You are on page 1of 31

Clean Hydrogen-rich Synthesis Gas

Contract No: SES6-CT-2004-502587 Report No. CHRISGAS October 2005_WP11_D89

Deliverable Number D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming)
Nader Padban and Velentin Becher

Work Package: WP11, Responsible/Contributing Partner: TPS Distribution: PU Revision history: Rev. no. 0 1 Date 2005-08-31 2005-10-05 Change information TPS TPS

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Table of contents
1 2
2.1 2.2

Introduction ..........................................................................................1 Routes for reforming of methane and hydrocarbons .......................3


Catalytic ........................................................................................................................ 4 Thermal ......................................................................................................................... 5

3
3.1 3.2

Performance of different catalysts in reforming of methane in product gas from biomass gasification and liquid fuels ..................5
Effect of additives and calcination temperature on the activity of the Nicatalysts ...................................................................................................................... 12 Other catalysts than Ni .............................................................................................. 13

4 5
5.1 5.2 5.3 5.4 5.5 5.6

Kinetic rates and experimental data .................................................16 Steam reforming reactors..................................................................18


Indirectly heated tube reactors................................................................................. 18 Directly heated partial oxidation (POX) reformer.................................................... 19 Autothermal Reforming (ATR) .................................................................................. 20 Gas Heated Reforming (GHR) ................................................................................... 22 Comparison between different reformer concepts................................................. 23 Novel reactor design within steam reforming......................................................... 24

6 Summary.............................................................................................25 References ......................................................................................................26

i (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

1 Introduction
Catalytic steam reforming is the main route to reduce the tars and other hydrocarbons in the product gas from gasifiers and thereby to generate a hydrogen-rich gas that can be further upgraded to other energy carriers. Methane is the most difficult hydrocarbon to decompose, thus if the methane concentration is low also the concentrations of other hydrocarbon will be low. Methane will exist at levels of 5-10% in the gasifier gas. This limits the yield of hydrogen and synthesis products. Therefore, as a target, the methane content after successful reforming should preferably be reduced to less than 0.5-1% in the reformer exit gas. In a steam methane reforming (SMR) process two main reactions are involved: CH4 + H2O = CO + 3H2, CO + H2O = CO2 + H2, H = 206 kJ/mol H = -41 kJ/mol (1) (2)

The first reaction is reforming itself, while the second is the water-gas shift reaction. The overall reaction is described as follows: CH4 + 2H2O = CO2 + 4H2, H = 165 kJ/mol (3)

Since the overall reaction is endothermic, it is necessary to supply the needed heat to the reaction by some route. In steam reforming of natural gas this is accomplished by combustion of a part of the fuel in a direct-fired or indirectly fired furnaces. The equilibrium concentrations for reaction (1) are shifted to the right at high temperature and low pressure. To achieve an almost complete conversion of the methane to CO and H2, a very high temperature and long residence time at this high temperature is necessary, meaning an overall energy loss and a huge size for the methane reforming reactor. By using catalysts it is possible to reduce the temperature necessary for total conversion of the methane and tars to below 1000C, within considerable short residence time. Most of the literature and reference plants on methane steam reforming is from applications within natural gas. The first patents on steam methane reforming of natural gas were awarded to BASF in 1926 and the first reforming plants were built in the 1930s. Large-scale production started only in the beginning of 1960s following the discovery of large gas fields in Europe and the subsequent changeover from use of coal to natural gas as a feedstock. The modern reforming processes are normally designed to operate at high pressure (~ 25 30 bar) and at the temperatures around 1000 C. The reason for the high operation pressure is to save compression energy in the downstream synthesis stage. However, due to different properties of the product gas from biomass gasification, the straightforward application of the results from natural gas steam reforming to the reforming of the product gas from biomass gasification can be misleading. Natural gas is free from solid particles and it is purified from impurities such as sulphur, which is a well-known catalyst poison before processing. The product gas from biomass gasification contains a considerable amount of higher hydrocarbons than methane (linear hydrocarbons and different classes of tars). These compounds compete with methane in endothermic reforming reactions, demanding a higher energy input than that calculated for only methane reforming. The high calorific value of the natural gas permits to burn a considerable amount of the gas to increase the temperature needed for reforming reactions. The start point for the reaction is a mixture of H2O/CH4, with a relatively high CH4 concentration. Since the original mixture contains the only combustible CH4, the result from the combustion of the gas is a natural decrease in the content of the methane. 1 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming The product gas from a gasifier has a higher content of H2 and CO, than methane. These two gases are the valuable products from the gasification and make the raw material in synthesis of upgraded vehicle fuels. In combustion of such a gas from the gasifier there will be a competition between H2, CO and CH4 in combustion reactions, and if the parameters are not chosen in an optimised way, the result of the combustion can be an overall decrease in the contents of the H2 and CO rather than thermal destruction of the CH4. The total concentration of the methane at the inlet of the reformer designed for natural gas application and the reformer designed for biomass product gas is also another parameter of vital importance. In the case of natural gas, the concentration of the methane after reforming is at the same level that can be considered as the inlet concentration for a process aimed to product gas from biomass gasification. The number of references within application of steam reforming to product gas from biomass gasification is very few. The only known demonstration plants are the Mino and IGT processes. These two were developed approximately at the same time, the first one at Studsvik, Sweden and the second one at the Institute of Gas Technology (Chicago, Illinois, USA) (1). A schematic drawing of the Mino process is given in Figure 1 and Table 1 and shows the product gas composition from the IGT gasifier during pressurised O2/ steam gasification of biomass. Concerning the performance of the reforming catalysts within these two demonstration plants there is no information available in the literature. According to the results presented in Table 1, a methane concentration of between 5 and 8 %Vol. can be expected in the product gas from pressurised gasification of biomass. Compared to atmospheric O2/steam gasification of biomass these figures are almost twice high, showing the pressure impact on formation/ or on incomplete destruction of methane. For the product gas from Vrnamo plant it can be expected that the level of methane will be in the range presented in Table 1. However the effect of the differences in the operational parameters in the gas composition should be studied thoroughly before a final conclusion.

Figure 1. Schematic drawing of a pressurised O2/steam biomass gasifier.

2 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Table 1.

Content of different gases in the product gas from pressurised O2/steam gasification of biomass. Source: (1; 2)
23 816 26,4 15,2 39,0 18,4 43,0 7,4 23 899 27,5 14,8 39,8 17,2 41,5 7,4 22 982 30,7 22,4 34,9 11,9 35,0 5,3 22 754 26,0 11,0 43,1 16,8 48,0 6,2 22 816 23,8 15,8 41,1 17,5 45,0 6,0 22 821 25,5 16,9 38,4 17,3 39,0 6,8

Pressure temp Content (% Vol.) H2 CO CO2 CH4 H2O CH4

Gas composition (dry, inert free)

Water content (%Vol, and concentration of methane in the wet gas)

2 Routes for reforming of methane and hydrocarbons


To achieve a high biomass to syngas conversion it is necessary that all compounds containing C and H in the product gas are converted to CO and H2 respectively. The compounds to be converted can be classified as follows: 1) Methane and other linear hydrocarbons 2) Benzene, toluene, light aromatics 3) Compounds classified within different tar classes There are two different pathways to eliminate these compounds from the product gas: catalytic and thermal. Experiences from tar removal from biomass gasifier show an almost complete conversion of the last two mentioned categories of compounds at temperatures around 900C over different type of catalysts, Ni-based catalysts being the most investigated one. The removal of these compounds is accomplished with the increase in the product gas yield and also an increase in the concentration of the hydrogen in the gas after the catalyst, showing the steam reforming reactions of these compounds. The ability of the catalysts in reforming of methane is less studied since this compound has not been considered as problematic during these investigations, which have had the focus to eliminate the condensable tars from a combustible gas. Concerning the thermal removal of the above mentioned compounds from the product gas there are not many literature references available, especially regarding the conversion of methane to CO and H2. In the following a short survey of available catalysts and routes for reforming processes is given.

