You are on page 1of 10

ARTICLE IN PRESS

Biomass and Bioenergy 29 (2005) 142151 www.elsevier.com/locate/biombioe

A H2S reactive adsorption process for the purication of biogas prior to its use as a bioenergy vector
L.V.-A. Truong, N. Abatzoglou
de Sherbrooke, 2500 boul. Universite , Faculty of Engineering, Chemical Engineering Department, Universite Sherbrooke, QC, Canada J1 K 2R1 Received 15 December 2003; accepted 3 March 2005 Available online 24 May 2005

Abstract This work studies, at lab scale, a reactive adsorption technology for the removal of the H2S from biogas produced in landlls or anaerobic digesters. The main phenomenon is an irreversible chemical reaction between the solid and gas phase. The study produces data on the efciency of the process as function of a number of variables including: the nature and properties of the adsorbent; the biogas ow rate and the contact time of the gas with the adsorbent; the geometry of the adsorption columns and the linear velocity of the ow; the concentration of the contaminant (H2S) and humidity in the biogas. The work reported in this paper focuses on a promising adsorbent available commercially and includes the breakthrough curves and the compilation of a phenomenological model for the process. These rst results show that the rates of the external diffusion, internal diffusion and surface reaction steps are relatively close and that the limiting step of the process changes with experimental conditions. Globally, the phenomenological model predicts that the rate of the process is near rst order with respect to the H2S concentration and zero order with respect to the solid reactant (adsorbent). The reported data constitute the basis for the scale-up of the unit at a commercial level. r 2005 Elsevier Ltd. All rights reserved.
Keywords: Biogas; Bioenergy; Adsorption; Hydrogen sulphide; Iron oxide

1. Introduction Anaerobic fermentation (AF) of organic waste produces a biogas with high concentration of methane (CH4). The biogas formed in AF plants consists of 5580vol% CH4, 2045vol%CO2,
Corresponding author. Tel.: 819 821 7904; fax: 819 821 7955.

E-mail address: Nicolas.Abatzoglou@USherbrooke.ca (N. Abatzoglou).

01.0vol%H2S, 00.05vol% NH3 and it is saturated with water [1]. Sulphate reducing bacteria grow in the digester and use acetic or propionic acid to produce H2S. This step occurs simultaneously with the methane production [2]. The generation of H2S poses serious problems of odour, toxicity for human and animal health, and corrosion. Additionally, hydrogen sulphide is known to be active in the presence of ferrous alloys and is a hydrogen embrittlement source (i.e. steel).

0961-9534/$ - see front matter r 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.biombioe.2005.03.001

ARTICLE IN PRESS
L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151 143

Moreover, whenever used for the production of energy, the biogas has to be conditioned because the combustion of hydrogen sulphide produces SOx, known for their detrimental role in the atmosphere and human health. Biogas contaminated with hydrogen sulphide can be puried by various methods [1,3]. Many small-scale AF plants have tried to use biogas as a bioenergy vector but the majority of these projects ended because of the corrosion and the high maintenance cost of the process [4]. The choice of the H2S removal process is, thus, site specic. In the case of this study, the process is designed for a farm-scale unit. As the process has to be simple and the product must be easy to handle and environmentally safe, the adsorption using a H2S scavenger was chosen.

particle size of the active components are shown in the SEM picture of Fig. 5. Apart from its specications, the manufacturer sells the product with: (a) a MSDS; (b) copies of ofcial documents proving its classication by EPA (USA) and the Alberta Environmental Protection Authority (Canada) as a non-hazardous material, in both its unreacted and dry, ready for disposal form; (c) a list of disposal options; (d) environmental data summary including the disposal approvals in various states in USA, Canada and Offshore. These documents are available upon request; a short summary of the main points follows:

