6

Gasification processes for syngas and hydrogen production

J.G. Speight    CD&W Inc., Laramie, WY, USA

Abstract

The choice of technology for synthesis gas production depends on the scale of the synthesis operation. Syngas production from solid fuels can require an even greater capital investment with the addition of feedstock handling and more complex syngas purification operations. The greatest impact on improving gas-to-liquids plant economics is to decrease capital costs associated with syngas production and improve thermal efficiency through better heat integration and utilization. Improved thermal efficiency can be obtained by combining the gas-to-liquids plant with a power generation plant to take advantage of the availability of low-pressure steam.

This chapter presents the means by which synthesis gas and hydrogen are produced from carbonaceous feedstocks.

Keywords

Steam-methane reforming

Autothermal reforming

Combined reforming

Partial oxidation

Hydrogen production

Product composition and quality

6.1 Introduction

Gasification processes are used to convert a carbon-containing (carbonaceous) material into a synthesis gas (syngas), which is a combustible gas mixture that typically contains carbon monoxide, hydrogen, nitrogen, carbon dioxide, and methane. The impure synthesis gas has a relatively low calorific value, ranging from 100 to 300 Btu/ft3. The gasification process can accommodate a wide variety of gaseous, liquid, and solid feedstocks and it has been widely used in commercial applications for the production of fuels and chemicals (Chapters 1 and 10). Conventional fuels such as coal and petroleum, as well as low- or negative-value materials and wastes such as petroleum coke, refinery residue, refinery waste, municipal sewage sludge, biomass, hydrocarbon contaminated soils, and chlorinated hydrocarbon by-products have all been used successfully in gasification operations (Speight, 2008, 2013a, 2013b). In addition, syngas is used as a source of hydrogen or as an intermediate in producing a variety of hydrocarbon products by means of the Fischer-Tropsch synthesis (FTS) (Table 6.1) (Chadeesingh, 2011). In fact, gasification to produce synthesis gas can proceed from any carbonaceous material, including biomass and waste.

Table 6.1

General carbon ranges and common names of hydrocarbons produced from synthesis gas by the Fischer-Tropsch process

Carbon number rangeCommon name
C1-C2SNG (Synthetic Natural Gas)
C3-C4LPG (Liquefied Petroleum Gas)
C5-C7Light petroleum
C8-C10Heavy Petroleum
C11-C17Middle distillate – kerosene, diesel
C18-C30Soft wax
C31-C60Hard wax

The synthesis of hydrocarbons from the hydrogenation of carbon monoxide was discovered in 1902 by Sabatier and Sanderens, who produced methane by passing carbon monoxide and hydrogen over nickel, iron, and cobalt catalysts. At about the same time, the first commercial hydrogen from syngas produced from steam methane reforming was commercialized. Haber and Bosch discovered the synthesis of ammonia from hydrogen and nitrogen in 1910 and the first industrial ammonia synthesis plant was commissioned in 1913. The production of liquid hydrocarbons and oxygenates from syngas conversion over iron catalysts was discovered in 1923 by Fischer and Tropsch. Variations on this synthesis pathway were soon to follow for the selective production of methanol, mixed alcohols, and iso-hydrocarbon products. Another outgrowth of FTS was the hydroformylation of olefins discovered in 1938.

In principle, synthesis gas can be produced from any hydrocarbon feedstock, which include natural gas, naphtha, residual oil, petroleum coke, coal, biomass, and municipal or industrial waste (Chapter 1). The product gas stream is subsequently purified (to remove sulfur, nitrogen, and any particulate matter) after which it is catalytically converted to a mixture of liquid hydrocarbon products. In addition, synthesis gas may also be used to produce a variety of products, including ammonia, and methanol.

Of all of the carbonaceous materials used as feedstocks for gasification process, coal represents the most widely used feedstocks and, accordingly, the feedstock about which most is known. In fact, gasification of coal has been a commercially available proven technology (Speight, 2013a, 2013b) (Chapters 1 and 10). The modern gasification processes have evolved from three first-generation process technologies: (1) Lurgi fixed-bed reactor, (2) high-temperature Winkler fluidized-bed reactor, and (3) Koppers-Totzek entrained-flow reactor. In each case steam/air/oxygen are passed through heated coal, which may either be a fixed bed, fluidized bed or entrained in the gas. Exit gas temperatures from the reactor are 500 °C (930 °F), 900 to 1100 °C (1650 to 2010 °F), and 1300 to 1600 °C (2370 to 2910 °F), respectively. In addition to the steam/air/oxygen mixture being used as the feed gases, steam/oxygen mixtures can also be used in which membrane technology and a compressed oxygen-containing gas is employed.

However, the choice of technology for synthesis gas production also depends on the scale of the synthesis operation. Syngas production from solid fuels can require an even greater capital investment with the addition of feedstock handling and more complex syngas purification operations. The greatest impact on improving gas-to-liquids plant economics is to decrease capital costs associated with syngas production and improve thermal efficiency through better heat integration and utilization. Improved thermal efficiency can be obtained by combining the gas-to-liquids plant with a power generation plant to take advantage of the availability of low-pressure steam.

This chapter presents the means by which synthesis gas and hydrogen are produced from carbonaceous feedstocks.

6.2 Synthesis gas production

Both nonrenewable and renewable energy sources are important for production of synthesis gas and hydrogen. As energy carriers, hydrogen and synthesis can be produced from catalytic processing of various hydrocarbon fuels, alcohol fuels, and a variety of biofuels and biomass feedstocks.

In most cases, synthesis gas is produced from coal (gasification, carbonization), natural gas, and light hydrocarbons such as propane gas (steam reforming, partial oxidation, autothermal reforming, plasma reforming); petroleum fractions (dehydrocyclization and aromatization, oxidative steam reforming, pyrolytic decomposition); biomass (gasification, steam reforming, biologic conversion); and water (electrolysis, photo-catalytic conversion, chemical and catalytic conversion) (Liu, Song, & Subramani, 2010; Speight, 2008,, 2011a, 2013a, 2013b, 2014; Wesenberg & Svendsen, 2007). The relative competitiveness of different options depends on the economics of the given processes, which, in turn, depend on many factors such as the (1) suitability and availability of the feedstock, (2) efficiency of the catalysis, (3) scale of production, (4) required hydrogen purity, and (5) economics of feedstock production and the processing steps.

Current commercial processes for synthesis gas and hydrogen production largely depend on fossil fuels both as the source of hydrogen and as the source of energy for the production processing. Fossil fuels are nonrenewable energy resources, but they provide a more economical path to hydrogen production in the near term (next three decades) and perhaps they will continue to play an important role in the midterm (up to 50 years from now) (Speight, 2011b). Alternative processes need to be developed that do not depend on fossil hydrocarbon resources for either the hydrogen source or the energy source, and these processes need to be economical, environmentally friendly, and competitive. Efficient separation of the hydrogen from the gaseous products is also a major issue that must be addressed. In this respect, pressure swing adsorption (PSA) is used in current industrial practice. Furthermore, several types of membranes are being developed that would, when incorporated into the separation process, enable more efficient gas separation (Ho & Sirkar, 1992).

The process for producing syngas involves three individual components: (1) synthesis gas generation, (2) waste heat recovery, and (3) gas processing (Speight, 2013a, 2013b, 2014). Within each of the three listed systems are several options. Synthesis gas can be generated to yield a range of compositions ranging from high-purity hydrogen to high-purity carbon monoxide. Two major routes can be utilized for high-purity gas production: (1) pressure swing adsorption and (2) utilization of a cold box, where separation is achieved by distillation at low temperatures. In fact, both processes can also be used in combination as well. Unfortunately, both processes require high capital expenditure. However, to address these concerns, research and development is ongoing and successes can be measured by the demonstration and commercialization of technologies such as permeable membrane for the generation of high-purity hydrogen, which in itself can be used to adjust the H2/CO ratio of the synthesis gas produced.

6.2.1 Steam-methane reforming

Steam-methane reforming is the benchmark process that has been employed over a period of several decades for hydrogen production. The process involves reforming natural gas in a continuous catalytic process in which the major reaction is the formation of carbon monoxide and hydrogen from methane and steam:

CH4+H2O=CO+3H2ΔH298K=+97,400Btu/lb

si1_e

Higher molecular weight feedstocks can also be reformed to hydrogen:

C3H8+3H2O3CO+7H2

si2_e

That is,

CnHm+nH2OnCO+0.5m+nH2

si3_e

In the actual process, the feedstock is first desulfurized by passage through activated carbon, which may be preceded by caustic and water washes. The desulfurized material is then mixed with steam and passed over a nickel-based catalyst (730 to 845 °C, 1350 to 1550 °F and 400 psi. Effluent gases are cooled by the addition of steam or condensate to about 370 °C (700 °F), at which point carbon monoxide reacts with steam in the presence of iron oxide in a shift converter to produce carbon dioxide and hydrogen:

CO+H2O=CO2+H2

si4_e

The carbon dioxide is removed by amine washing; the hydrogen is usually a high-purity (> 99%) material.

Steam reforming of natural gas (sometimes referred to as steam-methane reforming, SMR) is the part of the gas-refining process where the natural gas is converted to syngas, which is further used in the synthesis to methanol or Fischer-Tropsch products. Hydrogen-rich synthesis gas can also be used directly for hydrogen enrichment. The technology for steam reforming is of great interest because this part of the process represents a substantial portion of the investment costs. The reforming section costs about 60 to 80% of the total cost of the entire gas-refining plant. Improvements and cost savings in the reforming section will therefore become very noticeable in the total plant cost.