3 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming 2.1 Catalytic There are several manufactures that produce catalysts for reforming of methane and higher hydrocarbons. In Table 2 a number of these catalyst suppliers are listed. These catalysts are normally designed for specific applications within natural gas or nafta reforming industries. The behaviour for some of these catalysts has been studied in biomass gasification but in air blown conditions and mostly within atmospheric applications. There is a lack of long term experiences from the test of these catalysts in biomass product gas environment which is characterized with high concentration of tars, moderate contents of sulphurous and chlorinated compound. The short-term performance of some of these catalysts will however be described in the following sections. Company BASF, http://www.basf.com/ Ecocat/Kemira, http://www2.kemira.com/metalkat/ Alvigo ltd http://www.alvigo.ee/ ICI (catalco) http://www.ici.com Haldor Topsoe http://www.topsoe.dk/ Methanation, reforming Application Type

Tar/ammonia reduction, Granulate, monolith steam reforming, shift Environmental application, trace reduction Metallic support gas

Reforming of naphtha Catalyst on Dycat and methane, ammonia support (ceramic), synthesis, H2S removal also on monolith Naphtha reforming Nicatalysts supported on ceramic rings steam Nibased on ceramic rings and pellets, even with high sulphur tolerance of High sulphur tolerance, fuel cell application

InnovaTec http://www.tekkie.com Norta NORTA, UAB Savanori pr. 290, LTKAUNAS United Catalysts (Sued Chemie)

Steam reforming natural gas/ diesel

Different applications Transition metals on within reforming and metallic film, plasma environmental sprayed

Naphtha reforming, Ni- catalyst steam reforming ceramic rings

on

Table 2 Catalyst manufacturers

4 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming 2.2 Thermal For the thermal reaction of methane with steam exist only very limited literature. In the reviews about catalytic methane reforming the only fact mentioned is that thermal conversion should be carried out over 1500C to show feasible conversion levels (3). Only two research groups analyzed the thermal conversion more closely. Gordon et al. (4) did some experiments. Karim et al. (5) proposed a reaction mechanism and kinetic rates derived from the extensive literature available on the combustion of methane and did a computational parameter analysis. The mechanism consists of 32 reaction steps and 14 chemical species. All are given with the corresponding kinetic parameters in their paper (5). A simple one step kinetic rate expression for equation (1) is not available in the literature. The results from these two papers are summarized in the following. Gordon et al. (4) run experiments in the range between 1225C and 1516C with steamnatural gas ratios of 1.5 and 5 and with contact times of 0.21 and 4.6 seconds. Carbon formation within the reactor tube occurred in every run. The runs at the highest temperature of about 1500C showed the highest conversion rates between 96.76% and 99%. The conversion was about 1% less for the shorter contact time. This work was the only reference for the mechanism of Karim et al. (5) in the absence of more experimental work. The reaction mechanism allowed only for homogenous reactions thus predicting of carbon formation was impossible. Variations of one parameter at a time were calculated and the results shown graphically. The reforming process showed no significant dependency on pressure. Equilibrium calculations of the effect of steam to methane ratio in the feed on the hydrogen to carbon monoxide ratio in the product gas were clearly shown to be largely inadequate. Over 1500C the methane conversion rates were fast enough to allow for reaction times of a few seconds until equilibrium was reached. The influence of oxygen addition was also studied. The heat of reaction was significantly reduced while the product gas composition showed no significant change with up to 5 vol.-% oxygen added.

3 Performance of different catalysts in reforming of methane in product gas from biomass gasification and liquid fuels
A high susceptibility to deactivation of the catalyst by coke formation and thermal sintering was observed by Baker et al. (6). They investigated four different catalysts (see 0). The catalysts showed high activity for the removal of hydrocarbons and methane above 740C. In a second research project (7), they tested three catalysts and the effect of their placement. The catalysts were mixed into the main gasifier, downstream in a fixed bed or downstream in a fluidized bed. The mixing into the main gasifier and the secondary fixed bed showed a constant loss of activity with catalyst lifetime. The secondary fluidized bed showed an initial loss of activity but a levelling off after 12 15 hrs on stream on a constant level for lifetimes up to 50 hrs.

5 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Table 3. Investigated catalysts for gas upgrading from biomass gasification.

Investigator Baker et al. (6)

Investigated catalysts Harshaw 3266; United Catalyst G90C; Girdler C-13-3; PNL Ni-Co-Mo alloy

Size

Baker et al. (7) Kinoshita et al. (8) Koningen et al. (9) Aznar et al. (10; 11) Corella et al. (12)

PNL NCM; United Catalyst G-90C; Spheres, 0.5 1.0 cm ICI-46-1 chunks, fine mesh United Catalyst G-90B United Catalyst C11-9-061 UC-11-9-62; BASF G1-25S; TOPSOE Crushed RKS-1; Nickel A - D BASF BV0170 Monolith 20 40 mesh

Kinoshita et al. (8) investigated a variation of operation conditions for a secondary fluidized bed for product gas upgrading. They used one commercial methane steam reforming catalyst. They found that all tars and all methane could be converted with a variation of operation conditions space time, temperature and steam/biomass ratio. For space times over 1.3 s at a temperature of 750C all methane could be converted. They didnt mention deactivation or the duration of the test runs. Aznar et al. (10; 11) investigated 8 different commercial methane steam reforming catalysts for tar removal and gas upgrading from steam / oxygen blown biomass gasification (see 0). They found a conversion rate for methane of over 95% for reactor temperatures higher than 720C for all tested catalyst. They observed severe deactivation of the catalyst resulting in decrease of the conversion rate with time-on-stream. This happened instead of their use of a dolomite guard bed to avoid catalyst deactivation through tars. Wang et al. made experiments with naphthalene as tar model compound. Naphthalene is the least reactive tar molecule (13). They were using a Ni-Dolomite catalyst and had severe coke formation. Rapagna et al. made recently test with tri-metallic perovskite catalysts (LaNi0.3Fe0.7O3) (14). They found CH4 conversion rates up to 90% and significant experimental evidence to be able to reduce tar to a very low level (< 0.2 g Nm-3 of dry gas). They found no carbon on the catalyst after the 150 h test run. Brown et al investigated steam reforming of tar on three metal catalysts (ICI 46-1, Z409, and RZ409) in the gas from a thermally ballasted biomass gasifier (2). According to the authors all three proved effective in eliminating heavy tars (>99% destruction efficiency) and in increasing hydrogen concentration by 6-11 vol-% (dry basis). Space velocity had little effect on gas composition while increasing temperature boosted hydrogen yield and reduced light hydrocarbons (CH4 and C2H4), thus suggesting tar destruction is controlled by chemical kinetics. The study of the catalyst effect on methane conversion in this reference shows that at a temperature of 800C and a space velocity of 3000h-1, the conversion of methane over the steam reforming catalyst was around 20%. Although reactivity of the steam reforming system towards tars did not diminish during the 12 18 hours of testing, measurement of surface area and pore size distribution indicated the conversion of small pores into larger pores during high temperature operation (Figure 2).