2. Materials and methods 2.1. Materials A commercially available adsorbent was chosen for this study. The adsorbent is the Sulfatreat 410HPs (www.sulfatreat.com) .The active ingredients of the adsorbent is a combination of iron oxides (Fe2O3, Fe3O4) and an activator oxide attached to a calcined montmorillonite carrier matrix; the latter is thought to be enhancing catalytically the reactive adsorption phenomenon. Based on data available from the manufacturer, it is known that the amount of activator is 0.1255%w/w of the adsorbent. The activator is constituted on one or more oxides of a group of metals consisting of platinum, gold, silver, copper, cadmium, nickel, palladium, lead, mercury, tin and cobalt [5]. The active ingredients are supported on a non-porous silica (SiO2) matrix containing small amounts of alumina (Al2O3); an aluminosilicate coming from montmorillonite. Thus, it can be said that the adsorbent is a two-dimension (2D) formulation in which the active ingredients are at the surface of a coarse size support matrix; the silica particles diameter varies between 4.06.5 mm. These data are corroborated by SEM analyses which are provided within the Figs. 1, 2(a), 2(b) and 3. Fig. 4 depicts the sulphur distribution at the surface of the adsorbent after use. The morphology and

The material is non-toxic and non-hazardous. Both its unreacted and spent forms can be disposed of safely by landlling or even offshore spilling, if the gas treated does not contain dangerous contaminants other than sulphurous compounds. Disposal options, other than landll, include: recycle into sludge treatment and disposal processes; use as a road material; in agricultural fertilizers for product enhancement; in ceramic bricks manufacturing. Although there are no analytical data on the lifecycle of this material, the information available by the manufacturer indicates that even the use of the spent material is safe.

Fig. 1. Elements mapping by SEM on adsorbent (active+ support) particles; silicium (Si) in spots.

ARTICLE IN PRESS
144 L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151

Fig. 3. Elements mapping by SEM on adsorbent (active+ support) particles; aluminium (Al) in spots.

Fig. 2. (a, b) Elements mapping by SEM on adsorbent (active+support) particles; iron (Fe) in spots.

Fig. 4. Elements mapping by SEM on adsorbent (active+ support) particles; adsorbed sulphur in spots.

The manufacturer also provided the information that the active material has a relatively high internal porosity ( 0:75). This lm of active material seems to be physically attached to the inert matrix because it produces dust easily upon handling. The powder has been also analyzed by SEM; it is found that (a) the size of the powder particles varies between 1 and 10 mm and (b) that the powder is mainly composed of iron oxides with small quantities of Cu and Mn. Qualitative evidence for the presence of these elements has been obtained through SEM mapping.

The apparent density of the dry adsorbent is 1000 kg/m3. The experimental data have proven that 1 g of adsorbent can adsorb up to 0.11 g of H2S. The saturation was dened experimentally as the point at which the concentrations of the inlet and outlet gas in H2S were identical C OUT H2 S C IN H2 S . The obtained breakthrough curve was integrated to nd the mass of H2S adsorbed by the media. The specic surface is 5.4 m2/g. The adsorbent is environmentally safe in both nonreacted and H2S saturated forms and no odour

ARTICLE IN PRESS
L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151 Table 1 List of experiments H2S conc. (ppmv) Contact time (s) Column diameter (cm) 6.35 3.81 6.35 3.81 6.35 3.81 6.35 6.35 3.81 6.35 3.81 3.81 Water saturation 145

Fig. 5. Particle size of active component.

10 000 10 000 10 000 10 000 10 000 10 000 10 000 3000 3000 3000 3000 3000

30 30 30 30 60 60 60 30 30 60 60 30

Y Y N N Y Y Y Y Y Y Y N

emanating from it has been observed before or after its use. The hydrogen sulphide (H2S) adsorption from a simulated biogas (H2S: 3000 ppmv or 10 000 ppmv, CO2: 29% and CH4: balance) has been studied under dynamic conditions as function of contact time, column geometry, linear velocity, initial H2S concentration and water content in gas. Two columns of 3.81 and 6.35 cm diameter, respectively, have been used at constant gas ow rate of 20 L gas/h, operating at atmospheric pressure and ambient temperature. The contact time between the gas and media has been dened by varying the height of the adsorbent bed. Two contact times were tested (30 and 60 s) and reported in this paper. Two different H2S content gas mixtures, both dry and water saturated have been tested. The saturation, wherever required, was reached by bubbling the initial dry gas through a H2S saturated aqueous solution. The solubility of H2S in water at 20 1C is 3.98 g/L of water. The saturation is conrmed by analyzing the inlet and outlet concentrations of the H2S stream. The rst target of the experimentation was to trace the breakthrough curves and to determine the limiting step of the phenomenon (diffusion or reactive adsorption at the surface of the adsorbent). A list of the experiments is presented in Table 1. The experimental protocol is the following: a chosen gas ow rate of 20 L/h is fed through the xed bed of the adsorbent media. The composition