Steam reforming is an exothermic reaction that is carried out by passing a preheated mixture comprising methane (sometimes substituted by natural gas having high methane content) and steam through catalyst-filled tubes. The products of the process are a mixture of hydrogen, carbon monoxide, and carbon dioxide. To maximize the conversion of the methane feed, primary and secondary reformers are often used – the primary reformer elicits a 90 to 92% v/v conversion of methane. Here, the hydrocarbon feed is partially reacted with steam at 900 °C (1650 °F) at 220 to 500 psi over a nickel-alumina catalyst to produce a synthesis gas in which the hydrogen/carbon monoxide (H2/CO) ratio is on the order of 3:1. Any unconverted methane is reacted with oxygen at the top of a secondary autothermal reformer, which contains a nickel catalyst in the lower region of the vessel.

In autothermal (or secondary) reformers, the oxidation of methane supplies the necessary energy and carries out either simultaneously or in advance of the reforming reaction (Brandmair, Find, & Lercher, 2003; Ehwald, Kürschner, Smejkal, & Lieske, 2003; Nagaoka, Jentys, & Lecher, 2003). The equilibrium of the methane steam reaction and the water-gas shift reaction determines the conditions for optimum hydrogen yields. The optimum conditions for hydrogen production require high temperature at the exit of the reforming reactor (800 to 900 °C; 1470 to 1650 °F); high excess of steam (molar steam-to-carbon ratio of 2.5 to 3); and relatively low pressures (below 450 psi). Most commercial plants employ supported nickel catalysts for the process.

One way of overcoming the thermodynamic limitation of steam reforming is to remove either hydrogen or carbon dioxide as it is produced, hence shifting the thermodynamic equilibrium toward the product side. The concept for sorption-enhanced methane steam reforming is based on in situ removal of carbon dioxide by a sorbent such as calcium oxide (CaO).

CaO+CO2CaCO3

si5_e

Sorption enhancement enables lower reaction temperatures, which may reduce catalyst coking and sintering, while enabling use of less expensive reactor wall materials. In addition, heat release by the exothermic carbonation reaction supplies most of the heat required by the endothermic reforming reactions. However, energy is required to regenerate the sorbent to its oxide form by the energy-intensive calcination reaction:

CaCO3CaO+CO2

si6_e

Use of a sorbent requires either that there be parallel reactors operated alternatively and out of phase in reforming and sorbent regeneration modes, or that sorbent be continuously transferred between the reformer/carbonator and regenerator/calciner (Balasubramanian, Ortiz, Kaytakoglu, & Harrison, 1999; Hufton, Mayorga, & Sircar, 1999).

Synthesis gas produced from natural gas (or coal or other carbonaceous feedstocks) is the building block in the synthesis of ammonia, methanol, Fischer-Tropsch (FT) fuels, hydrogen for hydrocracking at oil refineries, Oxo-alcohols, and other fine chemicals. The gas composition varies with the intended use of the syngas; ammonia production requires a molar H2/N2 ratio of 3, and for hydrogen production, the H2 content should be as high as possible. Because of the active shift reaction, both carbon monoxide and carbon dioxide are reactants in the methanol synthesis and in high temperature FTS. The syngas composition is therefore specified by a stoichiometric number [SN = (H2 – CO2)/(CO + CO2)] which should be close to 2. On the other hand, only carbon monoxide is a reactant for the low-temperature FTS, and the synthesis gas should have a H2/CO ratio close to 2. These different syngas compositions are achieved by using different types of reactor technology and by varying the amount of added steam and possibly oxygen or air, which is discussed later. The synthesis gas composition is also dependent on the feed gas composition and on the outlet temperature and pressure of the reforming reactor (Chadeesingh, 2011).

The higher molecular weight hydrocarbons that are also constituents of natural gas (Speight, 2007, 2014) are converted to methane in an adiabatic pre-reformer upstream of the steam reformer. In the pre-reformer, all higher hydrocarbons (C2 +) are converted into a mixture of methane, hydrogen, and carbon oxides:

CnHm+nH2OnCO+n+m/2H2

si7_e

3H2+COCH4+H2O

si8_e

CO+H2OH2+CO2

si9_e

The pre-reforming process utilizes an adiabatic fixed-bed reactor with highly active nickel catalysts. The reactions take place at temperatures of approximately 350 to 550 °C (650 to 1020 °F), which makes it possible to preheat the steam reformer feed to higher temperatures without getting problems with olefin formation from the higher hydrocarbons. Olefins are unwanted in the steam reformer feed because they generally cause coking of the catalyst pellets at high temperatures. Preheating of the steam reformer feed is of great advantage because the reformer unit can be scaled down to a minimum size (Aasberg-Petersen et al., 2001; Aasberg-Petersen, Christensen, Stub Nielsen, & Dybkjær, 2002; Hagh, 2004).

The reactions are catalyzed by pellets coated with nickel and are highly endothermic overall. Effective heat transport to the reactor tubes and further into the center of the catalytic fixed bed is therefore a very important aspect during design and operation of steam reformers. The reactions take place in several tubular fixed-bed reactors of low diameter-to-height ratio to ensure efficient heat transport in radial direction. The process conditions are typically 300 to 600 psi bar with inlet temperature of 300 to 650 °C (570 to 1200 °F) and outlet temperature of 700 to 950 °C (1290 to 1740 °F). There is often an approach to equilibrium of about 5 to 20 °C, which means that the outlet temperature is slightly higher than the equilibrium temperature calculated from the actual outlet composition (Rostrup-Nielsen, Christiansen, & Bak Hansen, 1988).

In a pre-reformer, whisker carbon can be formed either from methane or higher molecular weight hydrocarbons. The lower limit of the H2O/C ratio depends on a number of factors, including the feed gas composition, the operating temperature, and the choice of catalyst. In a pre-reformer operating at low H2O/C-ratio, the risk of carbon formation from methane is most pronounced in the reaction zone where the temperature is highest. Carbon formation from higher molecular weight hydrocarbons can take place only in the first part of the reactor with the highest concentrations of C2+compounds.

In addition, two water-gas shift (WGS) reactors are used downstream of the secondary reformer to adjust the H2/CO ratio, depending on the end use of the steam reformed products. The first of the two shift reactors utilize an iron-based catalyst that is heated to approximately 400 °C (750 °F), whereas the second shift reactor operates at approximately 200 °C (390 °F) and contains a copper-based catalyst.

The deposition of carbon on the catalyst (coking) can be an acute problem with the use of Ni-based catalysts in the primary reformer (Alstrup, 1988; Rostrup-Nielsen, 1984, 1993). The carbon deposition reactions occur in parallel with the reforming reactions and are undesirable, as they cause poisoning of the surface of the catalyst pellets. This leads to lower catalyst activity and the need for more frequent catalyst reloading. The coking reactions are the CO reduction, methane cracking, and Boudouard reaction, given by the respective equilibrium reactions:

CO+H2+H2OC+H2O

si10_e

CH4C+2H2

si11_e

2COC+CO2

si12_e

Thus, low steam excess can lead to critical conditions causing coke formation – equilibrium calculations of the coking reactions can be a useful tool for predicting the danger for catalyst poisoning but the reaction kinetics may nevertheless be so slow that coking is no concern. A complete analysis should therefore also involve kinetic calculations, which will be feedstock-dependent expressions for these reactions.

Traditionally, steam reformers have been run with a steam/carbon ratio of 2 to 4 to ensure low coking potential. It is desirable to reduce this ratio for methanol and FTS purposes because it will give great cost savings in form of smaller reformer units with higher methane conversion. Technical developments such as new noble metal catalysts and the use of pre-reformers are continually decreasing the feasible team/carbon ratio.

A successful technique is to use a steam/carbon ratio in the feed gas that does not allow the formation of carbon, but the process has a lower efficiency. Another approach is to use sulfur passivation, which led to the development of the SPARG process (Rostrup-Nielsen, 1984, 2006; Udengaard, Hansen, Hanson, & Stal, 1992). This technique utilizes the principle that the reaction leading to the deposition of carbon requires a larger number of adjacent nickel atoms on the catalyst surface than does steam reforming. A third method is to use Group VIII metals (such as platinum) that do not form carbides.

The most common reactor concept for steam reforming of natural gas is the fired steam reformer. Natural gas and the tail gas from the synthesis loop are burned in a firebox where several reactor tubes are placed in rows with a number of 40 to 400 tubes. The reactor tubes are about 33 to 40 feet with diameters of about 4 to 5 inches. The reactions for conversion of natural gas to synthesis gas take place over the catalytic beds in the reactor tubes. The burners can be located in different places: on the roof, on the floor, on leveled terraces on the walls, or on the walls (side-fired heating).

The top-fired steam reformer must be operated carefully because the tube wall temperature and heat flux show a peak in the upper part of the reformer. The bottom-fired reformers achieve a stable heat flux profile along the tube length, which causes high tube skin temperatures at the reactor outlet. The terrace wall-fired reformer is a modification of the bottom fired reformer and has some smaller problem with high metal temperatures. The side-fired reformer has the most effective design and is also the most flexible reformer, both in design and in operation (Dybkjær, 1995). This configuration has the highest total heat flux possible combined with the lowest heat flux where the tube skin temperature is at its highest. In this type of reformer, it is possible to combine a low steam-to-carbon ratio with a high outlet temperature. The most critical operation parameter is the maximum temperature difference over the tube wall, not the maximum heat flux (Aasberg-Petersen et al., 2001).

6.2.2 Autothermal reforming

Autothermal reforming (ATR) uses oxygen and carbon dioxide or steam in a reaction with methane to form synthesis. The reaction takes place in a single chamber where the methane is partially oxidized. The reaction is exothermic due to the oxidation. The main difference between autothermal reforming and steam-methane reforming is that steam-methane reforming does not use or require oxygen. The advantage of autothermal reforming is that H2/CO can be varied, which is particularly useful for producing certain second-generation biofuels such as dimethyl ether synthesis, which requires a 1:1 H2/CO ratio.

