Sei sulla pagina 1di 13

Separation and Purification Technology 210 (2019) 80–92

Contents lists available at ScienceDirect

Separation and Purification Technology


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

Biogas cleaning: Trace compounds removal with model validation T


a,b,⁎ b a
Davide Papurello , Silvia Silvestri , Andrea Lanzini
a
Department of Energy (DENERG), Politecnico di Torino, Corso Duca degli Abruzzi, 24, 10129 Turin, Italy
b
Fondazione Edmund Mach, Biomass and Renewable Energy Unit, Via E. Mach, 1, 38010 San Michele a/A, Italy

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

Keywords: High energy efficiency SOFCs generators can be adopted for local and distributed micro-generation systems to
Biochar promote the reduction of greenhouse gas emissions with their high fuel flexibility, long-term stability also at
VOCs removal partial load and low noise. One of the main drawbacks for such generators, fed by biogenous gas is the impact of
Biogas trace compounds on the anode compartment. For this reason, a gas clean-up section is mandatory. The effect of
Solid Oxide Fuel Cell (SOFC)
temperature and gas moisture was investigated through experiments on the removal performance of the tested
Carbon
Adsorption modeling
sorbents. An increase in the operating temperature caused lower values of the maximum capacity of the ad-
sorbent. The decrease of removal performance considering a humidified gas is connected to the interference of
water in the pores of activated carbons. Biochar, compared to the other commercial sorbent materials showed
the lowest removal performance, even if with activated biochar the adsorption capacity growth to commercially
available materials. The highest adsorption capacity at 1% of the initial concentration was showed by com-
mercial carbons with 1.75 mg/g for H2S and 20.4 mg/g for HCl. Experimental data were employed in a porous
particle diffusion model to estimate the breakthrough time. Low values of errors validate the model in the first
part of the breakthrough curve, even for competitive adsorption case.

1. Introduction living organisms and conversion to less harmful forms).

The renewable sources coupled to distributed micro-generation Physisorption deploys weak Van der Walls forces, whereas chemi-
systems could be a plausible scenario for the energy future exploitation sorption involves the stronger covalent and ionic bonds. Van der Walls
according to the respect of the environment. A plausible ingredient in forces require polar or polarizable compounds. Typically, compounds
this complicated recipe is the exploitation of biomass for the energy with no dipole moment (0 D) are classified as non-polar, and else as
production with high-efficiency generators, such as SOFCs. One of the polar [19].
main negative aspects of such micro-generators, fed by biogas is the Chemical absorption of sulfur in aqueous solutions involves the
impact of trace compounds on their performance. oxidation of S2− to S0 and on the capture of S2− through precipitation
Requirements are related to 1–2 ppm(v) for sulfur compounds of its metallic salts. For example, at the first method belongs the oxi-
[1–7], 20–50 ppm(v) for chlorines and tars compounds [3,8–14] and as dative absorption of H2S and O2 by iron-chelated solutions [20]. A
low as possible for siloxanes concentration [7,15–18]. These limits are detailed review of the absorption methods are reported by Abatzoglou
related to the reversible operation for SOFC systems, even if for long and Boivin [20]. Literature studies have demonstrated how all the ad-
term operation SOFC stack lifetime is shortened accordingly to the sorption processes exploit semi-batch processes. This is due to the
current density, operating temperature and trace compound con- continuous gas-stream flow and fixed bed of adsorbent which is gra-
centration [18]. dually saturated with the adsorbate. These methods are hardly em-
For this reason, a gas clean-up section is strongly required. Biogas ployed for large-scale plants mainly due to the lifespan of sorbent
purification methods can be divided into two broad categories: material and regeneration problems [19,21,22]. Despite these issues, a
possible solution could be the lead and lag configuration. This config-
• those involving physicochemical phenomena (adsorption or ab- uration ensures the continuity of service [23]. Considering, the material
sorption); adopted, iron oxides adsorbents are recognized in literature for their
• those involving biological processes (contaminant consumption by sulfur removal capacity [20]. Operational experience indicates that


Corresponding author at: Department of Energy (DENERG), Politecnico di Torino, Corso Duca degli Abruzzi, 24, 10129 Turin, Italy.
E-mail address: davide.papurello@polito.it (D. Papurello).

https://doi.org/10.1016/j.seppur.2018.07.081
Received 23 January 2018; Received in revised form 13 July 2018; Accepted 27 July 2018
Available online 29 July 2018
1383-5866/ © 2018 Elsevier B.V. All rights reserved.
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

Nomenclature NLDFT non localized density functional theory


PDMS polydimethylsiloxane membrane
1%C/C0 pass-through ratio at 1% PID proportional integrative and derivative controller
10%C/C0 pass-through ratio at 10% Q volumetric flow rate (cm3/min)
C* gas concentration at the equilibrium within the pores qe equilibrium adsorption capacity (mg/gAC)
(ppm(v)) qmax maximum adsorption capacity function of sorbent mate-
C gas bulk concentration (ppm(v)) rial and contaminant (mg/gAC) – when it is reached
C0 initial concentration (ppm(v)) C* = C(x,t)
Carbox activated carbon from Airdep RH relative humidity
co covapors or simulatenuous concentration of H2S and HCl RST3 activated carbon from Norit
cow covapors or simulatenuous concentration of H2S and HCl SEM scanning electron microscopy
and humidified gas SOFC solid oxide fuel cell
D unit of dipole moment (Debye) SSA specific surface area (m2/g)
Dax axial dispersion term T operating temperature (K)
DFT density functional theory tb breakthrough time (min)
EDS energy-dispersive X-ray spectroscopy u gas interstitial velocity (m/s)
H2S hydrogen sulfide VFUEL volumetric flow rate (SNml/min)
HCl hydrogen chloride Vmicro microporous volume (cm3/g)
kfk is the boundary layer mass transfer coefficient VOCs volatile organic compounds
kl langmuir constant function of sorbent material, con- Vtot total volume (cm3/g)
taminant and temperature (m3/mg) w relative humidity 30%
kv overall adsorption rate coefficient (min−1) We equilibrium adsorption capacity (gpollutant/gcoal)
ms sorbent mass (g) x and t are respectively spatial and time coordinates
MSE mean squared error εb bed porosity (or void fraction inter-particle)
MW molecular weight (g/mol) ρb bulk density of the carbon bed (g/cm3)