of the inlet gas (H2S and water content) is also set (3000 and 10 000 ppmv). The linear velocity of the gas ow is dened by keeping the same ow rate and changing the diameter of the column: 0.18 cm/s for the 6.35 cm diameter column and 0.49 cm/s for the 3.81 cm diameter column. Finally, the water content depends on whether bubbling through the H2S saturated aqueous solution is used. The H2S concentration at the inlet and outlet of the xed bed column is measured using precision detector tubes (colorimetric method) from Matheson-Kitagawa. This colorimetric method was chosen because there are no possible interferences in this system of simulated biogas. The sampling and analysis were carried out at regular time intervals. The spent gas was purged to the atmosphere after removal of the residual H2S by means of a NaOH solution bubbler. The spent media were disposed of after use by the Universitys hazardous waste management service. Fig. 6 presents a simplied schematic of the experimental set-up. 2.2. Adsorption This process is an irreversible reactive adsorption through a xed bed. Since no data on isotherm, equilibrium or chemical composition were given, theoretical calculations were used to predict the controlling step of the global phenomenon. Then, an experimental study was performed

ARTICLE IN PRESS
146 L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151
Apparatus and procedure

3-WAY VALVE FOR SAMPLING

FLOWMETER

FIXED BED ADSORPTION COLUMN

SATURATED SOLUTION H2S

GAS CYLINDER (CH4,CO2,H2S)

3-WAY VALVE FOR SAMPLING TO PURGE

NaOH SOLUTION

Fig. 6. Simplied schematic of the experimental set-up.

to validate or invalidate the predictions. Since the adsorption is irreversible, the three principal steps are: the external diffusion, the internal diffusion and the adsorption on the active site. 2.3. External diffusion calculations The mass transfer coefcient for the external diffusion is calculated from the following equation: kD Sh DH2 S_gas , dP
1=2

Table 2 Results of kD and yD Linear velocity of gas (cm=s External mass transfer (diffusion) coefcient (cm=s kD 0:98 kD 1:25 Time for external mass transfer (diffusion) (s) yD 0:114 yD 0:089

u6.35 cm: 0.18 u3.81 cm: 0.49

(1)

porosity of the product. The exchange surface of a particle of the adsorbent was calculated using next equation and the result is aP 9 cm1 ; where p d 2  6 aP 1 P 3 . dP 6 pd P (3)

where the Sherwood number is Sh 2:0 1:8 Re Sc


1=3

(2)

The Reynolds number was calculated using the linear velocity of gas and the average diameter of the media particles. Diffusivity and viscosity are calculated at P 1 atm and T 293 K for gas mixtures using the equations proposed in [6]. The kD, Sh, Re and Sc were calculated using these parameters DH2 S-gas 0:1413l cm2 =s, the diffusivity of H2S in the uid, mgas 1:267 104 g=cm s, the viscosity of the gas, rgas 0:001 g=cm3 , the density of the uid, d P 0:5 cm, the average diameter of media particle, P 0:75, the internal

The value kD represents the speed of diffusion and yD is the time required to perform this diffusion. y 1 . k D ap (4)

A comparison between the linear velocity u and the constant kD shows that in both cases, presented in Table 2, the speed of the external mass transfer (kD) is between 2.5 and 5.5 times the velocity of the gas front through the column

ARTICLE IN PRESS
L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151 147

(u 0:1820:49 cm=s). This leads to the assumption that the external diffusion step is faster than the rate of the gas crossing the column; thus, theoretically, the external diffusion might not be the controlling step. 2.4. Internal diffusion The internal diffusion is the transfer of the adsorbate, the H2S, inside the adsorbent particle. Otherwise, it is the migration of the H2S molecule from the external surface of the adsorbent into the internal pores of the active ingredients [7]. To calculate the effective diffusivity, it is assumed that the mean free path of the gas molecules is smaller than the diameter of the pores (d) of the adsorbent so molecular gas diffusion will occur. RT l p 2 , 2 pd N A P (5)