The process was developed in the 1950s and is used in commercial applications to provide syngas for ammonia and methanol synthesis. In the case of ammonia production, where high H2/CO ratios are needed, the process is operated at high steam/carbon ratios. In the case of methanol synthesis, the required H2/CO ratio is provided by manipulating the carbon dioxide recycle. In fact, development and optimization of this technology has led to cost-effective operation at very low steam/carbon feed ratios to produce CO-rich syngas, for example, which is preferred in FTS. These are the advantages of using the autothermal reactor: (1) the reactor is compact in design and therefore has a smaller footprint; (2) it has flexibility in its operation, with short startup periods and fast load changes; and (3) it is a soot-free operation. In addition, the reactor offers a better economic profile.

In the process, an organic feedstock (such as natural gas) and steam (there may also be low amounts of carbon dioxide in the feed) are mixed directly with oxygen and air in the reformer. The reformer itself comprises a refractory-lined vessel that contains the catalyst, together with an injector located at the top of the vessel. Partial oxidation reactions occur in the combustion zone of the reactor and the gaseous mixture then flows through a catalyst bed where the actual reforming reactions occur. Heat generated in the combustion zone from partial oxidation reactions is utilized in the reforming zone, so that in the ideal case, it is possible that the process can be in complete heat balance.

The autothermal reforming reactor consists of three zones: (1) the burner, where the feed streams are mixed in a turbulent diffusion flame; (2) the combustion zone, where partial oxidation reactions produce a mixture of carbon monoxide and hydrogen; and (3) the catalytic zone, where the gases leaving the combustion zone attain thermodynamic equilibrium. Key elements in the reactor are the burner and the catalyst bed – the burner provides mixing of the feed streams and the natural gas is converted in a turbulent diffusion flame:

CH4+3/2O2CO+2H2O

si13_e

When carbon dioxide is present in the feed, the H2/CO ratio produced is on the order of 1:1, but when the process employs steam, the H2/CO ratio produced is 2.5:1.

2CH4+O2+CO23H2+3CO+H2O

si14_e

4CH4+O2+2H2O10H2+4CO

si15_e

The risk of soot formation in an ATR reactor depends on a number of parameters, including feed gas composition, temperature, pressure, and especially burner design. Soot precursors may be formed in the combustion chamber during operation, so it is essential that the design of the burner, catalyst, and reactor is such that the precursors are destroyed by the catalyst bed to avoid soot formation.

Many observers consider the combination of adiabatic pre-reforming and autothermal reforming at low H2O/C ratios to be the preferred layout for production of synthesis gas for large gas-to-liquids plants.

6.2.3 Combined reforming

Combined reforming incorporates the combination of both steam reforming and autothermal reforming. In such a configuration, the hydrocarbon (e.g., natural gas) is first only partially converted, under mild conditions, to syngas in a relatively small steam reformer (Wang, Stagg-Williams, Noronha, Mattos, & Passos, 2004). The off-gases from the steam reformer are then sent to an oxygen-fired secondary reactor, such as an autothermal reactor, where the unreacted methane is converted to syngas by partial oxidation followed by steam reforming.

Another configuration requires that the hydrocarbon feed be split into two streams that are then fed in parallel, to a steam-reforming reactor and an autothermal reactor (gas-heated reforming). This process is an alternative to the fired steam reformer and has been commercially proven. There is also interest for the gas-heated steam reformer in relation to the FTS of hydrocarbons and methanol production.

6.2.4 Partial oxidation

Partial oxidation is the process in which the feed fuel, such as methane or a suitable hydrocarbonaceous fuel, reacts exothermically in the presence of a small amount of air (Vernon, Green, Cheetham, & Ashcroft, 1990; Rostrup-Nielsen, 2002; Zhu, Zhao, & Deng, 2004). Because incomplete combustion occurs, a gas containing hydrogen and carbon monoxide is produced. The hydrogen can be used to extend the lean limit of diesel, for instance, which indicates a higher efficiency of the fuel and lower pollutants emissions.

Partial oxidation (POX, POX) reactions occur when a substoichiometric hydrocarbon-air mixture is partially combusted in a reformer:

CnHm+2n+m/2O2nCO+m/2H2O

si16_e

Thus, for coal or any carbonaceous feedstocks the reaction can be represented simply as (understanding the in reality the reaction is extremely complex):

CHcoal+O2CO+H2

si17_e

In the process, the feedstock is partially burned in a simple pre-combustion chamber in the presence of a small amount of air and converted into carbon monoxide and hydrogen. Because partial oxidation is an exothermic reaction, some of the heat of combustion is released. The released energy is converted into heat, which brings the temperature of the gas to approximately 870 °C (1600 °F). The temperature of the gas needs to be lowered before entering the combustion engine. Otherwise, the density of the gas is too low to have a good volumetric efficiency. The resulting gas can be burned in a gas engine.

A thermal partial oxidation reactor is similar to the autothermal reactor with the main difference being no catalyst is used. In the process, the feedstock, which may include steam, is mixed directly with oxygen by an injector that is located near the top of the reaction vessel. Both partial oxidation as well as reforming reactions occur in the combustion zone below the burner.

The principal advantage of the partial oxidation process is the ability of the system to accommodate almost any carbonaceous feedstock, which can comprise very high molecular-weight organic constituents such as petroleum residual and petroleum coke (Gunardson & Abrardo, 1999; Speight, 2014). Additionally, because the emission of nitrogen oxides (NOx) and sulfur oxides (SOx) are minimal, partial oxidation does not leave a large environmental footprint.

On the other hand, very high temperatures, approximately 1300 °C (2370 °F), are required to achieve near complete reaction. This necessitates the consumption of some of the hydrogen and a greater than stoichiometric consumption of oxygen (i.e., oxygen-rich conditions).

Partial oxidation cannot be used for gasifying gasoline, diesel, methanol, or ethanol, because of the decrease in energy content of the fuel. However, the hydrogen-rich gas (hence, the preference for this type of process in the petroleum industry) that is produced by partial oxidation may be used to enrich other fuels. For the production of hydrogen, partial oxidation is often used in combination with steam reforming, using the heat of the partial oxidation for the endothermic steam reforming. However, given that steam reforming can be accomplished by using the energy from the exhaust gases coming out of the combustion engine, there is no need to partially oxidize the fuel first. Doing so would result in loss of the heating value of the fuel, and thus an overall energy loss for the process.

A possible means of improving the efficiency of syngas production is by use of the catalytic partial oxidation (CPOX, CPOX) technology, which has the potential to offer several advantages over steam reforming and thermal particle oxidation, particularly higher energy efficiency (Enger, Lødeng, & Holmen, 2008). The reaction is not endothermic – as is the case with steam reforming – but slightly exothermic. Furthermore, an H2/CO ratio of close to 2.0 (i.e., the ideal ratio for the Fischer-Tropsch and methanol synthesis) is produced by this technology, which can occur by either of two routes: direct or indirect.

The indirect route comprises total combustion of methane to carbon dioxide and water, followed by steam reforming and the water-gas shift reaction in which equilibrium conversions can be greater than 90% at ambient pressure. However, in order for an industrial process for this technology to be economically viable, an operating pressure in excess of 300 psi would be required. Unfortunately, at high pressures, equilibrium conversions are lower and, because of the exothermic combustion step, process control is more difficult and there is the potential for temperature runaways.

The direct route occurs by a mechanism involving only surface reaction on the catalyst:

CH4+0.5O2CO+2H2

si18_e

Compared with conventional synthesis gas production methods, the direct route would drastically reduce the amount of catalyst used, making it possible to use compact reactors.

6.2.5 Membrane reactors

An innovative technology for combining air separation and natural gas reforming processes is using membrane technology, which has the potential to reduce the cost of syngas generation and hydrocarbon products (Carolan, Chen, & Rynders, 2002; Khassin, 2005). The technology (oxygen transport membranes) can combine five unit operations currently in use: (1) oxygen separation, (2) oxygen compression, (3) partial oxidation, (4) steam methane reforming, and (5) heat exchange. The technology incorporates the use of catalytic components with the membrane to accelerate the reforming reactions.

A patented a two-step process for synthesis gas generation has been developed (Nataraj, Moore, & Russek, 2000) that can be utilized to generate synthesis gas from several feedstocks, including natural gas, associated gas (from crude oil production), light hydrocarbon gases from refineries, and medium-weight hydrocarbon fractions such as naphtha. The first stage comprises conventional steam reforming with partial conversion to synthesis gas and is followed by complete conversion in an ion transport ceramic membrane (ITM) reactor. This combination resolves any issues associated with steam reforming for feedstocks with hydrocarbons higher in molecular weight than methane, because the higher molecular weight hydrocarbons tend to crack and degrade both the catalyst and membrane.

By shifting the equilibrium in the steam reforming process through removal of hydrogen from the reaction zone, membrane reactors can also be used to increase the equilibrium-limited methane conversion. Using Pd-Ag alloy membrane reactors. methane conversion can reach as close to 100 % (Shu, Grandjean, & Kaliaguine, 1995).