only about 85% of removal can be achieved (0.56 kgH2S/kgFe2O3), as Some studies suggest alternative solutions [29], for instance, Yuan
reported by Taylor [24]. Hence, iron oxides are mainly used for a first et al. [30] reports that the surface of sludge derived adsorbents can be
sulfur abatement section. used for the desulfurization of digester gas. The removal capacity is
VOCs removal by adsorption on activated carbons represents a valid comparable to catalytic activated carbons [30].
and economical solution [25]. The main properties of activated carbon Another alternative solution is the use of biomass or waste derived
filters for the effective removal of VOCs are high porosity, high super- materials char from pyrolisis or gasification processes [31,32]. In Hervy
ficial area (1500 m2 g−1), high volume, pore distribution [26] and et al.[31], two chars were produced by the pyrolysis of used wood
treatment with metal ions. Impregnated activated carbons, mainly with pallets, and a 50/50% mixture of food waste and coagulation-floccu-
caustic bases (NaOH, Na2CO3, NaHCO3, KI), show typical H2S loading lation sludge. The most efficient material was the steam activated char
capacities around 0.15 g/g [20]. Barelli et al. showed how the choice of from food waste and coagulation-flocculation sludge, with a removal
the impregnating agent is important in the removal of sulphur com- capacity of 65 mgH2S/g under dry syngas [31].
pounds. In fact, when the carbon is activated with KI and KOH, the role In summary, a first roughing step for the trace compounds limitation
of water on the adsorption capacity has a positive effect [27]. In Pa- should be followed by at least two guard beds to ensure the SOFC fuel
purello et al. this effect was not highlighted due to the absence of such requirements and continuity of operation. The first step could be cov-
agents and the role of water remained negative [28]. The non-im- ered by an absorption process or by an iron oxides bed while the ultra-
pregnated activated carbon show H2S-loading capacities around 0.02 g/ filtration process can be ensured by solid sorbents, such as activated
g [20]. carbons or similar [20,33–35].

Mass spectrometer

Blank line
Heated trap
Biogas (CH4/
PDMS filter
CO2)
P-4
+
H2S/HCl
+
H 2O AC filter

Fig. 1. Experimental set-up.

81
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

Table 1 Table 2
Elements identified with SEM-EDS analysis for the samples tested. Surface and volume analysis.
Atomic % Virgin Tested Sample SSA (m2/g) Vtot (cm3/g) Vmicro (cm3/g)

Element Carbox RST3 Biochar Carbox RST3 Biochar Carbox 1237 (Langmuir) 0.407 0.328
RST3 1117 (Langmuir) 0.447 0.300
C 80.9 91.15 98.85 86.96 93.83 99.65 Biochar 75 (Langmuir) 0.04 0.02
O 9.01 5.3 8.37 4.6 Biochar activated 593 (Langmuir) 0.26 0.17
Mg 1.19 0.2 0.1 0.08
Al 1.98 0.68
Si 1.4 1.03
shaker (Fritsch, Germany).
P 0.81
S 0.11 0.57 0.27 0.72 The biochar was activated in temperature to increase the starting
Cl 0.33 0.63 active surface area. The biochar sample was placed in a furnace at
K 0.78 0.78 0.34 0.21 0.37 0.11 750 °C (500 We, C.I.T.T., Milan, Italy), in a quartz tube with id 6 mm
Ca 1.65 0.76 0.8 1.03 0.4 0.24 and od 10 mm, length 60 cm. The sample was fluxed with carbon di-
Mn 0.41
oxide, 100 Sml/min for 1 h.
Fe 1.43 0.72
Zn 1.15 Data from the VOCs removal are used to calculate the adsorption
Total: 100 100 100 100 100 100 capacity for sulfur and chlorine using the following formula:
tt
C0 (H2 S, HCl) . 60 1
It is crucial to predict well the breakthrough time to foresee when
qe =
24.45
·MW ·VFUEL· 6 · ·
10 ms
∫ ⎛1− CC0 ⎞ dt
⎜ ⎟

0 ⎝ ⎠ (1)
the sorbent material needs to be changed. In this article, a porous
particle diffusion model was considered to find the equilibrium con-
2.3. Physicochemical characterization of sorbent materials
centration for two main contaminants contained in the biogas: H2S and
HCl. Experimental tests validate this model.
The physicochemical features of sorbent materials were carried out
in order to individuate the active surface area, the total volume, the
2. Material and methods
microporous volume and the pore distribution. Elemental composition
measurements were performed by scanning electron microscopy (SEM)
2.1. Materials
(FEI Inspect, Philips 525 M, The Netherlands) coupled with EDS ana-
lysis (SW9100 EDAX, The Netherlands). Characteristics of tested sor-
The materials tested are two different commercial activated carbons
bent materials are given in Table 1 and 2.
and a char obtained from the pyrogasification of forestry wood. Norit
The adsorption isotherms for N2 at 77 K were determined using a
RST3 (Cabot corporation, USA) is 3 mm extruded; steam activated
Quantachrome Autosorb 1. Samples were outgassed at 423 K overnight
carbon with catalytic properties, impregnated with calcium hydroxide
before the adsorption measurements. The experimental equipment al-
and potassium hydroxide. Airdep Carbox (Airedepuration plants, Italy)
lows measurement of the relative pressure until 10−6. Langmuir
is an extruded activated carbon with water vapor and impregnated with
equation has calculated specific surface areas for the carbon-based
iron oxide, calcium oxide and potassium oxide. Biochar is obtained
samples in the relative pressure range 0.04–0.1. Micropore volumes
from the pyro-gasification of wood from the waste sorting. The pyro-
were determined by means of the t-plot method in the relative pressure
gasification plant is realized near Turin, (Biesse di Leinì, Italy) with an
range 0.15–0.3. In this method the amount adsorbed is plotted against t,
electrical power of 400 kW and thermal power of 600 kW.
the multilayers thickness calculated from the standard isotherm ob-
tained with a non-porous solid reference. Carbon black was used as
2.2. Experimental apparatus
non-porous reference material for the t-plot.
For carbon-based materials the pore size has been evaluated
The experimental apparatus consists of a teflon tube piping (PTFE)
through the DFT method (Density Functional Theory), using the NLDFT
(i.d. 6 mm), a glass filter with a heated trap and mass spectrometer
equilibrium model for slit/cylindrical pores. The pore distribution for
instrument. Into the glass filter is located the sorbent material, main-
the two different commercial activated carbons, and biochar is reported
taining L/D value constant to 6.6 and a GHSV value constant to
in the following figure. A multimodal distribution is depicted due to the
1636 h−1. This is placed inside an electric oven (500 We, C.I.T.T.,
presence of several families of micropores (below 2 nm). The small
Milan, Italy), where the temperature is controlled at 30 °C by a PID
contribution above 2 nm is ascribed to inter-particles mesopores (see
regulator (Horst, Germany). The sorbent materials were tested with
Fig. 2).
simulated biogas (CH4/CO2 = 1.5) and pollutants such as, H2S con-
centration in the range 10–200 ppm(v), and HCl concentration in the
2.4. Adsorption model theory
range 10–377 ppm(v). Known concentrations of pollutants were avail-
able from certified gas cylinder (Siad spa, Italy). Demineralized water
The following section describes the assumptions and equations used
was co-fed to the mainstream using a liquid mass flow controller and a
for calculating the adsorption of selected contaminants on activated
controlled evaporator mixer (Bronkhorst, The Netherlands). Relative
carbons.
humidity was fixed at 30%, and it was changed from 0 to 100% when
A porous particle diffusion model, which involves the material
the impact of RH was investigated in the gas mixture. The biogas
balance equations in both of the gas-phase and the pore phase, was
mixture was fed to the micro-reactor through mass flow controllers
proposed to simulate the adsorption process [26]. The model adopted is
(Bronkhorst, The Netherlands).
composed by the mass balance equation for the bulk phase in a packed
Fig. 1 depicts the experimental set-up adopted. Red color represents
bed, the mass balance equation within the particle, the isotherm
heated lines at 40 °C with heated strings (isopad Thermocoax, Ger-
equation to describe the adsorption capacity of material and the ideal
many) controlled via a PID regulator (Horst, Germany). A PDMS
adsorption solution theory to consider the competitive adsorption.
(20 μm) membrane was inserted between the filter line and the heated
Several assumptions were made:
trap, to protect the mass spectrometer from the carbon particles. The
mass spectrometer adopted was an HPR 20 (Hiden Ltd., UK). The sor-
1. the adsorption process is isotherm, the temperature of the process is
bent materials were grounded up to 100–180 μm with a vibratory sieve