6 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Figure 2. BJH pore volume dV/dlog D versus pore diameter of Ni-based catalyst peak Source: Brown et al. (2). Yilmaz et al (15) investigated the steam reforming activity of UCI G90B catalyst in the gas from biomass. 200 mg of sample catalyst at VHSV (Volumetric Hourly Space Velocity) ~5000, 800C, UCI G90B was used. According to the authors methane concentration in the exit stream reached 44% of its inlet concentration after 31 hours from the start, indicating that almost half of the active sites responsible for steam reforming of CH4 were deactivated. The decrease in the reactivity is probably due to coke formation on the catalyst surface according to the authors. The results from tests of different catalysts in pressurised air gasification of biomass (16) show a close to equilibrium conversion of methane at 910C, and a space velocity of 2200 l/h, over the Ni- monolith catalyst (Table 4).

Table 4. Gas composition at reactor inlet and outlet. Fixed bed conditions: temperature 900C, SV 1900/l/h, pressure: 1 bar. Monolith reactor conditions: temperature: 901C, SV 2200 1/h, pressure 5 bar (16). Sours:(16)

7 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming Analysis of the data from pressurised air gasification of biomass and biomass/plastic waste mixture in a 90 kW gasifier (17), gives the following methane conversion over commercial Ni- catalyst (Table 5). In this work the H2S content of the gas was between 50 and 150 ppm. The tar content in the gas exceeded 10 g/m3, which gave a steam to tar ratio of about 4/11.
Table 5. Conversion of methane over Ni- catalyst. Pressure: 12 bar, space velocity: ~3500h1. Source: (17)

Sawdust Temperature C CO, Vol% CO2, Vol% H2, Vol% CH4, Vol% CH4 conversion
809 819 In ~10 ~16 ~7,5 ~5,5 out ~18 ~12 ~13 ~1,8 68 874 877 In 8 - 10 ~16 ~7 6,5 8,5 out ~21 11 8 ~16 ~2 79

Sawdust + 20% plastic waste


820 830 In ~10 ~15 ~3 ~6 out ~20 ~10 ~9 3,5 4 41 870 880 In ~10 ~16 ~3 ~5,5 out ~22 ~10 ~9 2,8 1,5 60

Studying the catalytic conversion of tars, Ising et al (18) showed that it is possible to eliminate the methane in the product gas from biomass gasification (Figure 3). Their experiment was however conducted at atmospheric pressure and with air as gasification agent.

Tar conversion vs. residence time in the catalytic reactor, T=880C 920C

8 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Gas composition, heating value and gas rate at the exit of the catalytic reactor vs. residence time., T=890C 920 C
Figure 3. Performance of Ni- monolith at air gasification of biomass. Source: Ising, et al.(18).

As seen in the figure above, with residence times greater than 0,3 seconds the complete conversion of methane is achievable at temperatures above 890C, and at atmospheric conditions. The approach to equilibrium is reported in conversion of fuels other than biomass, e.g. gasoline. Ahmed et al. (19) could observe a near to equilibrium concentration profile in conversion of gasoline to syngas (Figure 4).

Figure 4. Pressure: Atmospheric, Catalyst: 2g, Fuel: gasoline 0,04 ml/min, water: 0,04ml/min, oxygen: 24,7 ml/min. Source: (19).

9 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Figure 5 Results from reforming of grade 1 diesel fuel (20): GHSV: from 2000 to 10,000 h-1 at 840C and from 2000 to 4000 h-1 880C. Oxygen : fuel ratio = 8, water:fuel: 20.

The results from test of sulphur resistant catalysts in methane conversion is rare in the literature and especially in biomass gasification application. The early experiments with this kind of catalysts is conducted in coal gasification application and with the aim to reduce the content of ammonia in the syngas. Gangwal et al (21), tested several catalysts containing Ni, Co, Mo, and W (with Al O , TiO , and other oxides as supports) by themselves or in combination with a zinc titanate sorbent in sulphur rich gas, in micro reactors and in the real gas environment from Texaco coal gas containing up to 7,500 ppmv H2S. The list of some commercial catalysts used in this work is given in Table 6.

Table 6. List of the commercial catalysts tested by Gangwal et al. (21)

Some highlights of the results within this work were the following:

10 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming HTSR-1 exhibited excellent activity for NH3 decomposition in simulated Texaco gas without H2S, at 725 C. With H2S the catalyst was poisoned but the activity could be restored at 800 C even in the presence of H2S. The catalysts containing Ni, Co, Mo, and W on a high surface area TiO support showed moderate activity (typically 10 to 20 percent decomposition) for NH3 decomposition at 725C. The TiO2 support sintered extensively at 725 C and required stabilization with ZrO2. Mixing the Ni, Co, Mo, and W catalysts with zinc titanate sorbent allowed the catalysts to function longer. As the sorbent got loaded with H2S, the exit H2S level increased, thereby decreasing the activity for NH3 decomposition.

The results of tests at 11.2 atm showed that under identical conditions, the HTSR-1 was better than C-100N. The decomposition over HTSR-1 was initially about 75 percent and stabilized at around 50 to 60 percent. But, the decomposition on C-100N decayed exponentially, indicating that under these conditions the catalyst was continuing to deactivate. During the tests for a period of time, H2S was removed from the coal gas. This increased the activity to around 90 percent decomposition. When H2S was restored, the activity fell back very quickly to its earlier steady-state value. This result clearly indicates partial but reversible poisoning of NH3 decomposition sites at 850 C by H2S. Koningen (9), who studied the sulphur tolerance of two different catalysts in a artificially mixed syngas (15 vol.% H2, 5 vol.% CH4, 15 vol.% CO, 15 vol.% CO2 and 25 vol.% N2) at atmospheric pressure concluded that in comparison with the C11-9-061 catalyst, the reactor temperature could be lowered by 60C to reach the same level of methane conversion (Figure6).

Figure 6 Comparison between the C11-90-061 and HTSR1 catalyst. Methane conversion for various H2S concentrations. 0 ppm, 100 ppm, 200 ppm. Open symbols refer to C11-9-061, closed symbols to HTSR1. Source: Koningen (9).