6.3 Hydrogen production

Throughout the previous section there has been, of necessity, frequent reference to the production of hydrogen as an integral part of the production of carbon monoxide, because the two gases make up the mixture known as synthesis gas. Hydrogen is indeed an important commodity in the refining industry because of its use in hydrotreating processes, such as desulfurization, and in hydroconversion processes, such as hydrocracking. Part of the hydrogen is produced during reforming processes but that source, once sufficient, is now insufficient for the hydrogen needs of a modern refinery (Ancheyta & Speight, 2007; Speight, 2000; Speight, 2014; Speight & Ozum, 2002). In addition, optimum hydrogen purity at the reactor inlet extends catalyst life by maintaining desulphurization kinetics at lower operating temperatures and reducing carbon laydown. Typical purity increases resulting from hydrogen purification equipment and/or increased hydrogen sulfide removal, as well as tuning hydrogen circulation and purge rates, may extend catalyst life up to about 25%. Indeed, as hydrogen use has become more widespread in refineries, hydrogen production has moved from the status of a high-tech specialty operation to an integral feature of most refineries (Raissi, 2001; Vauk et al., 2008).

The gasification of residue and coke to produce hydrogen and/or power may increase in use in refineries over the next two decades (Speight, 2011b), but several other processes are available for the production of the additi onal hydrogen that is necessary for the various heavy feedstock hydroprocessing sequences (Speight, 2014). This section presents a general description of these processes. These gasification processes, which are often referred to the garbage disposal units of the refinery, have not been described earlier.

6.3.1 Heavy residue gasification and combined cycle power generation

Heavy residues are gasified and the produced gas is purified to fuel gas that is free of contaminants (Gross & Wolff, 2000). As an example, solvent deasphalter residuum (deasphalter bottoms) is gasified by partial oxidation method under pressure of about 570 psi and at a temperature between 1300 and 1500 °C (2370 °F and 2730 °F). The high temperature generates gas stream flows into a waste heat boiler, in which the hot gas is cooled and high pressure saturated steam is generated. The gas from the waste heat boiler is then heat exchanged with the fuel gas and flows to the carbon scrubber, where unreacted carbon particles are removed from the generated gas by water scrubbing.

The gas from the carbon scrubber is further cooled by the fuel gas and boiler feed water and led into the sulfur compound removal section, where hydrogen sulfide (H2S) and carbonyl sulfide (COS) are removed from the gas to obtain clean fuel gas. This clean fuel gas is heated with the hot gas generated in the gasifier and finally supplied to the gas turbine at a temperature of 250 to 300 °C (480 to 570 °F).

In order to decrease the nitrogen oxide (NOx) content in the flue gas, two methods can be applied. The first method is the injection of water into the gas turbine combustor. The second method is to selectively reduce the nitrogen oxide content by injecting ammonia gas in the presence of de-NOx catalyst that is packed in a proper position of the heat recovery steam generator. The latter is more effective that the former to lower the nitrogen oxide emissions to the air.

6.3.2 Hybrid gasification process

In the hybrid gasification process, a coal/residual oil slurry is injected into the gasifier where it is pyrolyzed in the upper part of the reactor to produce gas and chars. The chars produced are then partially oxidized to ash. The ash is removed continuously from the bottom of the reactor.

In this process, coal and vacuum residue are mixed together into slurry to produce clean fuel gas. The slurry fed into the pressurized gasifier is thermally cracked at a temperature of 850 to 950 °C (1560 to 1740 °F) and is converted into gas, tar, and char. The mixture of oxygen and steam in the lower zone of the gasifier converts the char to gaseous products. The gas leaving the gasifier is quenched to a temperature of 450 °C (840 °F) in the fluidized-bed heat exchanger, and is then scrubbed to remove tar, dust, and steam at around 200 °C (390 °F).

Ash is discharged from the gasifier and indirectly cooled with steam and then discharged into the ash hopper. It is burned with an incinerator to produce process steam. Coke deposited on the silica sand is removed in the incinerator.

6.3.3 Hydrocarbon gasification

The gasification of hydrocarbons to produce hydrogen is a continuous, noncatalytic process that involves partial oxidation of the hydrocarbon and one of several processes that are used for gasification of carbonaceous fuels to gaseous products (Breault, 2010).

In the process, air or oxygen, with steam or carbon dioxide, is used as the oxidant at 1095 to 1480 °C (2000 to 2700 °F). Any carbon produced (2 to 3% w/w of the feedstock) during the process is removed as a slurry in a carbon separator and pelletized for use either as a fuel or as raw material for carbon-based products.

6.3.4 Hypro process

Due to its abundance and high H/C ratio (highest among all hydrocarbons), methane is an obvious source for hydrogen. The steam reforming of methane represents the current trend for hydrogen production (Hypro process). Other popular methods of hydrogen production include autothermal reforming and partial oxidation. However, if hydrogen is the desire product, all these processes involve the formation of large amounts of unwanted carbon monoxide and carbon dioxide (COx) as a by-product.

Hydrogen production routes, which do not require complex COx removal procedures, are therefore desired for production of high-purity hydrogen. Thus, there is much interest in the catalytic decomposition of natural gas, whose major constituent is methane, for production of hydrogen. Given that only hydrogen and carbon are formed in the decomposition process, the separation of products is not an issue. The other main advantage is the simplicity of the methane decomposition process as compared to conventional methods. For example, the high- and low-temperature water-gas shift reactions and carbon dioxide removal step (involved in the conventional methods) are completely eliminated. Catalyst regeneration is extremely important for the practical application of the clean hydrogen production process.

The hypro process is a continuous catalytic method for hydrogen manufacture from natural gas or from refinery effluent gases, especially the decomposition of methane to hydrogen and carbon (Choudhary & Goodman, 2006; Choudhary, Sivadinarayana, & Goodman, 2003):

CH4C+2H2

si19_e

Hydrogen is recovered by phase separation to yield hydrogen of about 93% purity, and the principal contaminant is methane.

6.3.5 Pyrolysis processes

There has been recent interest in the use of pyrolysis processes to produce hydrogen. Specifically, the interest has focused on the pyrolysis of methane (natural gas) and hydrogen sulfide.

Natural gas is readily available and offers a relatively rich stream of methane with lower amounts of ethane, propane, and butane also present. The thermocatalytic decompositon of natural gas hydrocarbons (c.f., hypro process) offers an alternate method for the production of hydrogen (Dahl & Weimer, 2001; Uemura, Ohe, Ohzuno, & Hatate, 1999; Weimer et al., 2000):

CnHmnC+m/2H2

si20_e

The production of hydrogen by direct decomposition of hydrogen sulfide has also been proposed (Clark & Wassink, 1990; Donini, 1996; Luinstra, 1996; Zaman & Chakma, 1995). Hydrogen sulfide decomposition is a highly endothermic process, and equilibrium yields are poor (Clark, Dowling, Hyne, & Moon, 1995). At temperatures less than 1500 °C (2730 °F), the thermodynamic equilibrium is unfavorable toward hydrogen formation. However, in the presence of catalysts such as platinum-cobalt at 1000 °C (1830 °F), disulfides of molybdenum (Mo) or tungsten (W) at 800 °C (1470 °F) (Kotera, Todo, & Fukuda, 1976), or other transition metal sulfides supported on alumina at 500 to 800 °C (930 to 1470 °F), decomposition of hydrogen sulfide proceeds. In the temperature range of about 800 to 1500 °C (1470 to 2730 °F), thermolysis of hydrogen sulfide can be treated simply:

H2SH2+1/xSxΔH298K=+34,300Btu/lb

si21_e

where x = 2. Outside this temperature range, multiple equilibria may be present, depending on temperature, pressure, and relative abundance of hydrogen and sulfur (Clark & Wassink, 1990).

In addition, the steam-iron process is an established process, which was used for the production of hydrogen from cokes at the beginning of the twentieth century. However, the process could not compete with the later-developed steam reforming of methane, and so the process fell into disuse. The renewed interest in the development of the steam-iron process is mainly focused on the use of renewable energy sources, like biomass. In this thesis, the production of hydrogen by the steam-iron process from pyrolysis oil is studied. Pyrolysis oil, obtained from the pyrolysis of biomass, is used to facilitate transportation and to simplify gasification and combustion processes, before being processed to hydrogen. The benefit of the steam-iron process compared to other thermo-chemical routes of biomass is that hydrogen can be produced in a two-step redox cycle, without the need of any purification steps (such as pressure-swing adsorption) (Bleeker, 2009; Bleeker, Kersten, & Veringa, 2007).

6.3.6 Shell gasification process

The Shell gasification process (partial oxidation process) is a flexible process for generating syngas, principally hydrogen and carbon monoxide, for the ultimate production of high-purity, high-pressure hydrogen, ammonia, methanol, fuel gas, and town gas, or for reducing gas by reaction of gaseous or liquid hydrocarbons with oxygen, air, or oxygen-enriched air. Traditionally, petroleum residues have been sold as marine bunker fuel or used on-site as furnace fuel. However, with changing legislation, refineries are under pressure to reduce both their emissions and the sulfur content of their products. In addition, the market for fuel oil is shrinking. The Shell gasification process can be combined with other upgrading and treating technologies to convert a wide range of low-value residue into syngas.

The most important step in converting heavy residue into industrial gas is the partial oxidation of the oil using oxygen with the addition of steam. The gasification process takes place in an empty, refractory-lined reactor at temperatures of about 1400 °C (2550 °F) and pressures between 29 and 1140 psi. The chemical reactions in the gasification reactor proceed without catalyst to produce gas that contains carbon amounting to some 0.5 to 2% by weight, based on the feedstock. The carbon is removed from the gas with water, extracted in most cases with feed oil from the water, and returned to the feed oil. The high reformed gas temperature is utilized in a waste heat boiler for generating steam. The steam is generated at 850 to 1565 psi. Some of this steam is used as process steam and for oxygen and oil preheating. The surplus steam is used for energy production and heating purposes.