82
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

9.E-02

8.E-02

7.E-02

Pore size distribution (cc/nm/g)


6.E-02

5.E-02

4.E-02

Carbox
3.E-02 RST3
Biochar
2.E-02

1.E-02

0.E+00
6 9 14 22 35 54 85 134 210 329
Pore width (nm)
Fig. 2. Pore size distribution.

constant and the reaction of adsorption is exothermic meaning that for packed bed [38]. To obtain the rate of adsorption and complete the
the surface is heterogeneous; equation is also necessary to evaluate C*. This term is the gas con-
2. the axial dispersion is considered only longitudinally and not ra- centration at the equilibrium with qmax, or rather the contaminants
dially; quantity adsorbed on porous media. The evaluation of this parameter is
3. the adsorbent particles are spherical and homogeneous in size and obtained using the Langmuir isotherm.
density; The transport phenomenon on porous media can be simplified
4. the external transfer coefficient depicts the mass transfer across the considering two main mechanisms:
boundary layer;
5. the intra-particle mass transport is characterized by the effective • external transport, transport from gas bulk phase to gas film around
pore diffusion coefficient; the particle,
6. the linear velocity of the gas phase is independent of the con- • internal porous transport, it represents the resistance to diffusion of
centration; a contaminant in the intra-particle porosity.
7. a local equilibrium condition is established between the gas con-
centration adsorbed and the solid particle. For activated carbons, the linear adsorption equilibrium cannot be
chosen. The Langmuir isotherm is adapted to describe the non-linear
A similar model was developed by Rosen and by Rasmuson [36,37]. equilibrium behavior of sorbent material [39]. To estimate these
The model adopted is validated with experimental tests. The equation parameters, laboratory tests must be developed. C* is the solution of a
of the model is a partial differential equation in space and time. It is non-linear equation to quantify the rate of material adsorption. The
developed for a vertical dimensional adsorption layer. estimation is performed solving the Langmuir equation (Eq. (4).

∂2C (x , t ) ∂C (x , t ) ∂C (x , t ) 1−εb ∂q (x , t ) kl C ∗
−Dax +u + + =0 qe = qmax
∂x 2 ∂x ∂t εb ∂t (2) 1 + kl C ∗ (4)

The axial dispersion term is an undesirable effect. The aim of To evaluate the adsorption capacity for a mixture the ideal adsorbed
manufacturers and designers is to avoid this phenomenon compared to solution approach (IAS) is followed [40].
the advection term. This contribution can be neglected due to the lower Considering two-compound competitive adsorption the IAS equili-
Peclet number (around 500). This is generally attained by the gas flow brium conditions can be expressed using LeVan and Vermeulen equa-
across the adsorbent media. The accumulation term takes into account tions [41]. The following equation is adopted for the competitive ad-
the mass transfer of the contaminant from gas bulk phase to solid phase. sorption of two compounds.
It represents the rate of adsorption and it contains the transport kinetics
q1, m kl,1 C1 + q2, m kl,2 C2
of the contaminants, but also equilibrium conditions. These conditions qe =
are related to the type of isotherm chosen and related coefficients. The kl,1 C1 + kl,2 C2 (5)
necessity to know porous media characteristics are crucial to develop
The Langmuir equations of two compounds can be expressed with
mass balance within the particle. This term can be written as:
the following equations:
∂q (x , t )
= a (1−εb ) kfk (C −C ∗) qs kl, i Ci
∂t (3) qi = + Δ1,2
1 + ∑i = 1,2 kl, i Ci (6)
Boundary layer mass transfer coefficient (kfk) is calculated using the
empirical formulation of Wakao-Funazkri, as suggested by Xiao et al. where i = 1,2; and Δ1,2 can be expressed using (Eq. (2):

83
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

kl,1 C1 kl,2 C2 (v) for H2S (C0 = 76 ppm(v)) and 3.7 ppm(v) for HCl (C0 = 377 ppm
Δ1,2 = (q1, m−q2, m ) ln(1 + kl,1 C1 + kl,2 C2)
(kl,1 C1 kl,2 C2 )2 (7) (v)). These values are considered representatives of the reversible pol-
lutant concentration for SOFC applications, as reported elsewhere
Another possible model to be considered is based on the Wheeler- [4,9,47]. In the meantime, 10%C/C0 represents the irreversible condi-
Jonas equation. The Wheeler–Jonas equation allows the estimation of tion for SOFC applications. Above this value, the fuel feeding condition
the breakthrough time of granular activated carbon (GAC) bed [28,42]. can only generate irreversible issues for SOFCs. Hydrogen sulfide re-
This estimation is based solely on measurable and readily available moval with RST3 shows a breakthrough time (tb) with a pass-through
macroscopic parameters: ratio of 1%, around 31 min. For carbox the tb is close to 62 min. At
We·ms ρb ·We C −C 10%C/C0, the tb is 39 min and 112 min for RST3 and carbox, respec-
tb = − ·ln 0
Q·C0 k v·C0 C (8) tively. With these conditions, the sorbent material should be changed
between 1%C/C0 and 10%C/C0 in 8 min for RST3 and 50 min for
This equation is originally based on a continuity equation of the carbox. Carbox shows a more sigmoid behavior compared to RST3.
mass balance. This equation considers the mass balance between the Taking into account biochar as a sorbent material, the tb is close to
vapor entering the carbon bed and the sum of the amount adsorbed by 1.5 min and 2 min considering, respectively the 1% and 10% pass-
this bed plus the amount penetrating through it. through ratio. This material without any activation or impregnation
There are some limitations to this equation: process is useless for H2S removal. Hydrogen chloride removal with
RST3 shows a tb with a pass-through ratio of 1% around 44 min, while
• the flow pattern has to be a perfect plug flow with axial, but no for carbox the tb is close to 34 min. At the 10%C/C0 condition the tb for
radial diffusion; RST3 is 47 min and 37 min for carbox. With these conditions, the sor-
• the original equation was based on physisorption into micropores; bent material should be changed between 1%C/C0 and 10%C/C0 in
• the rate constant k has to be of first order concerning the number of
v 3 min for RST3 and 3 min for carbox. Taking into account biochar as a
gas molecules. For pure physisorption, this is usually true in the sorbent material, the tb is close to 1.6 min and 6.7 min considering,
first, convex part of the sigmoidal breakthrough curve. respectively the 1% and 10% pass-through ratio. This material without
any activation or impregnation process is useless also for HCl removal
As reported by several authors, this equation is reliable especially (see Fig. 3).
considering the single compound removal. In case of multiple con- The adsorption capacity evaluated at the starting of the break-
taminants, IAS approach must be adopted [42-46]. through (C/C0 = 1%) by Eq. (1) is reported in the following table:
The adsorption capacity evaluated with activated biochar show
3. Results and discussion comparable performance with commercial samples.
The removal performance in case of the single compound shows a
3.1. Removal of single compound linear correlation with the specific surface area, total volume and mi-
cropore volume. The correlation between adsorption capacity and mi-
To describe the pollutant removal using sorbent materials a pass- cropore volume is stronger with respect to specific surface area and
through ratio value (C/C0) is used. C, is the temporal evolution of the total pore volume. This is in line with the typology of the adsorbed
concentration along the filter. This term is evaluated at the end of the molecules, as reported in [48]. In the following figure are reported the
filter. C0, is the starting concentration to be removed from the filter. equations for the correlation between adsorption capacity and sorbent
This term is evaluated at the beginning of the filter. C/C0 identifies the characteristics (see Fig. 4).
concentration state of the gas flow at the output of the filter. The pass- There is a strong relationship between the adsorption capacity and
through ratio at 1% (1%C/C0) concentration corresponds to 0.76 ppm