De is equal to 2.649 106 m2/s, when s 1 and De 1:325 106 m2 =s, when s 0:5. To determine whether the reaction is controlled by this diffusion step or the chemical reaction the modulus of Thiele is used. In this case, where the rs value is not known, the modied Thiele modulus will be considered j0s Zs j2 s r L2 . De C s (7)

where l is the mean free path, T the temperature, P the pressure, d the molecular diameter of H2S, NA the Avogadro number, R the gas constant and p the known constant equal to 3.14159. , temperaUsing a molecular diameter of 10 A ture of 293 K and pressure of 1 atm, it is calculated that l 9 nm. This assumption is valid for d values typically at least 5 times higher than the mean free path (l), that is 50 nm. If this is the case, the diffusivity D is equal to the molar diffusivity DH2 S_gas 1:413 105 m2 =s. Otherwise the diffusivity is rather of the Knudsen type. Since there is no clear evidence from the SEM pictures (no visible pores at the 0.5 mm level in Figs. 11 and 12) the effective diffusivity is calculated by the following equation: De P sD , tP (6)

The values of r have been calculated by using the average rates measured experimentally. Since these values are lower than 1 and the j0s j2 s o1, the reaction must be mainly controlled by the chemical kinetics. But if the constriction factor is lower than 1 (i.e. s 0:5), the values are changing just as shown in Tables 3 and 4. Since j0s j2 s 1 the chemical kinetics and the internal diffusion rates are of the same order and they may both have an impact on the global reaction. 2.5. Kinetic study and modelling Fixed bed column reactors are most frequently modelled as plug ow reactors. However, we were unable to sample the reactor along its length, and used the semi-batch stirred reactor as the phenomenological model with the formulation that the reaction of H2S at concentration CA with the active sites at concentration CB is irreversible leading to A+B-C. In the classic stirred reactor kinetic formulation the correct concentration to be used in the equation is the exit concentration (as the system is well mixed); we have replaced this assumption with the mean concentration CA CA,INCA,OUT. The mass balance is F A;IN F A consumption of A in reactor;

where p is the porosity of the particle and tp is the tortuosity of the media. s is the constriction factor.
Table 3 Value of the modied Thiele modulus when s 1 CH2 S (ppmv) 10 000 3000 De (m2/s) 2.65 106 2.65 106 Cs (mol H2S/m3) 0.415 0.124

L (m) 8.33 103 8.33 103

r mol H2 S=m3 s 8.15 103 3.26 103

j0s 0.515 0.688

ARTICLE IN PRESS
148 L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151

Table 4 Values of modied Thiele modulus when s 0:5 CH2 S (ppmv) 10 000 3000 De (m2/s) 1.33 106 1.33 106 Cs(mol H2S /m3) 0.415 0.124 L (m) 8.33 103 8.33 103 r mol H2 S=m3 s 8.15 103 3.26 103 j0s 1.028 1.380

F A;IN F A consumption of A in reactor;


b FA;IN F A rA V kC a A CB V .

rA is the reaction rate (mg H2S reacted/cm3 adsorbent*h); k the reaction constant; CB the number of reactive sites consumed by the H2S (sites per cm3 adsorbent); a the order of reaction of CA, b the order of reaction of CB; V the volume of adsorbent (cm3 of adsorbent). According to this model, B is placed inside the semi-batch reactor and it is consumed with time while A, the H2S content in the gas, is fed into the reactor continuously. FA,IN is a constant for each experiment but FA changes with time. In this work, it was considered that the semi-batch reactor behaves as a stirred reactor (STR) while in reality the column behaves as a semi-batch Packed Bed Reactor (PBR). Thus, it is expected that the predicted efciencies will be lower than in real adsorption experiments because of the backmixing assumption. To respect the stirred reactor hypothesis, CA ought to be equal to the CA at the exit of the reactor. Since this is not sufciently close to the reality of the experimental runs and as there are no data so far to analyse the system as a PBR it was decided to use CA (CA inCA out) instead of CA CA out, which is a more realistic approach.

these calculations is that each reacted molecule of H2S consumed one adsorption site. The reaction rate is given by the following b equation: rA kC a A C B . A non-linear regression algorithm was used to evaluate the best-t kinetic parameters of this equation. The regression gave the following results:
:935 0:03 rA 0:0243 C 0 CB . A

(10)

The model proposed suggests that the reaction is close to rst order with respect to the H2S concentration and zero order with respect to the media. A comparison of the model prediction efciency is depicted in Fig. 7. A near zero order with respect to the adsorbent can be explained by the fact that these experiments are performed at conditions away from saturation. During the experiments, the adsorbent used was never saturated. The most advanced adsorption was stopped at 0.11 gH2S/gSulfatreat while saturation experiments have proven that the manufacturers claim for 0.2 gH2S/gSulfatreat is close to the reality.

3. Results and discussion In this study the parameters tested were, contact time, geometry of the column and linear velocity of the ow, initial concentration of hydrogen sulphide and the water content in the stream. The experimental runs performed led to the following results. 3.1. Contact time Contact time proved to be an important parameter. Even with a contact time of 60 s, the breakthrough curve resembles the ones characterizing the mass transfer controlled adsorption processes (no-pulse like adsorption front) and a

2.6. Estimation of the kinetic parameters The basic assumption is that the reaction rate is a function of the H2S concentration and of the sites available for the adsorption. Based on the available media saturation data provided by the experimental results, it has been calculated that 1 g of media contains 1,984 1021 available adsorption sites. The basic assumption in

ARTICLE IN PRESS
L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151
Experimental versus Model 1.8 1.6 -rA (mg H2S/cm3.h) 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0 20 40 60 80 Time (h) 100 120 140 Experimental Model
Cout/Cin 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0:00 12:00 24:00 36:00 48:00 60:00 72:00 84:00 96:00 Time (h) 30s 60s Breakthrough curves

149

Fig. 8. Breakthrough curves for two contact times tested.

Fig. 7. Experimental data versus model predictions.


500 450 400 350 H2S (ppmv) 300 250 200 150 100 50 0 0:00 12:00 24:00 36:00 48:00 60:00 72:00 Exp#5 Exp#8 Exp#18

part of the H2S is not retained by the bed and exits the column from the very rst hours of the process. The data reported in Fig. 8 were obtained by varying the height of the bed. It can be observed that the exit concentration is two times lower when the contact time is higher. Fig. 9 helps to examine the reproducibility of the tests. The differences observed are quite signicant. It is believed that the main source of discrepancy is the adsorbent itself. The SEM analyses on the adsorbent prove that the adsorbent is composed of a non-porous SiO2 matrix at the surface of which the active promoted iron oxides are physically retained. This means that the adsorbent is spatially rather non-homogeneous and it is composed of an iron-rich ne powder, in the range of 1020 mm, and iron-poor granules at the millimetre order of magnitude size. Given the heterogeneous nature of the bed material and its frangible character, we were unable to guarantee that each lling of the bed was with an identical composition of bed material especially in the ratio of reactive nes to larger and more inert matrix material.

Time (h)

Fig. 9. Reproducibility of the results.

3.2. Column geometry and linear velocity of the gas Fig. 10 presents the results obtained by maintaining the same ow rate (20L/h), contact time (30 s) and initial concentration (10 000 ppmv) and varying the bed diameter. Thus, for the same

contact time the gas stream has a lower linear velocity. From the data obtained during the 48 h test, it is shown that, with all other parameters kept constant, the column geometry and the linear velocity of the gas have a signicant impact on the efciency of the reactive adsorption. However, these data on the effect of linear velocity of the gas are equivocal. According to well-established theory, the linear velocity of the reacting gas neither affects the rate of the chemical reaction nor the internal diffusion. The only step, which can be affected positively by the increase of the velocity, is the external diffusion. Fig. 10 suggests that an increase in velocity decreases the efciency of the process, which is rather inexplicable. Since we did not do replicate experiments for these phenomena, we can only assume that the effects are within the noise shown in Fig. 9, which shows a very large variability between experiments. Our measurement

ARTICLE IN PRESS
150 L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151
Breakthrough curves 0.80 0.70 0.60 Cout/Cin 0.50 0.40 0.30 0.20 0.10 0.00 0:00 6:00 12:00 18:00 24:00 30:00 36:00 42:00 48:00 Time (h) 0.18 cm/s 0.49 cm/s Cout/Cin 1.00 0.90 0.80 0.70 0.60 0.50 0.40 0.30 0.20 0.10 0.00 0:00 24:00 48:00 72:00 96:00 120:00 10000ppmv 3000 ppmv Breakthrough curves

Time (h)

Fig. 10. Breakthrough curves for the two linear velocities at 10 000 ppmv H2S (Contact time of 30 s).

Fig. 11. Breakthrough curves for the two concentrations (Contact time of 30 s).

of breakthrough curves as a function of inlet concentration of H2S are more or less in line with the expectation that sites become saturated with increasing quantities of H2S and more rapidly at higher concentrations. The effect of water vapour is remarkable in enhancing the adsorption over the dry gas case. This deserves further study, as most of the anaerobic digestion processes have high humidity product gas streams.

Breakthrough curves 0.80 0.70 0.60 Cout/Cin 0.50 0.40 0.30 0.20 0.10 0.00 0:00 12:00 24:00 36:00 Time (h) 48:00 60:00 72:00 100% Saturation 0% Saturation

3.3. Initial concentration of H2S The comparison can be based on the concentration at the exit of the column after a period of time or the ratio COUT/CIN for the same period of time. When the contact time is 30 s and the linear velocity is maintained at 0.18 cm/s, for the rst hours of operation, the ratio COUT/CIN was similar for both concentrations. But upon time passing, the slope is more abrupt at high concentration of H2S. The speed of saturation increases more rapidly at 10 000 ppmv than at 3000 ppmv where the speed seems more constant. The adsorption rate at low concentration is decreasing more gradually. As seen in Fig. 11, the ratio COUT/CIN for 10 000 ppmv H2S is twice as high as that for 3000 ppmv H2S. At 125 h of operation, the media that treated the C IN 10 000 ppmv reached operational saturation and the media with C IN 3000 ppmv was only half consumed. Since the adsorbent has a xed number of available active sites, it is normal that the

Fig. 12. Breakthrough curves for the two levels of water presence.

saturation is faster if the initial concentration of H2S at the entrance is higher. 3.4. Water content in the stream The presence of water enhances the reaction between H2S and the media. Data from the experiments prove that under the same conditions (3000 ppmv, 60 s and 0.49 cm/s) the exit concentration of the treated gas was 10 times higher when the gas entered dry versus when it was saturated with water (see Fig. 12). This proves that the water participates in the chemical reaction with the media. The same prole was observed under other conditions (10 000 ppmv, 30 s and 0.18 cm/s). In the next steps of our work it will be examined whether the water acts as catalyst or as reactant; in the latter case kinetics must take into account the

ARTICLE IN PRESS
L.V.-A. Truong, N. Abatzoglou / Biomass and Bioenergy 29 (2005) 142151 151

presence of the water which is now hidden in the reaction constant (k).

proles along the reaction column and include non-ideal packing considerations instead of the ideal PBR system considered.

4. Conclusion An adsorption process for the removal of H2S from biogas has been tested at lab scale. The forms of the breakthrough curves obtained show that the rate of the process is not instantaneous. Both calculations and results suggest that the diffusion and the surface reaction rates are of the same order of magnitude and that the gas must be saturated in water to give optimum results. A preliminary kinetic model has been developed predicting that the rate of the process is rst order with respect to the H2S concentration and zero order with respect to the adsorption active sites. The latter is rather due to the fact that all our experiments were conducted far from the adsorbent saturation status. The process proved promising and a scaled-up unit has been built and it is actually tested at a swine farm in Quebec, Canada. Additional experimental work is underway to rene the model by measuring concentration References
[1] Schomaker A, Boerboom H, Visser A. Anaerobic digestion of agro-industrial wastes: information networks. Technical summary on gas treatment. August 2000. Project FAIRCT96-2083. [2] Vavillin VA, Vasiliev VB, Ritov SV, Ponomarev AV. Simulation model methane as a tool for effective biogas production during anaerobic conversion of complex organic matter. Bioresource Technology 1994;48(3):18. [3] Nagl G. Controlling H2S emissions. Chemical Engineering 1997;104(3):12531. [4] Lusk P. Methane recovery from animal manures: the current opportunities casebook. NREL/SR-25145. Golden Colorado. 1998. [5] Scranton Jr., Delbert C. Process and composition for increasing the reactivity of sulfur scavenging iron oxide. US Patent 5,792,438, accepted August 1998. [6] Bird RB, Stewart WE, Lighfoot EN. Transport phenomena. New York: Wiley; 1960. [7] Noll KE, Gounaris V, Hou W- S. Adsorption technology for air and water pollution control. Chelsea: Lewis Publisher Inc.; 1992 (347p).

You might also like