6.3.7 Steam-naphtha reforming

Liquid feedstocks, either liquefied petroleum gas or naphtha, can also provide backup feed for the steam-methane reformer, if there is a risk of natural gas curtailments (Breault, 2010; Rostrup-Nielsen & Christiansen, 2011). The feed-handling system needs to include a surge drum, feed pump, a vaporizer (usually steam-heated) followed by further heating before desulfurization. The sulfur in liquid feedstocks occurs as mercaptans, thiophene derivatives, or higher boiling compounds. These compounds are stable and will not be removed by zinc oxide, therefore a hydrogenation unit will be required. In addition, as with refinery gas, olefins must also be hydrogenated if they are present.

Thus, steam-naphtha reforming is a continuous process for the production of hydrogen from liquid hydrocarbons. In fact, it is similar to steam-methane reforming, which is one of several possible processes for the production of hydrogen from low-boiling hydrocarbons other than ethane (Brandmair et al., 2003; Find, Nagaoka, & Lercher, 2003; Muradov, 1998; Murata, Ushijima, & Fujita, 1997). A variety of naphtha-types in the gasoline boiling range may be employed, including feeds containing up to 35% aromatics. Thus, following pre-treatment to remove sulfur compounds, the feedstock is mixed with steam and taken to the reforming furnace (675 to 815 °C, 1250 to 1500 °F, 300 psi, where hydrogen is produced.

6.3.8 Texaco gasification (partial oxidation) process

The Texaco gasification (partial oxidation) process is a partial oxidation gasification process for generating synthesis gas (Breault, 2010). The characteristic of the process is to inject feedstock together with carbon dioxide, steam, or water into the gasifier. Therefore, solvent deasphalted residue or petroleum coke rejected from any coking method can be used as feedstock for this gasification process. The produced gas from this gasification process can be used for the production of high-purity, high-pressurized hydrogen, ammonia, and methanol. The heat recovered from the high-temperature gas is used for the generation of steam in the waste heat boiler. Alternatively the less expensive quench-type configuration is preferred when high-pressure steam is not needed or when a high degree of shift is needed in the downstream carbon monoxide converter.

In the process, the feedstock, together with the feedstock carbon slurry recovered in the carbon recovery section, is pressurized to a given pressure, mixed with high-pressure steam, and then blown into the gas generator through the burner together with oxygen.

The gasification reaction is a partial oxidation of hydrocarbons to carbon monoxide and hydrogen:

CxH2y+x/2O2xCO+yH2

si22_e

CxH2y+xH2OxCO+x+yH2

si23_e

The gasification reaction is instantly completed, thus producing gas mainly consisting of H2 and CO (H2 + CO = > 90%). The high-temperature gas leaving the reaction chamber of the gas generator enters the quenching chamber linked to the bottom of the gas generator and is quenched to 200 to 260 °C (390 to 500 °F) with water.

6.3.9 Recovery from fuel gas

Recovering of hydrogen from refinery fuel gas can help refineries satisfy high hydrogen demand. Cryogenic separation is typically viewed as being the most thermodynamically efficient separation technology. The basic configuration for hydrogen recovery from refinery gases involves a two-stage partial condensation process, with post purification via pressure swing adsorption (Dragomir et al., 2010). The major steps in this process involve first compressing and pre-treating the crude refinery gas stream before chilling to an intermediate temperature (–60 to –120°F). This partially condensed stream is then separated in a flash-drum after which the liquid stream is expanded through a Joule-Thompson valve to generate refrigeration and then is fed to the wash column. Optionally, the wash column can be replaced by a simple flash drum.

A crude liquefied petroleum gas (LPG) stream is collected at the bottom of the column, and a methane-rich vapor is obtained at the top. The methane-rich vapor is sent to compression and then to fuel. The vapor from the flash drum is further cooled in a second heat exchanger before being fed to a second flash drum where it produces a hydrogen-rich stream and a methane-rich liquid. The liquid is expanded in a Joule-Thomson valve to generate refrigeration, and then is sent for further cooling. Next, the hydrogen-rich gas is sent to the pressure swing adsorption unit for further purification. The tail gas from this unit is compressed and returned to fuel, together with the methane-rich gas.

6.4 Gasification products: composition and quality

The composition of the products from gasification processes is varied insofar as the gas composition varies with the type of feedstock and the gasification system employed (Chapters 1, 2, and 10). Furthermore, the quality of gaseous product(s) must be improved by removal of any pollutants such as particulate matter and sulfur compounds before further use, particularly when the intended use is a water-gas shift or methanation (Speight, 2007, 2008, 2013a, 2013b).

Generally, products from gasification processes can range from (1) high-purity hydrogen, (2) high-purity carbon monoxide, (3) high-purity carbon dioxide, and (4) a range of H2/CO mixtures (Chapter 10). In fact, the H2/CO ratio can be selected at will and the appropriate process scheme chosen, in part, by the product composition required. At one end of the scale, (i.e., if hydrogen is the desired product), the H2/CO ratio can approach infinity by converting (shifting) all of the carbon monoxide to CO2. By contrast, on the other end, the ratio cannot be adjusted to zero because hydrogen and water are always produced.

Low-Btu gas (low-heat content gas) is the product when the oxygen is not separated from the air and, as a result, the gas product invariably has a low-heat content (150 to 300 Btu/ft3). In medium-Btu gas (medium-heat content gas), the heating value is in the range 300 to 550 Btu/ft3 and the composition is much like that of low-heat content gas, except that there is virtually no nitrogen and the H2/CO ratio varies from 2:3 to approximately 3:1 and the increased heating value correlates with higher methane and hydrogen contents as well as with lower carbon dioxide content. High-Btu gas (high-heat content gas) is essentially pure methane and often referred to as synthetic natural gas or substitute natural gas. However, to qualify as substitute natural gas, a product must contain at least 95% methane; the energy content of synthetic natural gas is 980 to 1080 Btu/ft3. The commonly accepted approach to the synthesis of high-heat content gas is the catalytic reaction of hydrogen and carbon monoxide.

Hydrogen is also produced during gasification of carbonaceous feedstocks. Although several gasifier types exist (Chapter 2), entrained-flow gasifiers are considered most appropriate for producing both hydrogen and electricity from coal. This is because they operate at temperatures high enough (approximately 1500 °C, 2730 °F) to enable high carbon conversion and prevent downstream fouling from tars and other residuals.

There is also a series of products that are called by older (even archaic) names that evolved from older coal gasification technologies and warrant mention: (1) producer gas, (2) water gas, (3) town gas, and (4) synthetic natural gas. These products are typically low-to-medium Btu gases (Chapter 10).

6.4.1 Purification

The processes that have been developed for gas cleaning (Mokhatab, Poe, & Speight, 2006; Speight, 2007, 2008) vary from a simple once-through wash operation to complex multistep systems with options for recycle of the gases (Mokhatab et al., 2006). In some cases, process complexities arise because of the need for recovery of the materials used to remove the contaminants or even recovery of the contaminants in the original, or altered, form.

In more general terms, gas cleaning is divided into removal of particulate impurities and removal of gaseous impurities. For the purposes of this chapter, the latter operation includes the removal of hydrogen sulfide, carbon dioxide, sulfur dioxide, and products that are not related to synthesis gas and hydrogen production. However, there is also need for subdivision of these two categories as dictated by needs and process capabilities: (1) coarse cleaning whereby substantial amounts of unwanted impurities are removed in the simplest, most convenient, manner; (2) fine cleaning for the removal of residual impurities to a degree sufficient for the majority of normal chemical plant operations, such as catalysis or preparation of normal commercial products, or cleaning to a degree sufficient to discharge an effluent gas to atmosphere through a chimney; and (3) ultra-fine cleaning where the extra step (as well as the extra expense) is justified by the nature of the subsequent operations or the need to produce a particularly pure product.

Contrary to the general belief of some scientists and engineers, all gas-cleaning systems are not alike, and having a good understanding of the type of gaseous effluents from coal-based processes is necessary to implementing the appropriate solution. The design of a gas-cleaning system must always take into account the operation of the upstream installations, because every process will have a specific set of requirements. In some cases, the application of a dry dusting removal unit may not be possible and thus requires a special process design of the wet gas-cleaning plant. Thus, the gas-cleaning process must always be of optimal design – one for both the upstream and downstream processes.

Gas processing, although generally simple in chemical and/or physical principles, is often confusing because of the frequent changes in terminology and, often, lack of cross-referencing (Mokhatab et al., 2006; Speight, 2007,, 2008, 2013a, 2014). Although gas processing employs different process types, there is always overlap between the various concepts. And, with the variety of possible constituents and process operating conditions, a universal purification system cannot be specified for economic application in all cases.

Nevertheless, the first step in gas cleaning is usually a device to remove large particles of carryover (entrained) material coal and other solid materials (Mokhatab et al., 2006; Speight, 2007, 2008). This is followed by cooling, quenching, or washing to condense tars and oils and to remove dust and water-soluble materials from the gas stream. Water washing is desirable for simplicity in gas cleaning; however, the purification of this water is not simple.

Clean-up steps and their sequence can be affected by the type of gas produced and its end use (Mokhatab et al., 2006; Speight, 2007, 2008). The minimum requirement in this respect would be the application of low-heat value (low-Btu) gas produced from low-sulfur anthracite coal as a fuel gas. The gas may pass directly from the gasifier to the burners and, in this case, the burners are the clean-up system. Many variations on this theme are possible; also, the order of the cleanup stages may be varied.

The selection of a particular process-type for gas cleaning is not simple. Several factors have to be considered, not the least of which is the constitution of the gas stream that requires treatment. Indeed, process selectivity indicates the preference with which the process will remove one acid gas component relative to (or in preference to) another. For example, some processes remove both hydrogen sulfide and carbon dioxide; other processes are designed to remove hydrogen sulfide only (Mokhatab et al., 2006; Speight, 2007, 2014).