100

90

80

70

60
C/C0 (%)

50

40 RST3 - H2S
carbox - H2S
30 biochar - H2S
RST3 - HCl
carbox - HCl
20
biochar - HCl

10

0
0 50 100 150 200 250 300
Time (min)
Fig. 3. Single compound removal – breakthrough time.

84
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

2
H2S Carbox
1.8
RST3
1.6
1.4

Ads capacity (mg/g)


y = 0.0014x - 0.0347
1.2
Biochar act. R² = 0.9874
1
0.8
0.6
0.4
0.2 Biochar
0
0 200 400 600 800 1,000 1,200 1,400
S BET (m2/g)
2
H2S Carbox
1.8
RST3
1.6
1.4
Ads capacity (mg/g)

1.2 y = 4.1318x - 0.132


Biochar act. R² = 0.9586
1
0.8
0.6
0.4
0.2 Biochar
0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50

V tot (cm3/g)
2
H2S Carbox
1.8
RST3
1.6
1.4
Ads capacity (mg/g)

1.2 y = 5.4967x - 0.0641


Biochar act. R² = 0.9941
1
0.8
0.6
0.4
0.2 Biochar
0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35

V micro(cm3/g)
Fig. 4. Relationship between adsorption capacity and physical characteristics.

physical characteristics of sorbent materials. Specific surface area and 3.3. Concentration influence on removal performance
pore volumes are directly related with the adsorption capacity of the
sorbent. Two different commercial activated carbons are tested.
Experimental results data for the breakthrough time varying the
starting pollutant concentration for H2S and HCl are reported in Fig. 6.
3.2. Impact of the activation process on the removal performance of biochar Breakthrough time values modeled with Eq. (1) are in agreement with
experimental results, both in case of H2S, HCl and also for two different
Carbon dioxide was selected as activating agent. This selection commercial activated carbon filters. The results confirm a marked de-
wants to simulate an alternative use of a gas exhaust from SOFCs cline in the tb increasing the initial concentration of the contaminant. In
generators. The activation process increased the specific surface area dry condition, RST3 carbon showed better results than carbox sample.
from 75 m2/g to 593 m2/g and also increased the total volume and The biogas with a concentration of H2S around 10 ppm(v), with RST3
micropore volume. This process increased the ability to remove trace
compounds, such as hydrogen sulfide using biochar. The ability to re-
move H2S and HCl was proved according to results reported in Table 3. Table 3
In the following figure activated biochar is compared to the biochar Adsorption capacity for single compound removal.
sample for the removal of H2S (see Fig. 5). GHSV 1636 h−1 H2S (mg/g) HCl (mg/g)
The activated biochar shows comparable results with commercially
RST3 1.53 20.41
available carbons.
Carbox 1.75 12.09
Biochar 0.01 0.27
Activated biochar in Temperature 0.95 8.71

85
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

70

60

50

40
C/C0 (%)

30

20

biochar
10
biochar act

0
0 50 100 150 200 250 300
Time (min)
Fig. 5. Single compound removal – biochar versus activated biochar.

120
H2S Carbox
HCl Carbox
100 H2S RST3
HCl RST3
Breakthrough time (min)

80

60

40

20

0
0 50 100 150 200 250
Initial Concentration (ppmv)
Fig. 6. H2S and HCl removal – concentration dependence.

showed a tb around 60 min and 55 min, respectively in case of model compared to sulfur which affected the removal capacity of the sorbent
and experimental condition. Considering the same biogas conditions, material. The Wheeler-Jonas equation, adopted as predictive model
the carbox sample showed a performance (tb) around 54 min and worked well with an overestimation of 8% at lower concentrations,
50 min, respectively in case of model and experimental condition. The while at higher concentrations there was an overestimation around 1%.
biogas with a concentration of HCl around 10 ppm(v), with RST3,
showed a tb around 103 min and 91 min, respectively in case of model 3.4. The impact of moisture and temperature on the removal performance
and experimental condition. Considering the same biogas conditions,
the carbox sample showed a performance (tb) around 99 min and Fig. 7 depicts the breakthrough time trend for H2S varying the re-
91 min, respectively in case of model and experimental condition. lative humidity and the operating temperature (inlet H2S 10 ppm(v)). It
Above 50 ppm(v) of H2S the tb was less than 5 min. The breakthrough is highlighted how increasing the operating temperature the break-
time values, in case of HCl were higher compared to H2S for similar through time shows a decreasing rate. The relative humidity content
concentrations. This was due to the higher dipole moment of chlorine inside a fuel stream has a negative impact on the H2S removal

86
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

160 160
5°C 5°C
(a) 10°C
(b) 10°C
140 15°C 140 15°C
20°C 20°C
120 25°C 120 25°C
30°C 30°C
Breakthrough time (min)

Breakthrough time (min)


35°C 35°C
100 100 40°C
40°C
45°C 45°C
80 80

60 60

40 40

20 20

0 0
0 10 20 30 40 50 60 70 80 90 0 20 40 60 80 100
Relative Humidity (%) Relative Humidity (%)

Fig. 7. Breakthrough time versus relative humidity content of biogas for H2S and temperature – Norit, RST3 (a), Airdep, Carbox (b).

