol. Simul. 41 (2015) 1095–1109.
[5] L. Micoli, G. Bagnasco, M. Turco, H2S removal from biogas for fuelling MCFCs: [28] T. Montanari, E. Finocchio, I. Bozzano, G. Garuti, A. Giordano, C. Pistarino, G.
new adsorbing materials, Int. J. Hydrogen Energy 39 (2014) 1783–1787. Busca, Purification of landfill biogases from siloxanes by adsorption: a study of
[6] D. Papurello, A. Lanzini, P. Leone, M. Santarelli, S. Silvestri, Biogas from the silica and 13X zeolite adsorbents on hexamethylcyclotrisiloxane separation,
organic fraction of municipal solid waste: dealing with contaminants for a Chem. Eng. J. 165 (2010) 859–863.
solid oxide fuel cell energy generator, Waste Manage. 34 (2014) 2047–2056. [29] S.P. Zhdanov, T.I. Titova, L.S. Kosheleva, W. Lutz, Effect of hydrothermal
[7] L. Sigot, G. Ducom, B. Benadda, C. Labouré, Comparison of adsorbents for H2S dealumination of a synthetic faujasite by IR spectroscopy, Pure Appl. Chem. 61
and D4 removal for biogas conversion in a solid oxide fuel cell, Environ. (1989) 1977–1980.
Technol. doi: http://dx.doi.org/10.1080/09593330.2015.1063707. [30] E. Montarges-Pelletier, S. Bogenez, M. Pelletier, A. Razafitianamaharavo, J.
[8] A. Bagreev, H. Rahman, T.J. Bandosz, Thermal regeneration of a spent activated Ghanbaja, B. Lartiges, L. Michot, Synthetic allophane-like particles: textural
carbon previously used as hydrogen sulfide adsorbent, Carbon 39 (2001) properties, Colloids Surf. A 255 (2005) 1–10.
1319–1326. [31] G. Blyholder, D.O. Bowen, Infrared spectra of sulfur compounds adsorbed on
[9] T.J. Bandosz, On the adsorption/oxidation of hydrogen sulfide on activated silica-supported nickel, J. Phys. Chem. 66 (1962) 1288–1292.
carbons at ambient temperatures, J. Colloid Interface Sci. 246 (2002) 1–20. [32] D.D. Wagman, W.H. Evans, V.B. Parker, R.H. Schumm, I. Halow, The NBS tables
[10] M. Steijns, F. Derks, A. Verloop, P. Mars, The mechanism of the catalytic of chemical thermodynamic properties. Selected values for inorganic and C1
oxidation of hydrogen sulfide. II. Kinetics and mechanism of hydrogen sulfide and C2 organic substances in SI units, J. Phys. Chem. Ref. Data 11 (1982) 1–392.
oxidation catalyzed by sulfur, J. Catal. 42 (1976) 87–95. [33] M. Seredych, T.J. Bandosz, Reactive adsorption of hydrogen sulfide on graphite
[11] V. Meeyoo, D.L. Trimm, N.W. Cant, Adsorption-reaction processes for the oxide/Zr(OH)4 composites, Chem. Eng. J. 166 (2011) 1032–1038.
removal of hydrogen sulphide from gas streams, J. Chem. Tech. Biotechnol. 68 [34] A. Alonso-Vicario, J.R. Ochoa-Gómez, S. Gil-Río, O. Gómez-Jiménez-Aberasturi,
(1997) 411–416. C.A. Ramírez-López, J. Torrecilla-Soria, A. Domínguez, Purification and
[12] S.V. Mikhalovsky, Y.P. Zaitsev, Catalytic properties of activated carbons. I. Gas- upgrading of biogas by pressure swing adsorption on synthetic and natural
phase oxidation of hydrogen sulphide, Carbon 35 (1997) 1367–1374. zeolites, Microporous Mesoporous Mater. 134 (2010) 100–107.
[13] A. Primavera, A. Trovarelli, P. Andreussi, G. Dolcetti, The effect of water in the [35] T. Montanari, E. Finocchio, E. Salvatore, G. Garuti, A. Giordano, C. Pistarino, G.
low-temperature catalytic oxidation of hydrogen sulfide to sulfur over Busca, CO2 separation and landfill biogas upgrading: a comparison of 4A and
activated carbon, Appl. Catal. A 173 (1998) 185–192. 13X zeolite adsorbents, Energy 36 (2011) 314–319.

You might also like