Gas cleaning by absorption by a liquid or adsorption by use of a solid sorbent is one of the most widely applied operations in the chemical and process industries (Mokhatab et al., 2006; Speight, 2007). Some processes have the potential for sorbent regeneration, but in a few cases, the process is applied in a nonregenerative manner. The interaction between sorbate and sorbent may either be physical in nature or consist of physical sorption followed by chemical reaction. Other gas stream treatments use the principle of chemical conversion of the contaminants with the production of “harmless” (noncontaminant) products or to substances that can be removed much more readily than the impurities from which they are derived (Mokhatab et al., 2006; Speight, 2007, 2008).

Any gases, such as hydrogen sulfide and/or carbon dioxide, that are the products of coal processing can be removed by application of an amine washing procedure.

2RNH2+H2SRNH32S

si24_e

RNH32S+H2S2RNH3HS

si25_e

2RNH2+CO2+H2ORNH32CO3

si26_e

RNH32CO3+H2O2RNH3HCO3

si27_e

There are also solvent extraction methods for producing low-sulfur and low-mineral matter coal, but hydrotreatment of the coal extract is also required. In these methods, the organic material is extracted from the inorganic material in coal. A study has indicated that solvent-refined coal will probably not penetrate the power generation industry on a large scale for several years to come.

In addition to hydrogen sulfide and carbon dioxide, gas streams may contain other contaminants such as sulfur dioxide, mercaptans, and carbonyl sulfide. The presence of these impurities may eliminate some of the sweetening processes because some processes will remove large amounts of acid gas but not to a sufficiently low concentration. On the other hand, there are those processes that are not designed to remove (or are incapable of removing) large amounts of acid gases, yet they are capable of removing the acid gas impurities to very low levels when the acid gases are there in low-to-medium concentrations in the gas stream.

Many different methods have been developed for carbon dioxide and hydrogen sulfide removal, some of which are briefly discussed here. Concentrates of hydrogen sulfide obtained as by-products of gas desulfurization are often converted by partial oxidation to elemental sulfur (Claus process) (Mokhatab et al., 2006; Speight, 2007,, 2013a, 2014).

Pressure swing adsorption units use beds of solid adsorbent to separate impurities from hydrogen streams leading to high-purity, high-pressure hydrogen and a low-pressure tail gas stream containing the impurities and some of the hydrogen. The beds are then regenerated by depressuring and purging. Part of the hydrogen (up to 20% v/v) may be lost in the tail gas.

Pressure swing adsorption is generally the purification method of choice for steam reforming units because of its production of high-purity hydrogen. It is also used for purification of refinery off-gases, where it competes with membrane systems.

Many hydrogen plants that formerly used a wet scrubbing process for hydrogen purification are now using the pressure swing adsorption (PSA) for purification (Speight, 2007, 2014). The pressure swing adsorption process is a cyclic process that uses beds of solid adsorbent to remove impurities from the gas and generally produces higher-purity hydrogen (99.9% v/v purity compared to less than 97% v/v purity). The purified hydrogen passes through the adsorbent beds with only a tiny fraction absorbed, and the beds are regenerated by depressurization followed by purging at low pressure.

When the beds are depressurized, a waste gas (or tail gas) stream is produced and consists of the impurities from the feed (carbon monoxide, carbon dioxide, methane, and nitrogen) plus some hydrogen. This stream is burned in the reformer as fuel and reformer operating conditions in a pressure swing adsorption plant are set so that the tail gas provides no more than about 85% v/v of the reformer fuel. This gives good burner control because the tail gas is more difficult to burn than regular fuel gas and the high content of carbon monoxide can interfere with the stability of the flame. As the reformer operating temperature is increased, the reforming equilibrium shifts, resulting in more hydrogen and less methane in the reformer outlet and hence less methane in the tail gas.

Membrane systems separate gases by taking advantage of the difference in rates of diffusion through membranes (Brüschke, 1995, 2003). Gases that diffuse faster (including hydrogen) become the permeate stream and are available at low pressure, whereas the slower-diffusing gases become the nonpermeate and leave the unit at a pressure close to the pressure of the feedstock at the entry point. Membrane systems contain no moving parts or switch valves and have potentially very high reliability. The major threat is from components in the gas (such as aromatics) that attack the membranes, or from liquids, which plug them.

Membranes are fabricated in relatively small modules; for larger capacity, more modules are added. Cost is therefore virtually linear with capacity, making them more competitive at lower capacities. The design of membrane systems involves a tradeoff between pressure drop (or diffusion rate) and surface area, as well as between product purity and recovery. As the surface area is increased, the recovery of fast components increases; however, more of the slow components are recovered, which lowers the purity.

Cryogenic separation units operate by cooling the gas and condensing some, or all, of the constituents for the gas stream. Depending on the product purity required, separation may involve flashing or distillation. Cryogenic units offer the advantage of being able to separate a variety of products from a single-feed stream. One specific example is the separation of light olefins from a hydrogen stream.

Hydrogen recovery is in the range of 95% v/v, with purity above 98% v/v obtainable. In addition to the general description of the purification processes presented earlier, four of the major process techniques for achieving this level of purity are:

(i) Cryogenics plus methanation, which utilizes a cryogenic process whereby carbon monoxide is liquefied in a number of steps until hydrogen with a purity of on the order of 98% is produced. The condensed carbon monoxide, which would contain methane, is then distilled to produce pure carbon monoxide and a mixture of carbon monoxide and methane. The hydrogen stream is taken to a shift converter where the remaining carbon monoxide is converted to carbon dioxide and hydrogen. The carbon dioxide is removed and any further carbon monoxide or carbon dioxide can be removed by methanation. The resulting hydrogen stream typically has purity on the order of 99.7% v/v.

(ii) Cryogenics plus pressure swing adsorption (PSA), which utilizes the similar sequential liquefaction of carbon monoxide until hydrogen having 98% purity is obtained. Again, the carbon monoxide stream can be further distilled to remove methane until it is essentially pure. The hydrogen stream is then allowed to go through multiple pressure swing adsorption cycles until the hydrogen purity is as high as 99.999% v/v.

(iii) Methane-wash cryogenic process utilizes the principle of carbon monoxide absorption in a liquid methane stream so that the hydrogen stream produced contains only low levels (on the order of parts per million) of carbon monoxide but about 5 to 8% v/v methane. Hence, the hydrogen stream may have purity on the order of only 95% v/v. However, the liquid carbon monoxide/methane stream can be distilled to produce a pure carbon monoxide stream and a carbon monoxide/methane stream, which can be used as fuel.

(iv) COsorb process utilizes copper ions (cuprous aluminum chloride, CuAlCl4) in toluene to form a chemical complex with the carbon monoxide to separate it from hydrogen, nitrogen, carbon dioxide, and methane. This process can capture about 96% of the available carbon monoxide to produce a stream of greater than 99% purity. On the other hand, water, hydrogen sulfide, and other trace chemicals can poison the copper catalyst and must be removed prior to the reactor.

Although the efficiency of cryogenic separation decreases with a content of low carbon monoxide in the feed, the COsorb process is able to process gas streams with low carbon monoxide content much more efficiently.

6.4.2 Oil-water separation

The typical oil-water separation process occurs in a device designed to separate gross amounts of oil and suspended solids from the effluents of petroleum and gas processing. The most common type of separator is the API separator, which is a gravity separation device designed by using the specific gravity difference between the oil and water (depending on the pressure, the gas remains in the volatile state) while the oil and water separate from stream as liquids (Mokhatab et al., 2006; Speight, 2007). Based on that design criterion, any suspended solids settle to the bottom of the separator as a sediment layer, the oil will rise to the top of the separator, and the wastewater will be the middle layer between the oil on the top and the solids on the bottom.

Typically, the oil layer is skimmed off and subsequently reprocessed or disposed of, and the bottom sediment layer is removed by a chain and flight scraper (or similar device) and a sludge pump. The water layer is sent to further treatment consisting usually of a dissolved air flotation unit for further removal of any residual oil and then to some type of biological treatment unit for removal of undesirable dissolved chemical compounds.

Parallel plate separators are similar to API separators but include tilted parallel plate assemblies, and the underside of each parallel plate provides more surface for suspended oil droplets to coalesce into larger globules. Any sediment slides down the topside of each parallel plate. Such separators still depend on the specific gravity between the suspended oil and the water. However, the parallel plates enhance the degree of oil-water separation. The result is that a parallel plate separator requires significantly less space than a conventional API separator to achieve the same degree of separation.

6.5 Advantages and limitations

In the early days of the petroleum industry, the delayed coking unit was considered as the garbage can of the refinery in which any high-boiling petroleum-based feedstock (typically not much good for anything else) could be converted to distillates. For some time, gasifiers were considered in the same light, but that was not always the case. Advantages were obvious and disadvantages were not always obvious but they were not insurmountable. This section relates to the advantages and limitations of the gasification process that should be taken into consideration for efficient operation of a gasification plant – all relate to the production of the two major products (carbon monoxide and hydrogen) being given consideration here.

Gasification enables the capture – in an environmentally beneficial manner – of the value present in a variety of low-grade carbonaceous, wastes, or biomass. Without gasification, these materials would have to be disposed of by an alternate route that could potentially damage the environment and, equally important, ignore or discard a valuable source of energy. Although traditional feedstocks included coal and petroleum coke in large-scale industrial plants, there is an increasing use of municipal solid waste, industrial waste, and biomass in smaller-scale plants, converting that material to energy.