performance. In fact, contrarily to what reported from Sisani et al. [49] Increasing the operating temperature decreases the adsorption capa-
and from Barelli et al. [27], RST3 shows a continuous decreasing of tb city. Below 20 °C and above RH 50%, a substantial breakthrough time
trend increasing the RH content [49]. In accordance to Sisani et al. decrease is recorded. For RST3, the breakthrough time at RH 0% passes
[49], the higher potassium content in carbox favors the H2S removal from 272 min at 5 °C up to 109 min at 45 °C. Above RH 50% the
with an RH content lower than 10%, in case of RST3. Higher RH values breakthrough time is above zero compared to the removal of H2S. For
instead of favor the H2S removal causes the competition between sulfur carbox, the breakthrough time at RH 0% passes from 266 min at 5 °C up
and water in the absence of the activating agent (KOH). to 104 min at 45 °C. Here, compared to H2S removal case, above RH
For RST3, the breakthrough time at RH 0% passes from 150 min, at 50% the breakthrough time is above zero. The breakthrough time, at
5 °C up to 50 min, at 45 °C while above RH 50% and above 25 °C, the RH 80% shows values above 0 min for all the operating temperature
breakthrough time is around zero. The breakthrough time at RH 80% considered, both for RST3 and carbox sample. At a relative humidity of
shows different values to zero for temperature below 15 °C. Hence the 80% the tb passes from 174 min at 5 °C up to 2.5 min at 45 °C for RST3.
breakthrough time is zero when the operating temperature is above Considering the carbox sample, at a relative humidity of 80% the tb
20 °C. For carbox sample, the breakthrough time at RH 0% passes from passes from 171 min, at 5 °C up to 2.6 min at 45 °C. Considering carbox
135 min at 5 °C up to 41 min at 45 °C, while above RH 50% and above sample, bringing the relative humidity from 0% to 10%, there is an
25 °C, the breakthrough time is around zero. The breakthrough time at increasing trend on the breakthrough time of 6.0% at 5 °C, 6.3% at
RH 80% shows values different to zero for temperature below 15 °C. At 20 °C and 18.1% at 45 °C. When relative humidity is 10% the break-
a relative humidity of 80% the tb passes from 54.7 min at 5 °C up to through time shows a decreasing trend softer compared to RST3 case.
4.3 min at 10 °C. While, still for RST3, above 15 °C the tb is zero. Even if for RST3 sample, changing the operating temperature, the ab-
Considering the carbox sample, at a relative humidity of 80% the tb solute values of tb are higher respect to carbox sample. It results that
passes from 50.7 min at 5 °C up to 5.2 min at 10 °C, while above 15 °C RST3 shows better performance, especially at dry conditions compared
the tb is zero. to carbox sample. The carbox sample shows good performance at RH
Bringing the relative humidity from 0% to 10% there is an in- around 10%. The HCl removal case shows higher breakthrough time
creasing trend on the breakthrough time. values respect to the H2S case. This is due to the higher dipole moment
Fig. 8 depicts the breakthrough time trend for HCl (9 ppm(v)) of HCl (1.08 D) compared to H2S (0.95 D) [50,51]. The dipole moment
varying the relative humidity and the operating temperature. is a measure of the degree of polarity. The phenomenon explained as

300 300
5°C 5°C
(a) 10°C
(b) 10°C
15°C 15°C
250 20°C
250 20°C
25°C 25°C
30°C
Breakthrough time (min)

30°C
Breakthrough time (min)

200 35°C 200 35°C


40°C 40°C
45°C 45°C
150 150

100 100

50 50

0 0
0 20 40 60 80 100 0 20 40 60 80 100
Relative Humidity (%) Relative Humidity (%)

Fig. 8. Breakthrough time versus relative humidity content of biogas and temperature for HCl – Norit, RST3 (a), Airdep, Carbox (b).

87
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

follows: the higher the polarity of compounds of interest is the stronger removal performance for H2S are similar in single and multiple com-
electron-donating ability become [52]. pounds configuration. When the pass-through ratio is around 10%, the
The combination of the operating temperature and relative hu- breakthrough time is 38 min for both configurations (single and mul-
midity on the breakthrough time is reported in Figs. 8 and 9. This aspect tiple cases). At this point there is the inversion of the curve profile, and
causes the decreasing of the removal performance. Only some salts, from here forward the removal performance, for RST3 improves in the
adopted for the impregnation process help the removal performance for single compound configuration. The simultaneous presence decreases
sulfur compounds. This is the case of KOH and KI, which are able to the removal performance due to the competition of HCl for the removal
facilitate the removal of sulfur compounds, as reported in literature of H2S. Biochar results are not reported because the pass-through ratio
[27,49]. achieved the saturation value in 10 min.
Breakthrough time values modeled with Eq. (3) are in agreement The 1%C/C0 ranges from 44 min and 47 min for HCl with RST3,
with experimental results. This data are reported varying the operating respectively in single and multiple compounds configuration
temperature from 5 °C to 45 °C. Breakthrough time for HCl is in general (H2S + HCl). The removal performance for HCl are similar in single and
higher respect to H2S case, both for RST3 and carbox. The results multiple compound configuration. When the pass-through ratio is
confirm a marked tb decline increasing the operating temperature. The around 11% the breakthrough time is 49 min for both configurations.
maximum breakthrough time value for RST3 at 5 °C is 147 min, for H2S At this point, there is the inversion of the curve profile, and from here
(RH = 0%) and 278 min for HCl (RH = 0%). The minimum break- forward the removal performance, for RST3 improves in the single
through time value at 45 °C is 52.6 min for H2S (RH = 0%) and 110 min compound configuration. The simultaneous presence decreases the re-
for HCl (RH = 0%). Considering the carbox sample, the maximum tb is moval performance due to the competition of H2S for the removal of
264 min and 136 min, respectively for HCl and H2S at 5 °C and HCl.
RH = 0%. The minimum tb value is 105 min and 45 min, respectively Results reported in Fig. 11 consider the effect of humidified gas
for HCl and H2S at 45 °C at the same RH condition. (RH = 30%) with the simultaneous concentration of H2S and HCl. Here,
By increasing the operating temperature, the decreasing perfor- the addition of water for all the experimental tests leads to a decrease of
mance of the activated carbon is observed (Fig. 9). As the temperature removal performance. Considering the breakthrough at 1% of C/C0, the
rises, the vapor pressure of the adsorbate increases [22,38]. Hence, the time decreases of 77% for RST3 with humid biogas. The reduction of
energy level of the H2S and HCl molecules raises to overcome the Van removal performance adding water to the gas stream is due to the in-
der Waals attraction and migrate back to the bulk gas phase reducing terference of water in the pores of activated carbons.
the overall breakthrough time [53]. The effect is that the breakthrough Biochar results are not reported because the pass-through ratio
time decreases with increasing adsorbent bed temperature. Thus, an achieved the saturation value in 30 min.
increase in the operating temperature caused lower values of the The adsorption capacity evaluated by Eq. (1) is reported in the
maximum adsorption capacity. following table, considering the simultaneous concentration of H2S, HCl
and water:
3.5. Removal of multiple compounds and influence of humidified gas From single to multiple compounds the adsorption capacity for
RST3 decreases. The influence of water is very important as it can be
Generally sorbent materials face the contemporary presence of more seen from the adsorption capacity, table 4.
than one pollutant (co-vapors removal condition). Fig. 10 shows the
breakthrough time trend for the sorbent materials tested in both single 3.6. Porous particle diffusion model validation – single compound removal
and multiple compounds configuration (H2S and HCl). The 1%C/C0
ranges from 30 min and 35 min for H2S with RST3, respectively in Model validation was accomplished by comparing the experimental
single and multiple compounds configuration (H2S + HCl). The results with data obtained from the model described before. This

250

H2S Carbox
HCl Carbox
200 H2S RST3
HCl RST3
Breakthrough time (min)

150

100

50

0
275 280 285 290 295 300 305 310 315 320
Temperature (K)
Fig. 9. H2S and HCl removal – RST3 and Carbox – Temperature effect.

88
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

100

90

80

70

60
C/C0 (%)

50

40

30
RST3 - H2S RST3 - H2Sco

20
RST3 - HCl RST3 - HClco
10

0
0 50 100 150 200 250
Time (min)
Fig. 10. H2S and HCl removal – (co)-vapors effect.

100

90

80

70

60
C/C0 (%)

50

40

30
RST3 - H2Sco RST3 - H2Sco w

20

10 RST3 - Hclco RST3 - HClco w

0
0 50 100 150 200 250
Time (min)
Fig. 11. H2S and HCl removal – co-vapors effect – water influence.

Table 4 validation is necessary to understand the real ability to predict packed


Adsorption capacity for multiple compound removal and water influence. bed breakthrough. The model must be precise especially at an ultra-low
GHSV 1636 h−1 H2Sco (mg/ HClco (mg/ H2Sco-w HClco-w (mg/ concentration to be suitable for SOFC feeding requirements. In this
g) g) (mg/g) g) case, the focus is just on the first part of the breakthrough curve:
0–0.4% on the y-axis (C/C0). It is interesting to investigate the com-
RST3 1.48 11.19 0.42 0.06 petitive adsorption of H2S and HCl [6,54,55]. H2S single compound
Biochar 0.01 0.54 0.005 0.46
Activated biochar 1.09 6.4 0.47 0.08
removal appears closer to the experimental case.
Fig. 12 describe the breakthrough time for two different activated
carbons to remove 55 ppm(v) of H2S.
Porous particle diffusion model fits the experimental data very well,
espcially around the breakthrough point (0–0.4% C/C0). In fact, for

89
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

Fig. 12. Breakthrough time – H2S 55 ppm(v) - fitting evaluation for Norit RST3 (orange) and Airdep Carbox (blue). (For interpretation of the references to colour in
this figure legend, the reader is referred to the web version of this article.)

RST3 the value of mean squared error (MSE) is equal to 0.029%. The 4. Conclusions
breakthrough point is predicted by the model around 47 min and by the
experimental data at 48 min. The removal of selected trace compounds considering the covapor
The breakthrough point is predicted after 60 min, for carbox carbon condition, the operating temperature and also humid biogas conditions
while experiments underline a breakthrough after around 80 min of were studied on commercially available sorbents and waste-derived
operation. Despite this discrepancy between these values, the MSE material (biochar). The best removal performance, in single compound
value is fixed around 0.071%. removal condition (H2S or HCl) was achieved by RST3. Biochar, com-
Fig. 13 depicts the simultaneous removal of two trace compounds: pared to the other sorbent materials showed the lowest removal per-
H2S and HCl detected in biogas from organic waste. Results achieved formance, both for H2S and HCl. Activated biochar with carbon dioxide
from the model and compared to experimental points shows a C/C0 at 750 °C, showed interesting results comparable with commercially
trend accurate for H2S (35.1 min against 35.22 min), while under- available sorbents. This was due the increasing of specific surface area
estimate the HCl removal performance (breakthrough point expected and porous volume.
10 min later). Here the percentage error for the breakthrough point Operating temperature affects the removal performance for all the
estimation is 0.34% for H2S and 23% for HCl, Fig. 13. sorbents considered. The vapor pressure of the adsorbate increases with
In summary, low values of MSE validate the model in the first part the temperature. This favors the energy level rising for the trace com-
of the breakthrough curve, even for competitive adsorption. The prin- pounds to overcome the van der walls attraction bonds. This causes the
cipal result achieved from the porous particle diffusion model is the migration to the bulk gas phase reducing the breakthrough time. Thus,
excellent capacity to estimate the breakthrough time. This estimation is an increase in the operating temperature caused lower values of the
the most important point for SOFC feeding with biogas produced from maximum capacity of the adsorbent. The humidity contained in the
organic waste. biogas affects negatively the removal performance. This is connected to
the interference of water with trace compounds on the pores of

0.25

H2S
H2S model
0.20 HCl
HCl model

0.15
C/C0

0.10

0.05

0.00
0 10 20 30 40 50 60
Time (min)
Fig. 13. Breakthrough time – H2S 64 ppm(v) and HCl 307 ppm(v) - fitting evaluation for Norit RST3.

90
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

activated carbons. It was found that the commercially available sor- control, fate and removal of siloxanes on the energetic valorization of sewage
bents and waste-derived material tested are negatively affected by biogas – a review, Renew. Sustain. Energy Rev. 52 (2015) 366–381, https://doi.
org/10.1016/j.rser.2015.07.106.
water inside biogas. [18] D. Papurello, V. Chiodo, S. Maisano, A. Lanzini, M. Santarelli, Catalytic stability of a
Experimental data were employed in a porous particle diffusion Ni-Catalyst towards biogas reforming in the presence of deactivating trace com-
model to estimate the breakthrough time. Low mean squared errors pounds, Renew. Energy 127 (2018) 481–494, https://doi.org/10.1016/j.renene.
2018.05.006.
were achieved in the first part of the breakthrough curve, even for [19] T.J. Bandosz, On the adsorption/oxidation of hydrogen sulfide on activated carbons
competitive adsorption for SOFC feeding requirements. at ambient temperatures, J. Colloid Interface Sci. 246 (2002) 1–20, https://doi.org/
Future works will involve experiments on various methods of acti- 10.1006/jcis.2001.7952.
[20] N. Abatzoglou, S. Boivin, A review of biogas purification processes, Biofuels
vation for biochar samples and on the model validation for the con- Bioprod. Biorefining 6 (2009) 42–71, https://doi.org/10.1002/bbb.
temporary pollutants removal. [21] G. Coppola, D. Papurello, Biogas cleaning: activated carbon regeneration for H2S
removal, Clean. Technol. Mdpi (2018) 25–28, https://doi.org/10.3390/
cleantechnol1010004.
Acknowledgements
[22] M.J. Lashaki, M. Fayaz, H. Wang, Z. Hashisho, J.H. Philips, J.E. Anderson, et al.,
Effect of adsorption and regeneration temperature on irreversible adsorption of
This work is a part of the work for the demosofc project which is organic vapors on beaded activated carbon, Environ. Sci. Technol. 46 (2012)
carried out by Politecnico di Torino and other European partners (FCH- 4083–4090, https://doi.org/10.1021/es3000195.
[23] D.D. Papadias, S. Ahmed, R. Kumar, Fuel quality issues with biogas energy – an
JU). In addition, this research is part of the BWS project (Biowaste for economic analysis for a stationary fuel cell system, Energy 44 (2012) 257–277,
SOFCs), project funded by the Fondazione Cassa di Risparmio di Trento https://doi.org/10.1016/j.energy.2012.06.031.
e Rovereto. A special thanks to Ph.D. Fiorilli Sonia for her important [24] A. Kohl, R. Nielsen, Gas purification, fifth ed., Gulf Professional Publishing,
Houston, Texas, 1997, p. 1997.
and fundamental work and a special thanks to our technician Bressan [25] S.P. Hernández, F. Scarpa, D. Fino, R. Conti, Biogas purification for MCFC appli-
Maurizio. cation, Int. J. Hydrogen Energy 36 (2011) 8112–8118, https://doi.org/10.1016/j.
ijhydene.2011.01.055.
[26] R.T. Yang, Gas separation by adsorption processes, Butterworths (1988) 2, https://
References doi.org/10.1016/0950-4214(88)80042-2.
[27] L. Barelli, G. Bidini, N. De Arespacochaga, P. Laura, E. Sisani, Biogas use in high
[1] A. Hagen, J.F.B. Rasmussen, K. Thydén, Durability of solid oxide fuel cells using temperature fuel cells: enhancement of KOH-KI activated carbon performance to-
sulfur containing fuels, J. Power. Sources 196 (2011) 7271–7276, https://doi.org/ ward H 2 S removal 2 (2017) 10.1016/j.ijhydene.2017.02.021.
10.1016/j.jpowsour.2011.02.053. [28] D. Papurello, L. Tomasi, S. Silvestri, M. Santarelli, Evaluation of the Wheeler-Jonas
[2] J.F.B. Rasmussen, A. Hagen, The effect of H2S on the performance of Ni-YSZ anodes parameters for biogas trace compounds removal with activated carbons, Fuel
in solid oxide fuel cells, J. Power Sources 191 (2009) 534–541, https://doi.org/10. Process. Technol. 152 (2016), https://doi.org/10.1016/j.fuproc.2016.06.006.
1016/j.jpowsour.2009.02.001. [29] D. Papurello, L. Tomasi, S. Silvestri, I. Belcari, M. Santarelli, F. Smeacetto, et al.,
[3] A. Lanzini, D. Ferrero, D. Papurello, M. Santarelli, Reporting degradation from Biogas trace compound removal with ashes using proton transfer reaction time-of-
different fuel contaminants in Ni-anode SOFCs, Fuel Cells 17 (2017), https://doi. flight mass spectrometry as innovative detection tool, Fuel Process. Technol. 145
org/10.1002/fuce.201600184. (2016), https://doi.org/10.1016/j.fuproc.2016.01.028.
[4] D. Papurello, A. Lanzini, S. Fiorilli, F. Smeacetto, R. Singh, M. Santarelli, Sulfur [30] W. Yuan, T.J. Bandosz, Removal of hydrogen sulfide from biogas on sludge-derived
poisoning in Ni-anode solid oxide fuel cells (SOFCs): deactivation in single cells and adsorbents, Fuel 86 (2007) 2736–2746, https://doi.org/10.1016/j.fuel.2007.03.
a stack, Chem. Eng. J. (2016;283.), https://doi.org/10.1016/j.cej.2015.08.091. 012.
[5] D. Papurello, S. Silvestri, L. Tomasi, I. Belcari, F. Biasioli, M. Santarelli, Biowaste for [31] M. Hervy, D. Pham Minh, C. Gérente, E. Weiss-Hortala, A. Nzihou, A. Villot, et al.,
SOFCs, Energy Procedia 101 (2016), https://doi.org/10.1016/j.egypro.2016.11. H2S removal from syngas using wastes pyrolysis chars, Chem. Eng. J. 334 (2018)
054. 2179–2189, https://doi.org/10.1016/j.cej.2017.11.162.
[6] S. Rasi, J. Läntelä, J. Rintala, Trace compounds affecting biogas energy utilisation - [32] C. Gopu, L. Gao, M. Volpe, L. Fiori, J.L. Goldfarb, Valorizing municipal solid waste:
a review, Energy Convers. Manag. 52 (2011) 3369–3375, https://doi.org/10.1016/ waste to energy and activated carbons for water treatment via pyrolysis, J. Anal.
j.enconman.2011.07.005. Appl. Pyrolysis 133 (2018) 48–58, https://doi.org/10.1016/j.jaap.2018.05.002.
[7] D. Papurello, A. Lanzini, L. Tognana, S. Silvestri, M. Santarelli, Waste to energy: [33] D. Papurello, E. Schuhfried, A. Lanzini, A. Romano, L. Cappellin, T.D. Märk, et al.,
exploitation of biogas from organic waste in a 500 W<inf>el</inf&gt solid Proton transfer reaction-mass spectrometry as a rapid inline tool for filter efficiency
oxide fuel cell (SOFC) stack, Energy 85 (2015), https://doi.org/10.1016/j.energy. of activated charcoal in support of the development of Solid Oxide Fuel Cells fueled
2015.03.093. with biogas, Fuel Process. Technol. 130 (2015), https://doi.org/10.1016/j.fuproc.
[8] D. Papurello, A. Lanzini, P. Leone, M. Santarelli, The effect of heavy tars (toluene 2014.09.042.
and naphthalene) on the electrochemical performance of an anode-supported SOFC [34] D. Papurello, A. Lanzini, E. Schufried, M. Santarelli, S. Silvestri, Proton Transfer
running on bio-syngas, Renew. Energy 99 (2016), https://doi.org/10.1016/j. Reaction-Mass Spectrometry (PTR-MS) as a rapid online tool for biogas VOCs
renene.2016.07.029. monitoring in support of the development of Solid Oxide Fuel Cells (SOFCs), in: 6th
[9] D. Papurello, A. Lanzini, D. Drago, P. Leone, M. Santarelli, Limiting factors for Int PTR-MS Conf, vol. 130, 2013, pp. 144–150. https://doi.org/10.1016/j.fuproc.
planar solid oxide fuel cells under different trace compound concentrations, Energy 2014.09.042.
95 (2016), https://doi.org/10.1016/j.energy.2015.11.070. [35] D. Papurello, L. Tognana, A. Lanzini, F. Smeacetto, M. Santarelli, I. Belcari, et al.,
[10] H. Madi, A. Lanzini, D. Papurello, S. Diethelm, C. Ludwig, M. Santarelli, et al., Solid Proton transfer reaction mass spectrometry technique for the monitoring of volatile
oxide fuel cell anode degradation by the effect of hydrogen chloride in stack and sulfur compounds in a fuel cell quality clean-up system, Fuel Process. Technol. 130
single cell environments, J. Power Sources 326 (2016), https://doi.org/10.1016/j. (2015), https://doi.org/10.1016/j.fuproc.2014.09.041.
jpowsour.2016.07.003. [36] J.B. Rosen, Kinetics of a fixed bed system for solid diffusion into spherical particles,
[11] M. Santarelli, L. Briesemeister, M. Gandiglio, S. Herrmann, P. Kuczynski, J. Chem. Phys. 20 (1952) 387, https://doi.org/10.1063/1.1700431.
J. Kupecki, et al., Carbon recovery and re-utilization (CRR) from the exhaust of a [37] A. Rasmuson, Exact solution of a model for diffusion and transient adsorption in
solid oxide fuel cell (SOFC): analysis through a proof-of-concept, J. CO2 Util 18 particles and longitudinal dispersion in packed beds, AIChE J. 27 (1981)
(2017), https://doi.org/10.1016/j.jcou.2017.01.014. 1032–1035, https://doi.org/10.1002/aic.690270625.
[12] D. Papurello, C. Iafrate, A. Lanzini, M. Santarelli, Trace compounds impact on SOFC [38] Y. Xiao, S. Wang, D. Wu, Q. Yuan, Experimental and simulation study of hydrogen
performance: experimental and modelling approach, Appl. Energy 208 (2017), sulfide adsorption on impregnated activated carbon under anaerobic conditions, J.
https://doi.org/10.1016/j.apenergy.2017.09.090. Hazard. Mater. 153 (2008) 1193–1200, https://doi.org/10.1016/j.jhazmat.2007.
[13] D. Papurello, A. Lanzini, P. Leone, M. Santarelli, S. Silvestri, Biogas from the organic 09.081.
fraction of municipal solid waste: dealing with contaminants for a solid oxide fuel [39] A. Khazraei Vizhemehr, F. Haghighat, C.-S. Lee, Predicting gas-phase air-cleaning
cell energy generator, Waste Manag. 34 (2014), https://doi.org/10.1016/j.wasman. system efficiency at low concentration using high concentration results: develop-
2014.06.017. ment of a framework, Build. Environ. 68 (2013) 12–21, https://doi.org/10.1016/j.
[14] D. Papurello, A. Lanzini, SOFC single cells fed by biogas: experimental tests with buildenv.2013.05.023.
trace contaminants, Waste Manag (2017), https://doi.org/10.1016/j.wasman. [40] J.F. Scamehorn, Removal of vinyl chloride from gaseous streams by adsorption on
2017.11.030. activated carbon, Ind. Eng. Chem. Process. Des. Dev. 18 (1979) 210–217, https://
[15] H. Madi, A. Lanzini, S. Diethelm, D. Papurello, J. Van Herle, M. Lualdi, et al., Solid doi.org/10.1021/i260070a004.
oxide fuel cell anode degradation by the effect of siloxanes, J. Power Sources 279 [41] G. Guiochon, S. Golshan Shirazi, A.M. Katti, Fundamentals of Preparative and
(2015), https://doi.org/10.1016/j.jpowsour.2015.01.053. Nonlinear Chromatography, Academic Press, 1994.
[16] K. Sasaki, K. Haga, T. Yoshizumi, D. Minematsu, E. Yuki, R. Liu, et al., Chemical [42] P. Lodewyckx, L. Verhoeven, Using the Wheeler-Jonas equation to describe ad-
durability of Solid Oxide Fuel Cells: influence of impurities on long-term perfor- sorption of inorganic molecules: ammonia, Carbon N Y 41 (2003) 1215–1219,
mance, J. Power Sources 196 (2011) 9130–9140, https://doi.org/10.1016/j. https://doi.org/10.1016/S0008-6223(03)00052-6.
jpowsour.2010.09.122. [43] P. Lodewyckx, E.F. Vansant, Estimating the overall mass transfer coefficient k(v) of
[17] N. De Arespacochaga, C. Valderrama, J. Raich-Montiu, M. Crest, S. Mehta, the Wheeler-Jonas equation: a new and simple model, Am. Ind. Hyg. Assoc. J. 61
J.L. Cortina, Understanding the effects of the origin, occurrence, monitoring, (2000) 501–505, https://doi.org/10.1080/15298660008984561.

91
D. Papurello et al. Separation and Purification Technology 210 (2019) 80–92

[44] P. Lodewyckx, L. Verhoeven, Using the modified Wheeler-Jonas equation to de- constants – dipole moments, n.d.
scribe the adsorption of inorganic molecules: Chlorine, Carbon N Y 41 (2003) [51] M. Blaber, Dipole Moments, n.d.
1215–1219, https://doi.org/10.1016/S0008-6223(03)00052-6. [52] X. Yang, J. Li, T. Wen, X. Ren, Y. Huang, X. Wang, Colloids and surfaces a :
[45] J. Wu, O. Claesson, I. Fangmark, L.G. Hammarstrom, A systematic investigation of Physicochemical and engineering aspects adsorption of naphthalene and its deri-
the overall rate coefficient in the Wheeler-Jonas equation for adsorption on dry vatives on magnetic graphene composites and the mechanism investigation,
activated carbons, Carbon N Y 43 (2005) 481–490, https://doi.org/10.1016/j. Colloids Surf. A Physicochem. Eng. Asp. 422 (2013) 118–125, https://doi.org/10.
carbon.2004.09.024. 1016/j.colsurfa.2012.11.063.
[46] L.A. Jonas,J.A. Rehrmann , Predictive equations in gas adsorption kinetics, 1973, [53] H.S. Choo, L.C. Lau, A.R. Mohamed, K.T. Lee, Hydrogen sulfide adsorption by al-
pp. 59–64. https://doi.org/10.1016/0008-6223(73)90008-0. kaline impregnated coconut shell activated carbon, J. Eng. Sci. Technol. 8 (2013)
[47] D. Papurello, R. Borchiellini, P. Bareschino, V. Chiodo, S. Freni, A. Lanzini, et al., 741–753.
Performance of a Solid Oxide Fuel Cell short-stack with biogas feeding, Appl. [54] S. Rasi, A. Veijanen, J. Rintala, Trace compounds of biogas from different biogas
Energy 125 (2014), https://doi.org/10.1016/j.apenergy.2014.03.040. production plants, Energy 32 (2007) 1375–1380, https://doi.org/10.1016/j.
[48] W. Feng, S. Kwon, Adsorption of hydrogen sulfide onto activated carbon fibers : energy.2006.10.018.
effect of pore structure and surface, Chemistry 39 (2005) 9744–9749. [55] D. Papurello, C. Soukoulis, E. Schuhfried, L. Cappellin, F. Gasperi, S. Silvestri, et al.,
[49] E. Sisani, G. Cinti, G. Discepoli, D. Penchini, U. Desideri, F. Marmottini, Adsorptive Monitoring of volatile compound emissions during dry anaerobic digestion of the
removal of H2S in biogas conditions for high temperature fuel cell systems, Int. J. Organic Fraction of Municipal Solid Waste by Proton Transfer Reaction Time-of-
Hydrogen Energy 39 (2014) 21753–21766, https://doi.org/10.1016/j.ijhydene. Flight Mass Spectrometry, Bioresour. Technol. 126 (2012), https://doi.org/10.
2014.07.173. 1016/j.biortech.2012.09.033.
[50] M.L. McGlashan, National Physical Laboratory, Tables of Physical and chemical

92

Potrebbero piacerti anche