11 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming 3.1 Effect of additives and calcination temperature on the activity of the Ni- catalysts Courson et al (22) showed that in dry reforming of methane on a Niolivine catalyst with 6.4 wt.% of nickel oxide on olivine was very active. In addition, this catalyst meets the requirements of high syngas yield and stability with time on stream. In dry reforming of methane with CH4/CO2 = 1, methane conversion and hydrogen yield reached 95% at 800 C after the severe conditions of cycles and ageing temperature program. Natural olivine presented good characteristics to be used as biomass gasification catalyst in a fluidised bed reactor but also as nickel support. According to the authors, iron presence helps in stabilising nickel in reducible conditions. One part of nickel oxide seemed to be included into the olivine structure and maintains the reducible nickel oxide on the olivine surface. On the other hand, nickel integration in the olivine structure leads to an increase of free iron oxide, which favours reverse water gas shift reaction. This catalytic system meets all the requirements of activity, stability and attrition resistance for use in a fluidised bed for biomass steam gasification. Wang et al (23) studied the effects of promoters on catalytic activity and carbon deposition of Ni/ -Al2O3 catalysts in CO2 reforming of methane in a fixed-bed reactor. The used promoters were alkali metal oxide (Na2O), alkaline-earth metal oxides (MgO, CaO) and rare-earth metal oxides (La2O3, CeO2). In their tests CaO-, La2O3- and CeO2promoted Ni/ -Al2O3 catalysts exhibited higher stability whereas MgO- and Na2Opromoted catalysts demonstrated lower activity and significant deactivation. Metaloxide promoters (Na2O, MgO, La2O3, and CeO2) suppressed the carbon deposition, primarily due to the enhanced basicities of the supports and highly reactive carbon species formed during the reaction. In contrast, CaO increased the carbon deposition; however, it promoted the carbon reactivity. The effect of the calcinations temperature during preparation of the NiO/Al2O3 on the activity of the catalyst has been studied by Joo et al (24). In their work they tested the CH4/CO2 dry reforming on the Ni/Al2O3 catalysts calcined at temperatures 450 C and 850 C. According to their conclusions, the Ni/Al2O3 (850 C) catalyst had good activity and stability whereas the Ni/Al2O3 (450 C) catalyst showed lower activity and stability. The higher activity of the catalyst calcined at 850 C for 16 h (Ni/Al2O3 (850 C)) could be explained by the formation of the spinel structure of nickel aluminate, which was confirmed by TPR. The carbon formation rate on the Ni/Al2O3 (850 C) catalyst was very low till 20 h, and then steeply increased with reaction time without decreasing the activity for CH4 reforming. The Ni/Al2O3 (450 C) catalyst showed high carbon formation rate at the initial reaction time and then, the rate nearly stopped with continuous decreasing the activity for CH4 reforming. Even though the amount of carbon deposition on the Ni/Al2O3 (850 C) catalyst was higher than that on the Ni/Al2O3 (450 C) catalyst, the activity for CH4 reforming was also high, which could be attributed to the different type of the carbon formed on the catalyst surface. In the work of Dong et al (25) the Ni/MgO (Ni : 15 wt%) showed high activity and good stability in all the reforming reactions. Especially, it exhibited stable catalytic performance even in stoichiometric SRM (H2O/CH4 = 1). Based on the results from TPR and H2 pulse chemisorption studies, they could concluded that a strong interaction between NiO and MgO results in a high dispersion of Ni crystallite. Pulse reaction results could reveal that both CH4 and O2 were activated on the surface of metallic Ni over the catalyst, and then surface carbon species reacted with adsorbed oxygen to produce CO. 12 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming CH4 conversion on Ni/MgO catalyst, effect of O2/CH4 ratio. Source: Dong et al (25).

Roh et al (26) mentions the instability of Ni/ -Al2O3 catalysts at high temperature (>1000 K) because of the thermal deterioration of the - -Al2O3 support as well as phase transformation into -Al2O3. According to their results the Ni/ -Al2O3 has a higher carbon dioxide reforming (CDR) activity and stability than Ni/MgAl2O4, which has been used as commercial SRM catalyst.

Figure 7.

Comparison of CDR activity between 12% and 12% catalysts (reaction conditions: p=1 atm, T=1073 K, CH4/CO2/N2/=1/1/3, GHSV = 60,000 mL/gcat/h). Source: Roh et al(26).

3.2 Other catalysts than Ni Innovatek InnovaTek (27) http://www.tekkie.com has developed a proprietary catalyst formulation for the fuel processor that is being developed for use with polymer electrolyte membrane fuel cells. They have tested the catalyst for steam reforming of various hydrocarbons such as natural gas, iso-octane, retail gasoline, and hexadecane. During a 300 h continuous test the catalyst showed a stable performance for steam reforming of isooctane at 800 C with a steam/C ratio of 3.6. The same catalyst was also tested for steam reforming hexadecane (a surrogate of diesel) for 73 h as well as natural gas for over 150 h 13 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming continuously, without deactivation or carbon deposition. Sulphur tolerance of the catalyst was tested using iso-octane containing various concentrations of sulphur. There was no catalyst deactivation after a 220 h continuous test using iso-octane with 100-ppm sulphur. For comparison, a nickel catalyst (12 wt.% Ni/Al2O3) was also tested using different levels of sulphur in iso-octane. The results indicated that the InnovaTek catalyst has a substantially improved sulphur resistance compared to the nickel catalysts currently used for steam reforming. In addition, a variation of the catalyst was also used to reduce CO concentration to <1% by water gas shift reaction (27). Souza et al report about the steam reforming activity of the Pt/ZrO2/Al2O3 catalyst. At relatively low temperature (700 C), Pt/Zr catalysts show a high methane conversion capacity, approximately of about 0,27 mol/h/g catalyst (28).

Figure 8.

CH4 conversion of Pt catalysts during CO2/CH4 reforming as a function of temperature. Reaction conditions: CH4/CO2/He = 1/1/18; total feed flow rate = 200 cm3/min. Source: (28).

14 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Figure 9. Fit of the proposed kinetic model for CH4-CO2 reforming as function of methane and CO2 partial pressures on the Pt10Zr catalyst. Source: (28).

3.2.1.1 Dual Function DME Catalyst (Topsoe) Topsoe has developed a unique dual-function catalyst for low cost DME synthesis. The dual-function catalyst accelerates both of the sequential reactions in the DME synthesis; first the production of methanol from synthesis gas, subsequently the dehydration of methanol into DME. Simultaneously, the dual-function catalyst equilibrates the shift reaction, which makes it a truly flexible catalyst. The ability to equilibrate the shift reaction means that there are no limitations with respect to the synthesis gas used in the 15 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming system. If the synthesis gas originates from reforming of natural gas, where the amount of CO and CO2 is limiting the production, the multiple equilibria obtained ensure maximum conversion of the limiting species. Likewise, if the synthesis gas originates from a coal gasifier, the multiple equilibria ensures the reverse shift of CO and water to H2 and CO2, thereby maximizing the availability of hydrogen, the limiting species in case of coal derived synthesis gas The dual function catalyst has been tested in the Topsoe demonstration plant for more than 30,000 hours, and it has proven to have higher activity and better resistance to deactivation than conventional catalysts.

4 Kinetic rates and experimental data


For the industrial process of steam reforming of natural gas the gas composition is close to equilibrium for heterogeneous reactions, as the reaction rates are relatively fast. Mass transfer and diffusion to the catalyst surface and into the pores are the limiting factors. If the heat is supplied from outside, as is the case in an industrial steam reformer the heat transfer can be the limiting step. A heterogeneous reaction consists of seven steps (29):
1. 2. 3. 4. 5. 6. 7. Transport of reactants A, B, from the main stream to the catalyst pellet surface. Transport of reactants in the catalyst pores. Adsorption of reactants on the catalytic site. Surface chemical reaction between adsorbed atoms or molecules. Desorption of products R, S, . Transport of the products in the catalyst pores back to the particle surface. Transport of products from the particle surface back to the main fluid stream.

The main purpose of this paper is to analyse the kinetic reaction rates of catalytic and thermal steam reforming. Therefore the mass transfer steps are not look at more closely. It might be included in some of the given rates, depending on the experimental conditions. A lot of rate expressions for the methane steam reforming reaction (1) over nickel catalysts exist in the literature. The more recent ones are complex. Most are from research on steam reforming of natural gas. A good overview table with most models is given in (30). Under high temperature conditions (> 700C) a direct proportionality of the kinetic rate to the methane partial pressure can be observed.

d CH4 = k pCH 4 dt
with k as an Arrhenius expression:

(1)

E k = A exp a RT
16 (29)

(2)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming The differences between the models are listed in the table below:
Table 8. Kinetic models with first order dependence on methane.

Ref.
(31) (32) (33)

A
127 mol hr-1 (lb of
catalyst) -1 atm -1 hr-1 cm-2

[kJ/mole]

Ea

Catalyst
Nickel-onkieselguhr GIAP-3 7 wt% Ni/MgO

Temperature
336-637C 700-900C 550-750C

Pressure 0,1 MPa 0,1 MPa 5-450 kPa for partial pressures

70.77 81.2 102

2.1x104 liter gcatalyst 2.5x105 s-1 kPa-1 ->


turnover rate [s-1]

Earlier work at TPS and KTH came up with another kinetic rate expression. It relates the conversion of methane to the temperature. The expression is based on first order dependence on methane concentration and takes the deactivation resulting von H2S in the feed into account (9).

Wcat 1 E 0.9 = k exp a pH dx S 1 FCH4 x RT 2 0

(2)

Wcat: Catalyst amount [g]; FCH4: Molar flow of methane [mol/min]; k: Rate constant [mol/(g*min)*bar0.9]; Ea: Activation energy [kJ/mol]; pH2S: H2S partial pressure [bar]; R: Gas constant [kJ/mol*K]; T: Temperature [T]; x: CH4 conversion [-];

The activation energy was found to be between 220 kJ/mol and 285 kJ/mol depending on the surface coverage with Sulphur. There exist also a wide range of expressions with dependence on other gas compositions. A few are given in the Table 9.
Table 9 . Kinetic models for methane steam reforming.

Ref.

Ea [kJ/mole]

Catalyst

Temperature

Pressure

(34)

d CH4 Ea 1.25 = A exp pCH4 pH 2O dt RT 4.775 10 Ni/ZrO2 98 (N4Z4 cermet) Pa gcat-1 s-1 Pa-1

(3)

800 1000C

0.1 MPa

(35) 0.72 107 s-1

d CH4 Ea = A exp dt RT
75.3 GIAN-3 17 (29)

pCH4 + pH2O 2 pH + pCO 2


727 1027C 0.1 MPa

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

(36) 4.19

d CH4 Ea = A exp pCH4 pH2O dt RT


29 Ni/Al2O3

3 pCO pH 2 1 K p p eq CH4 H2O

mol gcat-1 s-1 Pa-1

n. s.

n. s.

Today the most widely used expression for modelling of natural gas steam reformers is from Fremont et al. (37). They use a complex Langmuir-Hinshelwood expression. Elanashaie et al. (38) did a parametric study on this model and found it to be the most accurate model over the widest range of parameters so far. A good summary of this model can be found in (39). But this expression is restricted to a temperature range from 500C 575C, a pressure range of 3-15 bar and molar H2O/CH4 ratios of 3-5 (40). As this is a considerably lower temperature range and higher partial pressures of methane as the conditions for methane reforming of product gas of biomass gasification, the rates are not directly useful for our purpose. More expressions can be found in the literature (30; 33; 41).

5 Steam reforming reactors


The Energy Centre of the Netherlands (ECN) has published a survey report on the steam reforming of the natural gas. This report covers fully the state of the art for the process, concept and reactor developments within the reforming of natural gas. The following is an executive summary of the ECN report (42). 5.1 Indirectly heated tube reactors In this kind of reactors a pre-heated mixture of natural gas and steam is passed through catalyst-filled tubes, allocated inside a direct heated furnace. A part of the fuel is combusted inside the furnace to generate the heat necessary for the endothermic reforming reactions inside the tubes. Depending on the position of the burners these reformers are categorised within top-fired, bottom-fired, side-fired, etc. In natural gas steam reforming, 35-50% of total energy input is absorbed by the reforming process, of which half is required for temperature rise and the other half for the reaction itself. The produced syngas leaves the reformer at a temperature of 800900C. The heat of the flue gases is usually utilised in the convective part of the reformer by generating steam and preheating the feedstock, thus bringing the overall thermal efficiency to over 85%.

Figure 10. Conventional steam methane reforming. Source: (42).

18 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming In the SMR configuration the needed energy for the endothermic reforming reaction must pass throughout to different barriers e.g. from the combusted fuel to the tube walls and from the tube walls to the catalyst inside the tubes. According to the results from Rajesh et al (43) who simulated the performance an existing side-fired steam reformer, the overall temperature difference between the two gases inside and outside the tubes (product gas to be reformed, flue gas used as the heat source), can be between 450C and 250C. Their simulation show that the largest temperature difference happens in the inlet of the catalyst tubes, due to the high molar conversion of the methane in this position. Application of SMR to biomass gasification system can be problematic, if a part of the product gas is planned to be used as the reformer fuel. The reasons are as follows: 1) The product gas has a much lower energy content than natural gas, meaning that a larger quantity of the product gas must be burned 2) A high degree of burning of the product gas affects directly the overall conversion of biomass to synthesis gas 3) The content of the inert compounds such as H2O and CO2 is high in the product gas. This will cause differences in the fluid-dynamic properties of the system resulting in an even lower heat flux and consequently a higher degree of combustion. Typical operating parameters of the SMR process are: pressure 20-26 bar, temperature 850950C H2/CO ratio 2,96,5. Complete conversion cannot be obtained in the SMR process: typically 65% of methane is converted, at best it is about 98% (44), so secondary reforming must be used if a higher conversion rate is desired. The SMR process is the most proven technology with a great deal of industrial experience; it requires no oxygen and produces syngas with a high H2/CO ratio. It also has relatively low operating temperatures and pressures in comparison to other technologies. Nevertheless, expensive catalyst tubing and a large heat recovery section make an SMR plant a costly investment that can only be justified for very large-scale production. Of the syngas production technologies, steam methane reforming is the most developed and commercialised. Lurgi, for example, has built more than 100 plants to date (45). Many engineering companies design and build SMR plants, among them M.W. Kellogg, Haldor Topse, ICI, Howe-Baker, KTI, Foster Wheeler, Kvrner. 5.2 Directly heated partial oxidation (POX) reformer In these types of reactors a part of the fuel is combusted inside the reformer to supply the heat necessary for the endothermic reforming reactions. A refractory-lined pressure vessel is fed with natural gas and oxygen at a typical pressure of 40 bar. Both natural gas and oxygen are preheated before entering the vessel and mixed in a burner. Partial oxidation reaction occurs immediately in a combustion zone below the burner. To avoid carbon deposition the reactants should be thoroughly mixed and the reaction temperature should not be lower than 1200C. Sometimes steam is added to the mixture to suppress carbon formation. In the case of catalytic partial oxidation steam is not required and the temperature can be below 1000C.

19 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Figure 11. Partial oxidation reformer. Source: (42).

The syngas produced leaves the reactor at temperatures of 13001500C. The syngas from the POX process has a H2/CO ratio between 1.6 and 1.8, so a shift converter or steam injection is necessary to increase this ratio, for instance, for methanol or DME synthesis. The non-catalytic process allows the use of a broad range of hydrocarbon fuels from natural gas to coal and oil residue and remains the only viable technology for heavy hydrocarbons. The catalytic process has a reduced size and consumes less oxygen, but runs the risk of catalyst destruction by local thermal stress. According to the SINTEF study (46), the investment costs for a POX-based syngas plant constitute 80% of the reference SMR plant. The oxygen costs can constitute 50% of operational costs of the syngas production at the POX plant (46). Syngas production via the POX route is an established technology. Texaco and Shell technologies have been employed for many years for partial oxidation of petroleum cuts and other heavy hydrocarbons. In the field of coal gasification, along with Texaco and Shell, other companies are active in this field such as Lurgi, Koppers, Foster Wheeler, British Gas, Starchem. In 1992, Texaco had more than 100 licensed commercial POX plants on their reference list, of which 28 were using gaseous and 62 were using liquid feedstock (47). Use of non- catalytic POX for reforming of the product gas from biomass gasification will at the first hand result in combustion of the existing H2 in the product gas. To achieve the high exit temperature, which is necessary for eliminating the methane, a considerable amount of the gas should be burned. 5.3 Autothermal Reforming (ATR) This process combines partial oxidation and steam reforming in one vessel, where the hydrocarbon conversion is driven by heat released in the POX reaction. Developed in the late1950s by Haldor Topse and Socit Belge de lAzote (48) the process is used for methanol and ammonia production. Both light and heavy hydrocarbon feed stocks can be converted. In the latter case, an adiabatic pre-reformer is required. In this process a preheated mixture of natural gas, steam and oxygen is fed through the top of the reactor. In the upper zone, partial oxidation proceeds at a temperature of around 1200C. After 20 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming that, the mixture is passed through a catalyst bed, where final reforming reaction takes place. The catalyst destroys any carbon formed at the top of the reactor. The outlet temperature of the catalyst bed is between 850 and 1050C. The main advantages of ATR are a favourable H2/CO ratio (1.6 to 2.6), reduction of emissions due to internal heat supply, a high methane conversion, and the possibility to adjust the syngas composition by changing the temperature of the reaction. However, it requires an oxygen source. The capital costs for autothermal reforming are lower than those of the SMR plant by 25%, as reported by Haldor Topse (49). Operational costs, however, are the same or even higher due to the need to produce oxygen. The SINTEF study (46) reported a capital-cost reduction of 35%, but an 8%-increase in operational costs for the ATR technology in comparison to the SMR process. ATR technology is commercially available, but still has limited commercial experience. The main licensors are Haldor Topse, Lurgi, ICI, Foster Wheeler.

Figure 12. Typical sketch of an autothermal reformer. Source:(50).

The heat transfer to the catalyst bed is more favourable in an autothermal reformer than in the externally heated tubular reformers, since in the former case the heat in the gas is supplied directly to the catalyst bed. This means that a high temperature in the catalyst bed can be achieved by burning only a small portion of the product gas. The quantity of the gas to be burned will be dependant to the inlet concentration of the methane and other reformable compounds (such as tars) in the gas. It is more likely that the initial 21 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming temperature increase in the combustion zone will reduce the concentration of the tars and other hydrocarbons sharply. However it must be taken to account that the combustion reaction will consume a part of the hydrogen that is present in product gas. The choice of the outlet temperature of such a reformer in pressurised biomass gasification case will be dependant on the composition of the gas and especially its steam content. At low steam contents in the gas a high outlet temperature should be chosen for avoiding the methanation reaction, which is favoured at high pressure and low temperature. According Waldheim et al (50) who simulated a pressurised autothermal reformer in pressurised biomass gasification application to achieve an outlet temperature of around 850C in the exit of the catalyst bed, it is necessary to have a temperature of around 1300C in the inlet of the bed (Figure 13).

Figure 13.Steam reformer temperature and conversion profiles with biomass gas (50).

5.4 Gas Heated Reforming (GHR) In the gas heated reformer (GHR) concept the heat for the endothermic reaction is supplied by cooling down the reformed gas from the secondary reformer. This technology, originally developed in the 1960s by ICI, was first demonstrated during 1988 at two ammonia plants in Severnside, UK. The feed in the gas-heated reformer is passed first to the primary reformer where about 25% of reforming takes place. The partially reformed gas is then passed to a secondary oxygen-fired reformer. The effluent of the latter is used to heat up the feed in the primary reformer. For start-up, an auxiliary burner is employed.

22 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Figure 13. Gas heated reformer. Source: (42).

The volume of a GHR is typically 15 times smaller than the volume of a fired reformer (SMR or CO2) for the same output (51). Overheating of hot metal parts and a poor temperature control can lead to problems concerning the reliable operation of heat exchange reformers. To overcome these problems, reformers usually use counter-current flows in the low-temperature part with effective heat transfer and co-current flows in the hot section for a better temperature control. Sogge et al (46) estimated that the GHR plant would cost about 40% less to build than a comparable SMR plant, while operational costs would be about the same. According to Abbott (52), the GHR scheme requires 33% less oxygen than the ATR plant. The main developer of GHR technology is ICI and its subsidiary Synetix. To date three ammonia plants (two in the UK, one in the USA) and one methanol plant (Australia) have been built using this syngas technology. This concept can be an attractive option within biomass gasification if the product gas is cooled down before the reforming step, e.g. during gas filtration. 5.5 Comparison between different reformer concepts
Table 10. Comparison between different reformer concepts. Source: (42; 42)
SMR Temperature, C Pressure, bar H2/CO ratio 800 900 20 30 36 POX 1000 1450 30 85 1,6 2 95 100 high optional 80 110 low small to large commercial ATR 850 1300 20 70 1,6 2,5 95 100 high high 65 80 low large commercial GHR Primary: 450, secondary: 1000 95 100 3,4 95 100 medium medium 60 80 low medium large 3 commercial units

CH4 conversion, % 65 95 Oxygen Steam consumption Capital cost, % Emissions Scale Development status none high 100 (refer) high large commercial

23 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

A comparison between the above-mentioned reforming concepts is given in Table 10 above. In a review report the Norwegian Statoil Research & Technology Centre (53) has made in a similar manner a comparison between different reforming technologies for production of FT- fuels from natural gas. Also according to these authors, the gas heated reformer concept has a relatively lower cost than the other investigated processes. The difference mentioned in this work is however much less than the figures presented by ECN.
Table 11. Advantages and disadvantages for different synthesis gas technologies. Source: (53)

Synthesis gas technology CSR (catalytic steam reforming) ATR

Cost, relative to ATR and include direct costs >100

No oxygen

Very high H2/CO High s/c- ratio, thus high CO2/CO

Near optimum H2/CO Low s/c ratio Near optimum H2/CO Low s/c ratio High CO/CO2

Oxygen need SN <2 Higher oxygen consumption than ATR, SN <2

100

POX

>100

GHR + ATR

Near optimum H2/CO Moderate oxygen consumption

Unproven at s/c-ratio 94 <2

5.6 Novel reactor design within steam reforming Fraunhofer Institute has developed steam reforming, autothermal reforming and partial oxidation that can be used for reforming and partial oxidation of natural gas and other hydrocarbons. This compact steam reformer consists of a heat exchanger for vaporizing the feed water and preheating the feed gas (natural gas and steam) to the reformers inlet temperature (450-600C), a two-part reforming reactor and a low-NOx radiation burner. The novel concept in the construction of FI reformer is the use of a commercial Pt/Al2O3 catalyst on a metal honeycomb support as the reforming catalyst. Use of such a metal support catalyst system enhances, the heat transfer properties compared to a conventional fixed bed catalyst system. The outer wall of the second reforming section is directly heated by radiation of the ceramic type burner, while the hot flue gas heats the inner reactor convectively (54). 24 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

Figure 14. Schematic description of FI compact reformer. Source:(54)

6 Summary
Production of a hydrogen rich gas from biomass requires conversion of all hydrocarbons present in the product gas to gaseous hydrogen and CO or CO2. The hydrocarbons to be converted are tars and linear hydrocarbons such methane, ethane and etc. In this subject methane, due to its thermal stability, becomes the most problematic compound. Thermal and catalytic routs can be applied in reforming of methane to hydrogen and carbonaceous products. The available literature on thermal conversion is very rare and from the applications within natural gas reforming. In these processes the partial combustion of the gas by an oxidant supplies the heat needed for thermal conversion. The gas to be reformed is initially free from other combustibles such as hydrogen and CO. As the reforming or partial combustion reactions exceed forward these products are produced rather than consumed. Since a product gas from gasification has already considerable amounts of these combustibles, the selectivity of the combustion reactions towards the important constituents in the gas such as H2 and CO in comparison to methane will become very important and must be studied thoroughly. Commercially there are several catalysts available for methane reforming. These catalysts are however designed for being used in clean gas environments, which is the case in natural gas processes. In literature several kinetic expressions is available for catalytic methane reforming but from conditions that are very different from the case for biomass gasification. The long-term stability of available catalysts on biomass gasification environment is not mentioned in the litterateur; also the data about deactivation of these catalysts by impurities in the gas is very rare. There are a variety of commercial reformer concepts available in the literature. Comparison of the economics for different configurations shows that ATR concept can be the best choice in O2/steam gasification. 25 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming

References
1. Robert J., E. B. S. "Operation results from a pressurised fluidized bed biomass asification process. In: "Energy from Biomass and Wastes IX; Publ: Institute of Gas Technology, Chicago, Illinois, USA. 573-593. 1985. 2. Brown, R. C., Smeenk, J., Sadaka, S., Norton, G., Zhang, R., Suby, A., Cummer, K., Ritzert, J., Xu, M., Lysenko, S., Nunez, J., and Brown, N., "Biomass-driven hydrogen from a thermally ballasted gasifier - Final Technical Report," Iowa State University, DOE Award Number: DE-FC36-01GO11091, Ames, IA, USA, 2005. 3. Rostrup-Nielsen, J. R., "Catalytic Steam Reforming," Catalysis - Science and Technology, Vol. 5, Springer Verlag, Berlin, 1984, pp. 1-117. 4. Gordon, A. S., "Uncatalyzed Reaction of natural gas and steam," Industrial & Engineering Chemistry, Vol. 38, No. 7, 1946, pp. 718-720. 5. Karim, G. A. and Metwally, M. M., "A kinetic Investigation of the reforming of natural gas for the production of hydrogen," International Journal of Hydrogen Energy, Vol. 5, 1980, pp. 293-304. 6. Sutton, D., Kelleher, B., and Ross, J. R. H., "Review of literature on catalysts for biomass gasification," Fuel Processing Technology, No. 73, 2001, pp. 155-173. 7. Baker, E. G., Mudge, L. K., and Brown, M. D., "Steam gasification of biomass with nickel secondary catalysts," Industrial & Engineering Chemical Research, Vol. 26, 1987, pp. 1335-1339. 8. Kinoshita, C. M., Wang, Y., and Zhou, J., "Effect of Reformer Conditions on Catalytic Reforming of Biomass-Gasification Tars," Industrial & Engineering Chemical Research, Vol. 34, 1995, pp. 2949-2954. 9. Koningen, J.. Upgrading and cleaning of gasified biomass for advanced applications. 1997. Royal Institute of Technology (KTH). 10. Caballero, M. A., Aznar, M. P., Gil, J., Martin, J. A., Frances, E., and Corella, J., "Commercial Steam Reforming Catalysts To Improve Biomass Gasification with Steam-Oxygen Mixtures. 1. Hot Gas Upgrading by the Catalytic Reactor," Industrial & Engineering Chemical Research, Vol. 36, 1997, pp. 5227-5239. 11. Aznar, M. P., Caballero, M. A., Gil, J., Martin, J. A., and Corella, J., "Commercial Steam Reforming Catalysts To Improve Biomass Gasification with Steam-Oxygen Mixtures. 2. Catalytic Tar Removal," Industrial & Engineering Chemical Research, Vol. 37, 1998, pp. 2668-2680. 12. Corella, J., Toledo, J. M., and Padilla, R., "Catalytic Hot Gas Cleaning with Monoliths in Biomass Gasification in Fluidized Beds. 1. Their Effectiveness for Tar Elimination," Industrial & Engineering Chemical Research, Vol. 43, 2004, pp. 24332445.

26 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming 13. Wang, T. J., Chang, J., Wu, C. Z., Fu, Y., and Chen, Y., "The steam reforming of naphthalene over a nickel - dolomite cracking catalyst," Biomass & Bioenergy, Vol. 28, 2005, pp. 508-514. 14. Rapagna, S., Provendie, H., Petit, C., Kiennemann, A., and Foscolo, P. U., "Development of catalysts suitable for hydrogen or syn-gas production from biomass gasification," Biomass & Bioenergy, Vol. 22, 2002, pp. 377-388. 15. Yilmaz, S. and Cowley, S. W., "Catalyst Lifetime testing for methanol syngas conditioning," Colorado School of Mines, Contract number: XAE-3-13442-01, Golden, CO, USA, 1996. 16. Simell, P., Stahlberg, P., Solantausta, Y., Hepola, J., and Kurkela, E., "Gasification gas cleaning with nickel monolith catalyst," Vol. 2, Chapman & Hall, London, 1997, pp. 1103-1116. 17. Padban N.. PFB Air Gasification of Biomass, Investigation of Product Formation and Problematic Issues Related to Ammonia, Tar and Alkali. 2000. Department of Chemical Engineering, Lund University. 18. Ising, M., Gil, J., and Unger, C., "Gasification of biomass in a circulating fluidized bed with special respect to tar reduction," Sevilla, Spain, 2000. 19. Ahmed, S., Krumpelt, M., Kumar, R., Lee, S. H. D., Carter, J. D., Wilkenhoener, R., and Marshall, C., "Catalytic partial oxidation reforming of hydrocarbon fuels," DOE: ANL/CMT/CP-96059, Argonne, IL, USA, 1998. 20. Pereira, C., Bae, J.-M., Ahmed, S., and Krumpelt, M., "Liquid Fuel Reformer Development: Autothermal Reforming of Diesel Fuel," Argonne National Laboratory, DOE: ANL/CMT/CP-102382, Argonne, IL, USA, 2000. 21. Gangwal, S. K., Gupta, R. P., Portzer, J. W., Turk, B. S., Krishnan, G. N., Hung, S. L., and Ayala, R. E., "Catalytic Ammonia Decomposition for Coal-Derived Fuel Gases," Research Triangle Institute, DOE: DOE/MC/29011-97/C0731, Research Triangle Park, NC, 1996. 22. Courson, C., Udron, L., Swierczynski, D., Petit, C., and Kiennemann, A., "Hydrogen production from biomass gasification on nickel catalysts Tests for dry reforming of methane," Catalysis today, Vol. 76, 2002, pp. 75-86. 23. Wang, S. and Lu, G., "Effects of promoters on catalytic activity and carbon deposition of Ni/ g-Al2O3 catalysts in CO2 reforming of CH4," Journal of Chemical Technology and Biotechnology, Vol. 75, 2000, pp. 589-595. 24. Joo, O.-S. and Jung, K.-D., "CH4 Dry Reforming on Alumina-Supported Nickel Catalyst," Bulletin of the Korean Chemical Society, Vol. 23, No. 8, 2002, pp. 1149-1153. 25. Dong, W.-S., Roh, H.-S., Liu, Z.-W., Jun, K.-W., and Park, S.-E., "Hydrogen Production from methane reforming reactions over Ni/MgO Catalyst," Bulletin of the Korean Chemical Society, Vol. 22, No. 12, 2001, pp. 1323-1327.

27 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming 26. Roh, H.-S., Jun, K.-W., Baek, S.-C., and Park, S.-E., "Carbon Dioxide Reforming of Methane over Ni/ q-Al2O3 Catalysts: Effect of Ni Content," Bulletin of the Korean Chemical Society, Vol. 23, No. 8, 2002, pp. 1166-1168. 27. Ming, Q., Healey, T., Allen, L., and Irving, P., "Steam reforming of hydrocarbon fuels," Catalysis today, Vol. 77, No. 1-2, 2002, pp. 51-64. 28. Souza, M. M. V., Aranda, D. A. G., and Schmal, M., "Reforming of Methane with Carbon Dioxide over Pt/ZrO2/Al2O3 Catalysts," Journal of Catalysis, Vol. 204, No. 2, 2001, pp. 498-511. 29. Froment, G. F. and Bischoff, K. B., Chemical Reactor Analysis and Design, Second ed., John Wiley & Sons, New York, 1990. 30. Elnashaie, S. S. E. H., Al-Ubaid, A. S., Soliman, M. A., and Adris, A. M., "On the Kinetics and Reactor Modelling of the Steam Reforming of Methane - A Review," Journal of Engineering Sciences, King Saud University, Vol. 14, No. 2, 1988, pp. 247-273. 31. Akers, W. W. and Camp, D. P., "Kinetics of the methane-steam reaction," AiChE Journal, Vol. 1, No. 4, 1955, pp. 471-475. 32. Bodrov, I. M., Apel'baum, L. O., and Temkin, M. I., "Kinetics of the reaction of methane with water vapor, catalyzed by nickel on a porous carrier," Kinetics and catalysis, Vol. 8, 1967, pp. 696-702. 33. Wei, J. and Iglesia, E., "Isotopic and kinetic assessment of the mechanism of reactions of CH4 with CO2 or H2O to form synthesis gas and carbon on nickel catalysts," Journal of Catalysis, Vol. 224, 2004, pp. 370-383. 34. Lee, A. L. and Zabransky, R. F., "Internal reforming development for solid oxide fuel cells," Industrial & Engineering Chemical Research, Vol. 29, 1990, pp. 766-773. 35. Nekrich, E. M., "Kinetics of partial oxidation of natural gas by a small amount of steam," Journal of Applied Chemistry (USSR), Vol. 43, No. 2, 1970, pp. 372-378. 36. Jin, W., Gu, X., Li, S., Huang, P., Xu, N., and Shi, J., "Experimental and simulation study on a catalyst packed tubular dense membrane reactor for partial oxidation of methane to syngas," Chemical Engineering Science, Vol. 55, 2000, pp. 2617-2625. 37. Xu, J. and Froment, G. F., "Methane Steam Reforming, Methanation and Watergas Shift: 1. intrinsic Kinetics," AiChE Journal, Vol. 35, No. 1, 1989, pp. 88-96. 38. Elnashaie, S. S. E. H., Adris, A. M., Al-Ubaid, A. S., and Soliman, M. A., "On the non-monotonic behaviour of methane-steam reforming Kinetics," Chemical Engineering Science, Vol. 45, No. 2, 1990, pp. 491-501. 39. Galluci, F., Paturzo, L., and Basile, A., "A simulation study of the steam reforming of methane in a dense tubular membrane reactor," International Journal of Hydrogen Energy, Vol. 29, 2004, pp. 611-617. 40. Rostrup-Nielsen, J. R. and Sehested, J., "Hydrogen and Synthesis Gas by Steamand CO2 Reforming," Advances in Catalysis, Vol. 47, 2002, pp. 65-139.

28 (29)

CHRISGAS October 2005_WP11_D89 Literature and State-of-the-Art review (Re: Methane Steam Reforming 41. Leinfelder, R.. Reaktionskinetische Untersuchungen zur Methan-DampfReformierung und Shift-Reaktion an Anoden oxidkeramischer Brennstoffzellen. Dissertation . 2004. Universitt Erlangen-Nrnberg. Ref Type: Thesis/Dissertation 42. Korobitsyn, M. A., van Berkel, F. P. F., and Christie, G. M., "SOFC as a separator," Energy Center of the Netherlands, ECN-C-00-122, 2000. 43. Rajesh, J. K., Gupta, S. K., Rangaiah, G. P., and Ray, A. K., "Multiobjective Optimization of Steam Reformer Performance Using Genetic Algorithm," Industrial & Engineering Chemical Research, Vol. 39, No. 3, 2000, pp. 706-717. 44. Appl, M., "Modern Ammonia Technology: Where Have We Got To. Where We Are We Going?," Nitrogen, Vol. 199, 1992, pp. 46-74. 45. Hydrocarbons, Gas and Chemicals. Lurgi . 1999. 5-4-2005. 46. Sogge, J., Strom, T., and Sundset, T., "Technical and Economic Evaluation of Natural Gas Based Synthesis Gas Production Technologies," SINTEF, STF21 A93106, Trondheim, Norway, 1994. 47. Falsetti, J. S., "Gasification process for maximising refinery profitability," Hydrocarbon Technology International, 1993, pp. 57-61. 48. Chauvel, A. and Lefebvre, G., Petrochemical Processes, Part 1: Synthesis-Gas Derivatives and Major Hydrocarbons, Editions Technip, Paris, France, 1989. 49. Dybkjaer, I. and Madsen, S. W., "Advanced Reforming Technologies for Hydrogen Production," The International Journal of Hydrocarbon Engineering, Vol. December-January 1997-1998, 1997. 50. Waldheim, L. B. M. N. T., "Gas cleaning for advanced applications," TPS, TPS 96/54, Studsvik, Sweden, 1996. 51. Kitchen, D. and Mansfield, K., "ICI's new synthesis gas technology," European Applied Research Conference on Natural Gas Eurogas '92, Trondheim, Norway, 1992. 52. Abbott, J.. GTL Syngas Generation Using Synetix Gas Heated Reforming Technology. 1998. Synetix.

53. Olsvik, O., Schanke, D., Sogge, J., de Bruyn, H. J., and Hansen, R., "Technology Development Trends and Challenges for New Mega GTL Plants," 2005. 54. Vogel, B. and Heinzel, A., "Reformer development at Fraunhofer ISE," Fuel Cells Bulletin, Vol. 1, No. 3, 1998, pp. 9-11.

29 (29)

You might also like