In fact, the increasing costs of conventional waste management and disposal options, and the desire, in most developed countries, to divert an increasing proportion of mixed organic waste materials from landfill disposal, for environmental reasons, will render the investment in energy from waste projects increasingly attractive. Indeed, gasification also offers more scope for recovering products from waste than incineration. When waste is burnt in an incinerator, the only practical product is energy, whereas the gases, oils, and solid char from pyrolysis and gasification not only can be used as a fuel but they may also be purified and used as a feedstock for petrochemicals and other applications. Rather than producing only ash, many processes also produce a stable granulate, which can be more easily and safely utilized. In addition, some processes are targeted at producing specific recyclables such as metal alloys and carbon black. From waste gasification, in particular, it is feasible to produce hydrogen, which many see as an increasingly valuable resource.

Most new projects for the recovery of energy from various carbonaceous feedstocks (including wastes such as municipal waste materials) will involve the installation of new purpose-designed incineration plant with heat recovery and power generation. However, advanced thermal processes for municipal solid waste that are based on pyrolysis or gasification processes are also being introduced. These processes offer significant environmental and other attractions and will likely have an increasing role to play, but the rate of increase of use is difficult to predict.

Despite the benefits of using coal as a gasification feedstock, there are several environmental challenges, including significant air quality, climate change, and mining impacts. However, coal gasification technologies have been demonstrated that provide order-of-magnitude reductions in criteria pollutant emissions and, when coupled with carbon capture and sequestration (CCS), the potential for significant reductions in carbon dioxide emissions. Therefore, although coal is a finite nonrenewable resource, coal-derived hydrogen with carbon capture and storage can increase domestic energy independence, provide near-term carbon dioxide and criteria pollutant reduction benefits, and facilitate the transition to a more sustainable hydrogen-based transportation system. Carbon capture and storage is one of the critical enabling technologies that could lead to coal-based hydrogen production for use as a transportation fuel. However, there are other risks to the environment that need to be addressed.

Although not a limiting factor of the process, many forms of biomass contain a high percentage of moisture (along with carbohydrates and sugars) and mineral constituents – both of which can influence the economics and viability of a gasification process. The presence of high levels of moisture in the biomass reduces the temperature inside the gasifier, which then reduces the efficiency of the gasifier. Therefore, many biomass gasification technologies require that the biomass be dried to reduce the moisture content prior to feeding into the gasifier. In addition, biomass can come in a range of sizes. In many biomass gasification systems, the biomass must be processed to a uniform size or shape to feed into the gasifier at a consistent rate and to ensure that as much of the biomass is gasified as possible.

Furthermore, the presence of mineral matter in the coal-biomass feedstock is not appropriate for fluidized-bed gasification. The low melting point of ash present in woody biomass leads to agglomeration, which causes defluidization of the ash as well as sintering, deposition, and corrosion of the gasifier construction metal bed (Vélez et al., 2009). Biomass containing alkali oxides and salts with the propensity of produce yield higher than 5% w/w ash causes clinkering/slagging problems (McKendry, 2002). Thus, it is imperative to be aware of the melting of biomass ash, its chemistry within the gasification bed (no bed, silica/sand, or calcium bed), and the fate of alkali metals when using fluidized-bed gasifiers.

Furthermore, the disposal of municipal and industrial wastes has become an important problem because the traditional means of disposal – landfill – has become environmentally much less acceptable than previously. New, much stricter regulations of these disposal methods will make the economics of waste processing for resource recovery much more favorable. One method of processing waste streams is to convert the energy value of the combustible waste into a fuel. One type of fuel attainable from wastes is a low-heating value gas, usually 100 to 150 Btu/ft3, which can be used to generate process steam or to generate electricity. Co-processing such waste with coal is also an option (Speight, 2008).

One of the major disadvantages of gasification plants in general, irrespective of the feedstock, is the environmental impact that has drawn increasing concern. Attention is not only focused on controlling pollutants such as sulfur dioxide (SO2), oxides of nitrogen (NOx), and particulates (PM) but also for controlling the emission of carbon dioxide (CO2). There is an increasing need to reduce the emissions of carbon dioxide to the atmosphere to alleviate the global warming effect. It induces significant challenges to generate electricity efficiently together with near-zero carbon dioxide emissions.

In the process, carbon dioxide, hydrogen, and other coal by-products are captured so they can be used for useful purposes. Evolving technologies are also making coal at existing plants cleaner – refined coal technologies remove many of the impurities contained in existing coal. New techniques are helping remove mercury and harmful gases while unlocking more energy potential.

In co-gasification of coal with other feedstocks or with a mixture of feedstocks (coal may be excluded), the technology varies and is usually site specific with high dependence on the feedstock (Brar, Singh, Wang, & Kumar, 2012). At the largest scale, the plant may include the well-proven fixed-bed and entrained-flow gasification processes. At smaller scales, emphasis is placed on technologies that appear closest to commercial operation. Pyrolysis and other advanced thermal conversion processes are included where power generation is practical, using the on-site feedstock produced (Chapter 1).

A major advantage of the gasification process is that it lends itself to the installation of a gasification refinery that would have, as the centerpiece, gasification technology or at least as a section of a conventional petroleum refinery (Speight, 2011b). The refinery would produce syngas (from the carbonaceous feedstock) from which liquid fuels would be manufactured using the FTS technology.

The manufacture of gas mixtures of carbon monoxide and hydrogen has been an important part of chemical technology for approximately 100 years. Originally, such mixtures were obtained by the reaction of steam with incandescent coke and were known as water gas. Eventually, steam reforming processes, in which steam is reacted with natural gas (methane) or petroleum naphtha over a nickel catalyst, found wide application for the production of synthesis gas.

As petroleum supplies decrease, the desirability of producing gas from other carbonaceous feedstocks will increase, especially in those areas where natural gas is in short supply. It is also anticipated that costs of natural gas will increase, allowing coal gasification to compete as an economically viable process. Research in progress on a laboratory and pilot-plant scale should lead to the invention of new process technology by the end of the 21st Century, thus accelerating the industrial use of coal gasification.

The conversion of the gaseous products of gasification processes to synthesis gas, a mixture of hydrogen (H2) and carbon monoxide (CO), in a ratio appropriate to the application, needs additional steps, after purification. The product gases – carbon monoxide, carbon dioxide, hydrogen, methane, and nitrogen – can be used as fuels or as raw materials for chemical or fertilizer manufacture.

Finally, the gas from any gasification process is inherently toxic because of essential components such as carbon monoxide and unwanted components. However, this inherent toxicity is not the reason for gas cleaning because the gas should never be released to the atmosphere directly.

References

Aasberg-Petersen K, Bak Hansen J-H, Christiansen TS, Dybkjær I, Seier Christensen P, Stub Nielsen C, et al. Technologies for Large-Scale Gas Conversion. Applied Catalysis A: General. 2001;221:379–387.

Aasberg-Petersen K, Christensen TS, Stub Nielsen C, Dybkjær I. Recent Developments in Autothermal Reforming and Pre-reforming for Synthesis Gas Production in GTL Applications. Preprints of Papers-American Chemical Society, Division of Fuel Chemistry. 2002;47(1):96–97.

Alstrup I. A New Model Explaining Carbon Filament Growth on Nickel, Iron, and Ni-Cu Alloy Catalysts. Journal of Catalysis. 1988;109:241–251.

Ancheyta J, Speight JG. Hydroprocessing of Heavy Oils and Residua. Taylor & Francis Group, Boca Raton, Florida: CRC Press; 2007 2007.

Balasubramanian B, Ortiz AL, Kaytakoglu S, Harrison DP. Hydrogen from Methane in a Single-Step Process. Chemical Engineering Science. 1999;54:3543–3552.

Bleeker MF. Pure Hydrogen from Pyrolysis Oil by the Steam-Iron Process. Enschede, Netherlands: Ipskamp Drukkers B.V; 2009.

Bleeker MF, Kersten SRS, Veringa HJ. Pure Hydrogen from Pyrolysis Oil Using the Steam-Iron Process. Catalysis Today. 2007;127(1–4):278–290.

Brandmair M, Find J, Lercher JA. Combined Autothermal Reforming and Hydrogenolysis of Alkanes. In: Proceedings of DGMK Conference on Innovation in the Manufacture and Use of Hydrogen. Dresden, Germany. October 15-17. Page 273-280; 2003.

Brar JS, Singh K, Wang J, Kumar S. Cogasification of Coal and Biomass: A Review. International Journal of Forestry Research. 2012;doi:10.1155/2012/363058 Article ID 363058; accessed July 31, 2013.

Breault RW. Gasification Processes Old and New: A Basic Review of the Major Technologies. Energies. 2010;3(2):216–240.

Brüschke H. Industrial Application of Membrane Separation Processes. Pure and Applied Chemistry. 1995;67(6):993–1002.

Brüschke H. Separation of Hydrogen from Dilute Streams (e.g. Using Membranes). In: Proceedings of DGMK Conference on Innovation in the Manufacture and Use of Hydrogen. Dresden, Germany. October 15-17. Page 47; 2003.

Carolan MF, Chen CM, Rynders SW. ITM Syngas and ITM H2: Engineering Development of Ceramic Membrane Reactor Systems for Converting Natural Gas to Hydrogen and Synthesis Gas for Liquid Transportation Fuels. In: Proceedings of the 2002 U.S. DOE Hydrogen Program Review NREL/CP-610-32405. National Renewable Energy Laboratory, Golden, Colorado; 2002.

Chadeesingh R. The Fischer-Tropsch process. In: Speight JG, ed. The Biofuels Handbook. London, United Kingdom: The Royal Society of Chemistry; 2011:476–517 (Part 3, Chapter 5).

Choudhary TV, Goodman DW. Methane Decomposition: Production of Hydrogen and Carbon Filaments. Catalysis. 2006;19:164–183.

Choudhary TV, Sivadinarayana C, Goodman DW. Production of Cox-Free Hydrogen for Fuel Cells via Step-Wise Hydrocarbon Reforming and Catalytic Dehydrogenation of Ammonia. Chemical Engineering Journal. 2003;2003(93):69–80.

Clark PD, Dowling NI, Hyne JB, Moon DL. Production of Hydrogen and Sulfur from Hydrogen Sulfide in Refineries and Gas Processing Plants. Quarterly Bulletin. 1995;32(1):11–28.

Clark PD, Wassink B. A Review of Methods for the Conversion of Hydrogen Sulfide to Sulfur and Hydrogen. Quarterly Bulletin. 1990;26(2/3/4):1.

Dahl J, Weimer AW. Preprints of Papers-American Chemical Society, Division of Fuel Chemistry. 2001;221.

Donini JC. Separation and Processing of Hydrogen Sulfide in the Fossil Fuel Industry. In: Minimum Effluent Mills Symposium; 1996:357–363.

Dragomir R, Drnevich RF, Morrow J, Papavassiliou V, Panuccio G, Watwe R. Technologies for Enhancing Refinery Gas Value. In: Proceedings of AIChE 2010 SPRING Meeting. San Antonio, Texas. November 7-12; 2010.

Dybkjær I. Tubular Reforming and Autothermal Reforming of Natural Gas – An Overview of Available Processes. Fuel Processing Technology. 1995;42:85–107.

Ehwald H, Kürschner U, Smejkal Q, Lieske H. Investigation of Different Catalysts for Autothermal Reforming of i-Octane. In: Proceedings of DGMK Conference on Innovation in the Manufacture and Use of Hydrogen. Dresden, Germany. October 15-17. Page 345; 2003.

Enger BC, Lødeng R, Holmen A. A Review of Catalytic Partial Oxidation of Methane to Synthesis Gas with Emphasis on Reaction Mechanisms over Transition Metal Catalysts. Applied Catalysis A: General. 2008;346(1–2):1–27.

Find J, Nagaoka K, Lercher JA. Steam Reforming of Light Alkanes in Micro-Structured Reactors. In: Proceedings of DGMK Conference on Innovation in the Manufacture and Use of Hydrogen. Dresden, Germany. October 15–17. Page 257; 2003.

Gross M, Wolff J. Gasification of Residue as a Source of Hydrogen for the Refining Industry in India. In: Proceedings of Gasification Technologies Conference. San Francisco, California. October 8–11; 2000.

Gunardson, H.H., and Abrardo, J.M. 1999. Produce CO-Rich Synthesis Gas. Hydrocarbon Processing, April: 87–93

Hagh BF. Comparison of Autothermal Reforming for Hydrocarbon Fuels. Preprints of Papers-American Chemical Society, Division of Fuel Chemistry. 2004;49(1):144–147.

Ho WSW, Sirkar K. Membrane Handbook. New York: Van Nostrand Reinhold; 1992.

Hufton JR, Mayorga S, Sircar S. Sorption-Enhanced Reaction Process for Hydrogen Production. AIChE Journal. 1999;45:248–256.

Khassin AA. Catalytic Membrane Reactor for Conversion of Syngas to Liquid Hydrocarbons. Energeia. 2005;16(6):1–3.

Kotera, Y., Todo, N., and Fukuda, K. 1976. Process for Production of Hydrogen and Sulfur from Hydrogen Sulfide as Raw Material. U.S. Patent No. 3,962,409. June 8.

Liu K, Song C, Subramani V, eds. Hydrogen and Syngas Production and Purification Technologies. Hoboken, New Jersey: American Institute of Chemical Engineers. John Wiley & Sons Inc.; 2010.

Luinstra E. Hydrogen from Hydrogen Sulfide – A Review of the Leading Processes. In: Proceedings of 7th Sulfur Recovery Conference. Gas Research Institute, Chicago. Page 149-165; 1996.

McKendry P. Energy Production from Biomass Part 3: Gasification Technologies. Bioresource Technology. 2002;83(1):55–63.

Mokhatab S, Poe WA, Speight JG. Handbook of Natural Gas Transmission and Processing. Amsterdam, Netherlands: Elsevier; 2006.

Muradov NZ. CO2-Free Production of Hydrogen by Catalytic Pyrolysis of Hydrocarbon Fuel. Energy & Fuels. 1998;12(1):41–48.

Murata, K., H. Ushijima, K. Fujita. 1997. Process for Producing Hydrogen from Hydrocarbon. United States Patent 5,650,132.

Nagaoka K, Jentys A, Lecher JA. Autothermal Reforming of Methane over Mono- and Bi-metal Catalysts Prepared from Hydrotalcite-like Precursors. In: Proceedings of DGMK Conference on Innovation in the Manufacture and Use of Hydrogen. Dresden, Germany. October 15-17. Page 171; 2003.

Nataraj, S., Moore, R.B., Russek, S.L. 2000. Production of Synthesis Gas by Mixed Conducting Membranes. United States Patent 6,048,472. April 11.

Raissi AT. Technoeconomic Analysis of Area II Hydrogen Production. Part 1. In: Proceedings of US DOE Hydrogen Program Review Meeting, Baltimore, Maryland; 2001.

Rostrup-Nielsen JR. Sulfur-Passivated Nickel Catalysts for Carbon-Free Steam Reforming of Methane. Journal of Catalysis. 1984;85:31–43.

Rostrup-Nielsen JR. Production of Synthesis Gas. Catalysis Today. 1993;19:305–324.

Rostrup-Nielsen JR. Syngas in perspective. Catalysis Today. 2002;71:243–247.

Rostrup-Nielsen JR. 40 Years in Catalysis. Catalysis Today. 2006;111:4–11.

Rostrup-Nielsen JR, Christiansen LJ. Concepts in Syngas Manufacture. London, United Kingdom: Imperial College Press, World Scientific Publishing (UK) Ltd; . Catalytic Science Series. 2011;Volume 10.

Rostrup-Nielsen JR, Christiansen LJ, Bak Hansen J-H. Activity of steam reforming catalysts: Role and assessment. Applied Catalysis A: General. 1988;43:287–303.

Shu J, Grandjean BPA, Kaliaguine S. Asymmetric Pd-Ag Stainless Steel Catalytic Membranes for Methane Steam Reforming. Catalysis Today. 1995;25:327–332.

Speight JG. The Desulfurization of Heavy Oils and Residua. 2nd ed. New York: Marcel Dekker Inc.; 2000.

Speight JG. Natural Gas: A Basic Handbook. Gulf Publishing Company, Houston, Texas: GPC Books; 2007.

Speight JG. Synthetic Fuels Handbook: Properties, Processes, and Performance. New York: McGraw-Hill; 2008.

Speight JG, ed. The Biofuels Handbook. London, United Kingdom: Royal Society of Chemistry; 2011a.

Speight JG. The Refinery of the Future. Elsevier, Oxford, United Kingdom: Gulf Professional Publishing; 2011b.

Speight JG. The Chemistry and Technology of Coal 3rd Edition. Taylor & Francis Group, Boca Raton, Florida: CRC Press; 2013a.

Speight JG. Coal-Fired Power Generation Handbook. Salem, Massachusetts: Scrivener Publishing; 2013b.

Speight JG. The Chemistry and Technology of Petroleum 5th Edition. Taylor & Francis Group, Boca Raton, Florida: CRC Press; 2014.

Speight JG, Ozum B. Petroleum Refining Processes. New York: Marcel Dekker Inc.; 2002.

Udengaard NR, Hansen JHB, Hanson DC, Stal JA. Sulfur Passivated Reforming Process Lowers Syngas H2/CO Ratio. Oil & Gas Journal. 1992;90:62–67.

Uemura Y, Ohe H, Ohzuno Y, Hatate Y. Carbon and Hydrogen from Hydrocarbon Pyrolysis. In: Proceedings of the International Conference on Solid Waste Technology Management; 1999 15: 5E/25-5E/30.

Vauk D, Di Zanno P, Neri B, Allevi C, Visconti A, Rosanio L. What Are Possible Hydrogen Sources for Refinery Expansion?. Hydrocarbon Processing. 2008;87(2):69–76.

Vélez FF, Chejne F, Valdés CF, Emery EJ, Londoño CA. Cogasification of Colombian Coal and Biomass in a Fluidized Bed: An Experimental Study. Fuel. 2009;88(3):424–430.

Vernon PDF, Green MLH, Cheetham AK, Ashcroft AT. Partial Oxidation of Methane to Synthesis Gas. Catalysis Letters. 1990;6(2):181–186.

Wang W, Stagg-Williams SM, Noronha FB, Mattos LV, Passos FB. Partial Oxidation and Combined Reforming of Methane on Ce-promoted Catalysts. Preprints of Papers-American Chemical Society, Division of Fuel Chemistry. 2004;49(1):132–133.

Weimer AW, Dahl J, Tamburini J, Lewandowski A, Pitts R, Bingham C, et al. Thermal Dissociation of Methane Using a Solar Coupled Aerosol Flow Reactor. In: Proceedings of the 2000 the DOE Hydrogen Program Review, NREL/CP-570-28890; 2000.

Wesenberg MH, Svendsen HF. Mass and heat transfer limitations in a heterogeneous model of a gas heated steam reformer. Industrial and Engineering Chemistry Research. 2007;46(3):667–676.

Zaman J, Chakma A. Production of Hydrogen and Sulfur from Hydrogen Sulfide. Fuel Processing Technology. 1995;41:159–198.

Zhu Q, Zhao X, Deng Y. Advances in the Partial Oxidation of Methane to Synthesis Gas. Journal of Natural Gas Chemistry. 2004;13:191–203.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset