Você está na página 1de 87

4 Liquefaction of Coal

4.1 Introduction
Today, the issue of energy security is essential for survival, i.e., for the sustainable development of the global community. The world population is expected to reach 7.0 billion by 2010, and over 55% of this number will be taken up by the population of Asia. Because the demand for petroleum and natural gas in the industrialized nations continues to rise, world petroleum supplies are anticipated to become either unreliable or inadequate in the near future. The local distribution of petroleum in the world has resulted in several crises and supply interruptions, especially since 1973 when the Organization of Petroleum Exporting Countries (OPEC) first achieved sufficient power to have a major impact on world oil markets (Wilson, 1980). Nowadays, there is no doubt that coal is a valuable resource and may be the major fuel of the 21st century. Coal can be considered to be one of the most attractive alternative sources of petroleum oil for the following two reasons. 1) The amount of coal deposits estimated worldwide is ten times larger than that for the other carbonaceous resources. 2) Coal resources are located more widely throughout the world than are oil reserves. Therefore, coal will be more widely available than crude oil in the future. It is necessary to develop new processes for transformation of solid coal into clean liquid from the viewpoint of energy security in the future. 4.1.1 Coal Liquefaction

Many scientists and engineers believe that in the future coal liquefaction will be required to meet the demand for liquid transportation fuels. Coal liquefaction means the conversion of solid coal to fuel liquids (Whitehurst et al., 1980). Coal composed mostly of carbon and hydrogen and has lower hydrogen content than petroleum. Therefore, the process of coal liquefaction produces liquid compounds containing hydrogen at levels of approximately 10 to 15% by weight (Wu and Storch, 1968). Typical compositions for coals, liquid fuels, and some hydrocarbons are given in Table 4.1. Table 4.1 Compositionof Typical Fuel Oils and Hydrocarbons Fuel Typical crude oil Fuel oil Gasoline Bituminous coal Subbituminous coal Benzene Naphthalene Carbon 86.0 86.0 85.0 78.0 71.9 92.3 93.7 Element, wt% (moisture and ash-free) Hydrogen Oxygen Sulfur 11.0 13.4 15.0 5.7 6.1 7.7 6.3 0.7 0.2 0.0 11.6 20.2 0.0 0.0 1.5 0.3 0.1 3.3 0.6 0.0 0.0 Nitrogen 0.5 0.1 0.0 1.0 1.0 0.0 0.0

182

4 Liquefactionof Coal

There are many technologies for coal liquefaction: indirect liquefaction, refinement of coal tar obtained in carbonization at 630-770 K, flash hydropyrolysis in the high pressure of hydrogen, extraction with solvent and/or critical gas, and direct liquefaction by hydrogenolysis in the present of catalyst, solvent and high pressure of hydrogen. Regarding the gasification of coal involving flash hydropyrolysis and carbonization is described in chapter 3. The chapter emphasizes on the direct liquefaction. In indirect liquefaction, coal is gasified at 1300 K or over in the presence of steam and oxygen to produce a synthesis gas containing mostly carbon monoxide and hydrogen. This synthesis gas (syngas), after being cleaned of impurities and adjusted to the desired H2/CO ratio (if required), is converted to liquid fuels in the presence of catalysts. A unique feature of the indirect liquefaction is the ability to produce a broad array of sulfur and nitrogen free products including motor fuels, methanol, oxygenates (octane enhancers), and chemicals with the use of different combinations of catalysts and process conditions. The conversion of syngas to motor fuels is known as Fischer-Tropsch (F-T) synthesis. Commercial indirect liquefaction plants in operation since 1955 have included coal based plants in South Africa and the U.S., and natural gas based plants in South Africa, New Zealand and Malaysia. In all the plants, the syngas is converted in gas phase reactors. Because of the high exotherm associated with the reactions, it has long been known that a liquid phase reactor could offer cost and operability advantages over gas phase reactors due to its superior heat transfer capabilities. Earlier efforts in developing a liquid phase F-T reactor after World War II were suspended in the late 1950s because of the availability of cheap petroleum crude (Poutsma, 1980). Interest in this area was revived in late 1970s with the rise in petroleum crude price. Scoping economics studies supported by the US Department of Energy (DOE) and the Electric Power Research Institute (EPRI) indicated that the capability of a liquid phase reactor to process a low H2/CO ratio syngas from advanced coal gasifier could offer significant cost advantages over gas phase reactors (Gray et al., 1980, Brown et al., 1982). In cooperation with industrial organizations, DOE in 1981 began to support a research and development program to advance the liquid phase reactor technology for coal-based syngas conversion beyond that of the late 1950s. The initial focus of this program has been on the liquid phase reactor technology development for methanol and F-T synthesis. The details of the program have been reviewed (Shen et al., 1996). Liquid phase methanol development was successfully completed at the proof-of-concept (POC) scale in 1989, and advanced to commercial demonstration in 1993 under the support of DOE Clean Coal Technology program. Development of liquid phase reactor technologies for F-T synthesis and for syngas conversion to oxygenates and chemicals have been under way at the POC unit. Direct coal liquefaction involves the conversion of solid coal to liquids without the production of synthesis gas, a mixture of carbon monoxide and hydrogen, as an intermediate step. Direct coal liquefaction should be the most energetically efficient method of producing liquids and this method makes it possible to obtain the highest oil yield. Although the development of coal liquefaction will depend on both economics and the reliability of petroleum and natural gas supplies from the Middle East and other main exporting areas, it is important to develop coal liquefaction technology to insure new sources of energy in the future. The direct liquefaction of coal has been studied extensively. Most of the current processes for the liquefaction of coal have been developed from early works (Bergius and Billiviller, 1918) and have some common features. In most of the conversion processes, as shown in Fig. 4.1 which is a flow diagram of the coal liquefaction process, coal is transport-

4.1 Introduction Coal Liquefaction


f

183

Light oil Hydrogenation r"~ Liquid products

Slurry preparation Hydrogen

~
v

Distillation

M ._ ._ Distillation } Heavy oil

Fig. 4.1 A flow diagramof the coal liquefaction process. ed to a coal slurry preparation process, in which the coal is dried and ground, then mixed with a hydrogen-donor solvent and/or a coal-derived recycle oil to form coal slurry. Then the slurry is pumped into the liquefaction process, in which the coal is liquefied by hydrogen at a high temperature and pressure in the absence/presence of a catalyst in one and/or several high-pressure reactors. The liquefaction product is separated in the distillation process. The heavy oil (residue) and a portion of light gas oil is hydrogenized in the solvent hydrogenation process and transferred to the hydrogen-donor solvent.

4.1.2 Mechanism of Coal Liquefaction


It has become almost axiomatic to formulate coal liquefaction as a free radical process. The concept has its origins in the contribution of Curran et al. (1967) who pointed out the significance of the relationship between the extent of the conversion of the intractable coal molecules to soluble products and the amount of hydrogen transferred to the liquid coal products. They proposed a five-step reaction [Eqs. (4.1)-(4.5)] sequence focused on the homolyses of carbon-carbon bonds in the coal molecules. In this reaction sequence, the radicals produced in the initial reaction, Ri, react with other coal molecules or with hydrogen-atom donor-solvent molecules, DH2, to form other radicals. A variety of recombination reactions terminate the chain reactions. Coal --->2Ri~ Ri" + DH2 ~ Rill q- DHRi~ -Jr- Coal--) Rill -I-- Rj. Ri9 -}- DH"---) Rill q- D
Ri ~ -4- Rj. ---->Rill -+- Argj

(4.1) (4.2) (4.3) (4.4) (4.5)

This view is supported by the general observation that free radical reactions control the pyrolysis chemistry of most organic substances. General kinetic features of coal liquefaction have also been used to support this view (Neavel, 1982). A detailed consideration of the chemical structure of coal and its reaction products also strongly suggests that free-radical reactions control coal chemistry. The aromatic and hydroaromatic units found in coal tars and liquids and presumed to be dominant structures in coal itself are known to be very reactive toward free radicals. Moreover, resonance-stabilized radicals derived from these

184

4 Liquefactionof Coal

structures are formed and react readily at coal decomposition temperatures (-350 ~ Methyl and hydroxy substituents serve to increase the overall free radical reactivity of the molecules to which they are attached. The present analysis accepts the importance of free radicals in coal chemistry and attempts to develop a more detailed and unified view of their structures and properties. In many respects, liquefaction is closely related to pyrolysis. They share an identical initial step - the thermal generation of radicals from the coal by way of homolytic bond scission. In pyrolysis, these radicals are either capped by an internally transferred hydrogen or they combine with carbon to form material of heavier molecular weight (char). These two events also occur in liquefaction, along with transfer of hydrogen to the radicals from a hydrogen source. The net effect is that liquefaction produces greater amounts of liquid and gaseous products than conventional pyrolysis, but at the expense of additional hydrogen consumption. Liquids from hydroliquefaction are substantially depleted of heteroatoms as compared with either the parent coal or pyrolysis liquids. A wide range of different techniques is used to make liquids from coal, even though they all share the thermal conversion step. These methods differ in whether the hydrogen is provided from an organic donor or from molecular hydrogen, either catalytically or non-catalytically. They also differ in whether a solvent, and what kind of solvent, is used. Thus, a study of the physical properties of solutions of coal macromolecules in various solvents as well as the colloidal nature of the solutions would be helpful. An understanding of the phase behavior at high temperatures and under high hydrogen pressures would help to elucidate the liquefaction process. Catalysis in liquefaction has received much attention, although thus far the use of such catalysts as cobalt-molybdenum has not altered process temperature or pressure requirements (Johnson, 1978). Research should be carded out to develop catalysts that will positively affect the initial coal conversion. It is relatively easy to affect the course of reactions after the primary products are out of the coal particle. However, by this time the product distribution may already have been determined. If a catalyst that could influence the product distribution of the primary products as they are formed could be found, entirely different types and quantities of products might result. It would be less important, but still valuable, to determine whether the use of catalysts in coal liquefaction improves the quality of the liquid product. Comparisons of the heteroatom content, aliphatic/aromatic ratios, viscosity and compatibility with petroleum liquids of catalyzed and noncatalyzed coal liquids would be valuable in determining the best disposition of various coal liquid fractions. It would also be valuable to: (1) understand the relationship between coal structure and its reactivity; (2) establish the ultimate dispositions of elements; (3) better understand the kinetics and mechanism of the reaction involved in the catalytic effect, hydrogen transfer, etc.; and (4) develop data which link coal characteristics to process conditions and product type, quality and ultimately, utilization. In this chapter, the advances in the areas mentioned above are described.

4.1.3 Hydrogen Transfer Reaction in Coal Liquefaction


The hydrogen-transfer reactions that occur during coal liquefaction reactions are essential for the conversion of intractable coal molecules into liquids and soluble products. Virtually all the practical processes for coal liquefaction, such as the solvent-refined Coal II process (Schmid and Jackson, 1981), the Exxon donor-solvent process (Furlong et al., 1976), the integrated two-stage liquefaction process (Whitehurst et al., 1980, Neuworth and Moroni, 1981), and the Chevron coal liquefaction process (Rosenthal et al., 1982), use a portion of the liquid coal products as a solvent for the dissolution reaction. In the recently developed

4.2 Coal Structure and Reactivity

185

Chevron coal liquefaction process, the liquefaction reaction is carded out in two separate, but closely coupled, reactors. A slurry of the coal in a portion of the coal liquid (recycle oil) is introduced into the first-stage reactor and the product of this phase of the reaction is then fed into the second-stage reactor. The large coal molecules are decomposed and, in part, dissolved in the first reactor and the initial product is refined catalytically in the second reactor to yield the coal liquefaction products which include a fraction suitable for use as the solvent for the reaction. The conversion reactions require not only the addition of hydrogen but also the redistribution of the hydrogen atom already present in the coal molecules. Thus, hydrogen transfer reactions occur between the coal molecules and the components of the reaction solvent and between the coal molecules and the added hydrogen. Hydrogen transfer reactions also take place between the liquid coal products, the solvent molecules, and gaseous hydrogen. Many of these reactions, particularly those that occur in the initial stages of the coal liquefaction process, are quite rapid even in the absence of catalysts. Indeed, some coals are such good hydrogen atom donors that they only need to be heated in a fluid medium to cause extensive degradation of the carbon skeleton with an attendant redistribution of the hydrogen atoms (Neavel, 1982). More often, however solvents that are good hydrogen-atom donors are used in coal liquefaction reactions to provide a fluid medium for the products as well as to provide a convenient, mobile source of hydrogen for the decomposing coal molecules. In addition, these solvent molecules enable the transfer of hydrogen atoms between the array of hydrogen donors in the solid, liquid, and gas phases and the reactive coal molecules. The reaction pathways important for the transfer of hydrogen atoms during coal liquefaction have been studied intensively in the past few years to establish a more secure basis for the development of efficient methods of coal liquefaction using the available hydrogen atoms in the coal molecules as well as the hydrogen atoms in donor-solvent molecules and added hydrogen. The recent work on this matter is also discussed in this chapter.

4.2 Coal Structure and Reactivity 4.2.1 Stages of Coal Liquefaction


Prior to liquefaction, coal is often washed to remove inorganic minerals and dried. This process sometimes changes the structure and assemblage of coal macromolecules, which profoundly influences the reactivity of coal (Mochida and Sakanishi, 1994). In the preheater, coal with or without catalysts is rapidly heated to reaction temperature in the presence of solvent and pressurized hydrogen extensive decarboxylation, formation of carbonates, and dehydration take place in the preheater (Neavel, 1976). Coal is believed to be substantially dissolved in the preheater at this stage. Rapid heating of up to several hundred degrees per minute is believed to be very essential in obtaining high oil yields and preventing retrogressive reactions, which may take place at the same time. Catalysts are not expected to be effective in the preheater stage due to insufficient contact time. Hydrogen donor solvents play an important role in suppressing the retrogressive reactions at this stage, and it is important that the capacity of the donor not be exceeded in the preheater. The amount of hydrogen consumed from solvent has been considered to be relative to the heating rate. Slow heating rates allow more solvent dehydrogenation (Derbyshire et al., 1986b). It is well known that the viscosity increases very rapidly with bituminous coals that are dissolved in the solvent rapidly in the preheater. This sometimes causes problems of slurry transportation in narrow preheater tubes. The preheated coal slurry (essentially liquefied) is sent to the reactor, where thermal and catalytic cracking, hydrogenation and

186

4 Liquefactionof Coal

hydrocracking take place. These reactions occur rather slowly because fewer reactive bonds are involved in this stage, which produces distillate range small molecules. In the earlier Bergius process, the reaction at this stage was performed under very high pressure at high temperature with disposable catalysts of low activity and was completed in a single step. Current liquefaction processes utilize two or three stages under more moderate conditions. Hydrogen donor solvents also assist in moderating the conditions required. Thus, the primary products in the first stage, together with the used solvent, are further hydrocracked and/or hydrorefined products as well as rehydrogenated solvent. Various types of feeds, distillates, nondistillable liquids free of minerals, the catalyst, preasphaltenes and unreacted coal of the first stage or whole products, including the catalyst and minerals, are charged to the second stage, depending on the liquid/solid separation procedure utilized and the durability of the catalyst in this stage. Staged heating is sometimes utilized in the first stage where the reaction temperature of each reactor is controlled separately to obtain the best oil yield with minimum formation of hydrocarbon gases and avoidance of coking (Burgess and Schobert, 1991, Davis et al., 1989). The oil is further refined in the following stages. Such a process scheme practiced at present is called multistage liquefaction. A series of reaction temperatures is expected to improve selectively specific reactions at different temperatures in a series of consecutive reactions. Higher degrees of desulfurization and denitrogenation, longer catalyst life, less sludge formation, and higher yields of distillate are reportedly obtained by the multistage processing and refining of petroleum products (Mochida et al., 1988a; 1990a). The function of the catalysts in the various liquefaction stages are described in the following sections. 4.2.2 C o a l D i s s o l u t i o n , D e p o l y m e r i z a t i o n , and R e t r o g r e s s i v e R e a c t i o n s

The liquefaction of coal is the conversion of an ensemble of macromolecules as described above into smaller hydrocarbon molecules that are distillable. Shinn (1984) has described the changes in representative molecular structures of intermediates in the three steps of liquefaction as shown in Fig. 4.2a-c. The first step in the liquefaction of solid coal is the formation of liquid phase. Small molecules of the coal fuse above 350 ~ to form a liquid phase together with solvent (if present); some macromolecules may be dissolved in this liquid phase (fusion and dissolution mechanisms). Other molecules undergo thermal fission at their weakest bonds, such as methylene and benzylether bonds, producing fragmented radicals (Whitehurstet et al., 1980). When the radicals are capped with hydrogen from the solvent or the catalyst, they form smaller molecules that are soluble in the solvent or even fusible by themselves (first mechanism) increasing the quantity of liquid phase (Poutsma, 1990). This pyrolysis continues while the reactive bonds and stabilizing hydrogen are available. Atomic or molecular hydrogen available in the reactor system can hydrogenate reactive sites on the aromatic rings. When the ipso-position of the strong aryl-aryl bond in the aromatic ring is hydrogenated, the bond becomes weakened and bond cleavage becomes possible via the first mechanism of depolymerization and facile stabilization (second mechanism) (McMillen et al., 1987; Kamiya et al., 1988; Mochida et al., 1988b). Very reactive hydrogen may attack the aryl-aryl bond directly, leading to its breakage (third mechanism) (Mochida et al., 1990b). Aromatic rings are very stable unless they are hydrogenated to naphthenic tings, which may be thermally or catalytically cracked to open the ring (fourth mechanism) (Malhotra and McMillen, 1990). Unless the fragmented radicals are stabilized, they recombine or react with other molecules, forming thermally stable bonds. Repetition of recombination reactions produces large molecules that have resistance toward depolymerization. Coking takes place when such large molecules remain at elevated temperatures

4.2 Coal Structure and Reactivity

187

OH HO

O O

OH

H 3 C ~

HO "~0 ( ~ ~ C H 3 ~"~"~-'~ 0 ~ . OH H O ' V N - ~ ~ ~ [ ~ )H ~.-.~'S'~~O-0 ~ . , ~ H O , ~ j [~J - ~ O I ~ ~ ~?H O O ~.H,~ ~.'~ I ON ~o ~o~ o

a.f~o~.js~
H,C~ ~ ~_~ ~ HO~~OOH - ~" k~ O
~

2.~ ~ e ~ ~
~ (H o ~ ~" N~O~L.~ C
" c~tT"_ o"

~OOH
"~"",Nvy

x-..,'m

H, OH

H20 20 / NNN~ OHOH

H20

O H N HO O O ~
(a)

HO

3H8 2 CH 4 H3C~~ ; ~ ~ ' ~ kH 3 C ~ 'c ~ H H '" - ~ ~ CH3 l / - O ~ ~ Y ~ O ~ ~ ~ O % ~ H O ~ I ~ o H s ~ C ~ H 3 ~H20 H20 H20

\ .,o= ~ . . ~ L o ~ O - l / . .
H20

~o't~,It'CH; H 2 S
(b)

Fig. 4.2 The Shinn model of a bituminous coal structure [From Shinn, J.H., Fuel, 63, 1190, 1191, 1194 (1984)]

for long time periods, for example, in the locations of low flow rate, such as near reactor walls, in bends of transfer lines, or on catalyst surfaces (retrogressive mechanism) (Bate and Harrison, 1992). When radicals are trapped in the cage of coal macromolecules, such

188

4 Liquefaction of Coal Table 4.2 Half-life Estimated for Carbon-Carbon Bond Homolysis in Some Representative Compounds at 400 ~ in Tetralina Compound b 1,2-Diphenylethane, C6HsCH2-CH2C6H5 2,3-Diphenylbutane, C6HsC(CH3)H-C(CH3)HC6H5 2,3-Diphenyl-2, 3-dimethylbutane, C6HsC(CH3)2-C(CH3)2C6H5 9-(1-Phenylethyl) anthracene, 9-CI4H9CH2-CH2C6H5 9-Benzyl-9,10-dihydrophenanthrene, (9,10-C14HI1)-9-CH2C6H5 Bitetrayl, (1-CIoHIl)-(1-Cl0Hll) Benzyl phenyl ether, C6HsCH2-OC6H5 Benzyl phenyl thioether, C6HsCH2-SC6H5
a

Half-life (min) 1680 20 0.2 5 14 1 0.7 1

The values were selected from the compilation provided by Stein (1981). bThe carbon-carbon, carbon-oxygen, or carbon-sulfur bond cleaved in the hemolytic reaction is indicated in the structural reprasentation. [Reproduced with permission from Stein, S. E., ACS Symposium Ser, No. 169. New Approaches in Coal Chemistry, 7, 104 (1981)]

retrogressive reactions come to act and are accelerated as the radicals frequently encounter each other. The cage hinders liberation of radicals and the participation of donors. Hence, the dissolution of depolymerized coal molecules to break the cage is very important and effective in the liquefaction process (Sakata et al., 1990). Strong dissociative properties of the solvent are important in minimizing macromolecular interactions of coal components or coal-derived products. 4.2.3 F r e e R a d i c a l s in C o a l L i q u e f a c t i o n

It is generally accepted that free radicals are key reactive intermediates in thermal coal chemistry. However, because of the chemical complexity of coal, structural and kinetic properties of these radicals cannot be directly obtained from experiments on coal itself. Therefore, the work that is based on results of well controlled "model" experiments and predictive theory has been carded out (stein 1981). This work extends a previous analysis of coal chemistry (Stein, 1981) in which basic predictive theory was first applied to coal-related molecules and reactions and then used to analyze selected results of model compound experiments. The present analysis uses these predictive methods and results along with relevant experimental data to divide coal-derived free radicals into classes according to their reactivity and to examine the probable behavior of each of these radical classes in coal conversion chemistry. The primary focus of this work is on reactions below 500 ~ To make this analysis tractable, the following working assumptions are made: 1) Coal is composed of randomly oriented, substituted, hydroaromatic clusters tied together by short covalent linkages (biphenyl, methylene, ether (Whitehurst et al., 1980) 2) Free radical reactions account for all covalent bond breaking and forming processes and most types of hydrogen transfer 3) All varieties of elementary free radical reactions of importance in coal conversion are already known (O'Neill and Benson, 1973) 4) Steric and diffusional effects do not influence reaction kinetics. In the past decade, many contributions have been presented to support the general reaction scheme as shown in Eqs. (4.1)-(4.5). The first reaction in the sequence involving the homolytic cleavage of carbon-carbon bonds in the coal molecules is the critical step. Studies of the kinetics of the decomposition of a variety of compounds with relatively weak carbon-carbon bonds have shown that the rates of decomposition of these substances by well-known free radical pathways are comparable with the rates of decomposition of coal

4.2 CoalStructure and Reactivity

189

molecules. This feature is well illustrated by the data presented in Table 4.2 (Stein, 1981), One of the best lines of evidence for the involvement of free radicals in these processes stems from the study of the rapid pyrolysis of coal in the cavity of EPR spectrometers (Petrakis and Grandy, 1981; Sprecher and Retcofsky, 1983). In one quite pertinent study. Sprecher and Retcofsky investigated the thermal decomposition or a bituminous coal suspended in silica (Sprecher and Retcofsky, 1983). The concentration of radicals increased by a factor of 4 in about 5 min at 470 ~ The addition of an equal weight of 9,10-dihydrophenanthrene to the reaction mixture inhibited the formation of radicals, whereas the addition of an equivalent amount of phenanthrene had no effect on the radical concentration. In addition, it was found that the radical concentration increased by an additional factor of 2 when the volatile products formed in the pyrolysis of this coal were allowed to escape from the reaction vessel. These results strongly support the idea that coal molecules decompose by homolytic reactions to yield transient reactive radicals. In addition, the results are compatible with the view that hydrogen-atom-abstraction reactions occur rapidly with effective donor molecules such as 9,10-dihydrophenanthrene. As illustrated in reactions (4.2) and

Rj,-+- ~

RjH + ~

(4.6)

(4.6), a new coal molecule is produced together with a mobile radical. These react with other coal molecules as illustrated in Eqs. (4.3) and (4.7) to yield new coal molecules and a different series of coal radicals resulting from the abstraction of hydrogen atoms rather than from the homolyses of carbon-carbon bonds.
Ri 9

+ volatile products ---->Rill

n t-R3"

(4.7)

The fourth reaction in the sequence describes the behavior of donor-solvent molecules, such as tetralin and 9,10-dihydrophenanthrene, that can readily be oxidized to aromatic compounds. The abstraction reactions of the dihydronaphthalenes [Eq. (4.9)] are more rapid than the abstraction reactions of tetralin. The removal of hydrogen atoms from the intermediate radicals by reactions with other coal radicals are also presumably very rapid processes. Some radicals derived from the coal and from the solvent engage in dehydrogenation reactions. The fifth reaction in the basic sequence illustrates the notion that hydrogen atoms are transferred among coal molecules to yield both hydrogen-rich and hydrogenpoor compounds during the coal liquefaction process (Neavel, 1982). These processes are

~ @ @o

+ Ri9 --t-Ri9 + Ri9

---> ~ ----) @ 9 ~ ~

--t--Rin -F-RiH -+Rill

(4.8)

(4.9)

(4.10)

complemented by another series of reactions which are initiated by the abstraction of hydrogen atoms from the coal molecules to provide another unstable series of radicals [Eq. (4.3)]. These reactions are certain to be important in coals with abundant hydroaromatic

190

4 Liquefactionof Coal

structures because the benzylic carbon-hydrogen bonds are relatively weak and the activation energies for the transfer of hydrogen atoms from these positions to other coal radicals are modest. While some of the radicals formed in benzylic hydrogen-atom abstraction eventually undergo aromatization, other radicals formed as outlined in Eq. (4.3) decompose by the very well-known zr-scission process (Sweeting and Wilshire, 1962; Collins et al., 1979; Poutsma and Dyer, 1982; King and Stock, 1984) to yield a new radical and a highly reactive alkene as illustrated for 1,3-diphenylpropane in Eq. (4.11). The fragmentation reactions decrease the molecular weight of the coal molecules and the reactive products propagate the liquefaction reaction. C6HsCHCH2CH2C6H5 -'-) C6HsCH29 + C6HsCH - CH2 (4.11)

While most discussions have focused attention on the dominant free radical processes that take place during the thermal decomposition of coal molecules during coal liquefaction, it is apparent that pericyclic reactions can make a major contribution to the degradation reactions of the complex molecules under appropriate conditions. There are a variety of plausible pericyclic reactions that must be considered in the development of an adequate theory for the noncatalytic thermal reactions of coal molecules. Moreover, processes of this kind are certain to be more important in the more severe reactions of coal molecules at temperatures in excess of 500 ~ The reactions include the unimolecular transfer of hydrogen from one hydrocarbon to another hydrocarbon. Doering and Rosenthal (1967) provided the classic example of the reaction when they showed that the Z isomer, rather than the E isomer, of 1,2-dimethylcyclohexane was obtained preferentially (6% yield) during the decomposition of the dihydronaphthalene. However, few other authentic examples of the reaction have been reported presumably because free radical chain reactions occur competitively obscuring the nonchain pericyclic reactions (Fleming and Wildsmith, 1970). For example, studies of the hydrogen-transfer chemistry of 1,2- and 1,4-dihydronaphthalene and other dihydroaromatic compounds unequivocally indicate that pericyclic processes are not involved in the product-forming stages of the reactions with E-stilbene, phenanthrene, and tetracene (King and Stock, 1981) and Heesing and Mullers (1980) demonstrated that the hydrogentransfer reactions which take place during the disproportionation of 1,2-dihydronaphthalene at 300 ~ do not occur in a pairwise fashion or stereo specifically. H H

H3C
H3C

H3C H U3C H

(4.12)

Unimolecular dehydrogenation is almost certainly a more important process. Evidence for this reaction which leads directly to aromatic compounds is much more abundant (Derbyshire et al., 1982). These processes may be regarded as termination reactions in coal liquefaction. ~ @ + H2 (4.13)

The notion that pericyclic reactions are important for carbon-carbon bond-cleavage reactions remains a matter of controversy. Virk and his co-workers have advanced the view that retroene reactions and related processes are important in the decomposition reactions of

4.3 Catalysisin Coal Liquefaction

191

simple hydrocarbons at the threshold temperature of liquefaction, 400 ~ (Virk, 1979). Tests of this proposal by the study of the rates and products of decomposition of 1,2diphenylethane, 1,3-diphenylpropane, and 1,4-diphenylbutane reveal that the retroene process is insignificant for the formation of the reaction products (Stein, 1981; Poutsma and Dyer, 1982; Hung and Stock, 1982). The situation is well illustrated by the pathway followed in the decomposition of labeled 1,4-diphenylbutane. The radical chain decomposition reaction predicts that unlabeled toluene [Eqs. (4.14)-(4.16)] be formed, whereas the C6HsCH2CD2CD2CH2C6H5 q- Ro --4 RD + C6HsCH2CDCD2CH2C6H5 C6HsCHzCDCD2CH2C6H5 ~ C6HsCH2CD=CD2-1-9 (4.14) (4.15) (4.16)

9CH2C6H5 + Tetralin --~ C6HsCH3 + 1-Tetralyl radical

retroene process requires the formation of toluene-2-d [Eqs. (4.17)-(4.18)]. No more than 2% toluene-2-d is obtained during the decomposition reaction in tetralin at 400 ~ Thus, reactions Eqs. (4.14)-(4.16) appear to be the principal toluene-forming reactions. While it seems clear that retroene processes are not responsible for the product-forming reactions, it is quite possible that such pericyclic processes may be responsibleindirectly for the

cvI:
C6HsCH2CD2CD2CH2C6H5 D H + C6HsCHzCD = CD2 (4.17)

[~
D

CH2 several steps > Toluene-2-d H

(4.18)

initiation of free radical chain reactions. This idea requires consideration because the retroene reactions often produce very unstable intermediates. To examine the concept, we recently studied the decomposition of 9-[3-(perdeuteriophenyl)propyl-3,3-d2]-phenanthrene (Stock, 1985). Group-additivityconsiderations suggest that the retroene reaction is CH2CH2CD2C6D5 _._) CH2 -+- C6DsCD ----CH2 (4.19)

endothermic by no more than 38 kcal mol-1. However, the products formed in this reaction are quite unstable under the conditions of coal liquefaction and may initiate free radical chain reactions with quite low effective activation energies. Thermal chemical analyses (Benson, 1976) suggest that the molecule-induced homolysis shown in Eq. (4.20) is endothermic by only 21 kcal mol-1. More important, the abstraction of the benzylic and CH2 + C6HsCH--CH2 ~ CH2 ~ -q- C6H5CHCH3 (4.20)

192

4 Liquefaction of Coal

allylic hydrogen atom in the 9-methylphenanthrene isomer is a very facile process [Eq. (4.21)] which leads to a new radical capable of initiating the decomposition of the original alkylphenanthrene by a conventional reaction sequence. Considerations of this kind support the view that certain low-energy pericyclic reactions may lead to the initiation of free radical chain reactions. In this sense, such reactions may influence the rates of the thermal decomposition reactions of coal molecules.
CH2 CH2 ~

+Ri" ~ ~

-'lRill

(4.21)

Other pericyclic reactions such as the decarbonylation of phenolic aromatic compounds [Eq. (4.22)] and the fragmentation of tetralin [Eq. (4.23)] apparently are too slow to be important under the conventional conditions used in liquefaction reactions (Cypres and Bettens, 1974; Berman et al., 1980). Phenol --+

Tetralin --+

~ ~

O -+ CO + Cyciopentadiene CH2 CH2 + CH2 = CH2 (4.22) (4.23)

4.3 Catalysis in Coal Liquefaction


The coal liquefaction proceeds even in the absence of the catalyst. There are essentially no use of catalysts in the processes developed in USA such as SRC, SRC-II, EXXON DONOR SOLVENT (EDS), in which the iron species present in ash of coal and hydrogen-donor solvent were utilized to enhance the liquefaction of coal (Suzuki and Ikenaga, 1998). On the other hand, the use of catalysts is considered to enhance the oil yield in coal liquefaction and has been spotlighted in a lot of studies, especially in Japan. In fact, the NEDOL process, which the catalyst and hydrogen-donor solvent was used, had showed higher oil yield than any present processes. The most conventional catalytic material is iron sulfide in various types. Iron-sulfur catalyst systems have been successfully employed for direct hydrogenation during coal liquefaction on a commercial scale (Wu and Storch, 1968). Nowadays, they are preferred because they are simple to use and because of economic reasons. Iron-sulfur catalyst systems have been widely investigated by several authors (Montano and Granoff, 1980; Montano et al., 1981; Cypres et al., 1981; Baldwin and Vinciguerra, 1983; Lambert, 1982; Stenberg et al., 1983; Mukherjee and Mitra, 1984; Trewhella, 1987). It was suggested that the highest conversion of coal to liquid products was associated with a pyrrhotite, which had the largest number of vacancies. Moreover, both H2S and pyrrhotite appeared to play a significant role in the coal liquefaction process (Montano and Granoff, 1980; Montano et al., 1981; Suzuki and Ikenaga, 1998). In most studies, it was recognized that the active form of iron-sulfur catalysts was pyrrhotite. There is an alternative interpretation that the catalyst was actually H2S produced from the reduction of pyrite (Lambert, 1982; Mukherjee and Mitra, 1984). Pyrite, pyrrhotite, and various nonstoichiometric sulfides are known, and pyrrotite is

4.3 Catalysisin Coal Liquefaction

193

postulated as the active form. Its precursors are red mud, residue of bauxite after the separation of alumina, iron ores of various sources, synthetic and natural pyrite, fine iron particles, iron dust from converters, iron sulfate, iron hydroxide, etc. (Montan et al., 1981; Suzuki et al., 1989). It is also important to elucidate the catalytic roles played by ironbased catalysts and sulfur, which are added during coal liquefaction. The next most widely used materials are Co-Mo and Ni-Mo sulfides, which have been widely used in petroleum refineries. They are usually supported on alumina of designed pore structures in which the pore diameter is usually larger than that for conventional petroleum residue (Stephens et al., 1985; Derbyshire, 1989). A third type of material is the chloride of transition metals, such as ZnC12 and SnC14 (Mobley and Bell, 1980; Mizumoto et al., 1985). This group of catalysts works in molten state in contrast to the solid state of the previous two groups. The corrosive nature and instability may exclude their practical application. No details are reviewed here. Ru has been used as an additive to Co-Mo and Ni-Mo (Hirschon et al., 1987) to improve their hydrogenation and denitrogenation activities. Hydrogen sulfide in the reaction atmosphere has been reported to accelerate liquefaction directly, in addition to controlling the extent of sultiding of iron, Ni-Mo, and Co-Mo catalysts (Ogawa et al., 1984; Hirschon and Laine, 1984). Recently, carbon black was reported to catalyze coal liquefaction (Farcasiu and Smith, 1990, 1991); this may initiate radical reactions of bond breakage. 4.3.1 Preparation of Catalysts

Solid liquefaction catalysts have been prepared by three procedures. 1) Fine Powder Catalysts. Most iron catalysts are used in powdered form. Since their particle size strongly influences their activity, fine powders are preferred. Natural products are ground extensively. Magnetite for the magnetic tape is needle-like crystal of which the diameter is less than 1 /2m. Recently, ultra-fine powders (nanometer to tens-of-nanometer size scale) of iron oxides and sulfides have been prepared by means of vapor-phase hydrolysis of volatile compounds in a hydrogen-oxygen flarhe to produce nanometer-sized iron oxides (aerosol) (Bacaud et al., 1993; Lacroix et al., 1989); rapid thermal decomposition of solutes (RTDS), such as Fe(NO3)3 solutions (Matson et al., 1993); laser pyrolysis of Fe(CO)5 and C2H4 to produce iron carbides followed by in situ sulfidation (Hager et al., 1993); precipitation/crystallization sequence from the sulfated and oxyhydroxides of iron (Pradhan et al., 1993); and a chemical reduction or an exchange/replacement reaction of iron salts solubilized in inverse micelles of reaction media (Martino et al., 1993). Finer powders of the iron sulfide are expected to be expensive as well as active. The cost/performance should be carefully evaluated. 2) Supported Catalysts. Co-Mo and Ni-Mo sulfides are usually supported on alumina. Selection of the specific alumina is conventionally studied on the basis of the pore size distribution and acidic characteristics. The supporting procedure and the amount of supported sulfides are very influential in catalyst activity. Alternative supports to alumina are the focus of current research. Titania and carbon have recently been examined as supports for iron and Ni-Mo sulfides (Mochida et al., 1984; Derbyshire et al., 1986a). Bifunctional and strong interactive roles of the support should be emphasized in addition to physical properties (Ehrburger et al., 1976; Showa, 1982). The search for additives such as phosphate and sulfate, which have been utilized for commercial Co-Mo and Ni-Mo base catalysts, has also been receiving much recent attention (Lewis and Kydd, 1991; Morales et al., 1984; DeCanio et al., 1991). 3) Highly Dispersed Catalysts on Coal. Sulfide catalysts have been dispersed directly on

194

4 Liquefaction of Coal

the coal surface. Very high dispersion on the catalyst may allow direct interactions between the catalyst and solid coal. The first application of this approach utilized molten chloride as the starting material. Later, oil- and water-soluble iron precursors were impregnated or ion exchanged onto the coal surface through the interaction with oxygen functional groups (Cugini et al., 1991; Hirschon and Wilson, 1991a; Curtis and Pellegrino, 1989; Snape et al., 1989; Pradhan et al., 1991). Recently, highly dispersed, highly active, or highly functional catalysts have been extensively investigated to reduce the amount of catalyst required for recovery and regeneration (Suzuki et al., 1984, 1987, 1991; Holloway and Nelson, 1977; Takemura et al., 1989; Curtis and Cahela, 1989; Ryan and Stacey, 1984; Mendez-Vivar et al., 1990). Very fine particles of iron sulfide are very promising catalysts because of lower cost and moderate activity. Presulfiding treatments for activation, ion exchange, and dispersed impregnation of catalysts or catalyst precursors are combined to enhance the catalytic activity and reduce the amount of catalyst required (Naumann et al., 1982; Yokoyama et al., 1983). The use of highly dispersed catalysts from soluble salts of molybdenum is another approach to the reduction of catalyst amount because of their excellent activity despite their higher price. Recently, metal carbonyl com-pounds, such as Fe(CO)5, Ru3(CO)12, and Mo(CO)6 have been investigated as metal cluster catalysts. Preparation involved their deposition and decomposition on catalyst support surfaces (Paradhan et al., 1991; Burgess and Schobert, 1991; Cowans et al., 1987). It has been reported recently that highly dispersed catalyst on coal grains can accelerate the liquefaction of the coal grains without supporting catalysts (Cugini et al., 1991; Pradhan et al., 1991). The fine powders of the catalyst are indicated to be mobile during the liquefaction, suggesting no importance of direct interaction. Recoverable catalysts also offer a promising way to economize the cost of liquefaction catalysts (Joseph, 1991; Pelofsky, 1979). Dow designed a process that utilized fine powders of MoS2 that were reported to be recoverable by hydroclone; however, specific details have not been published (Whitehurst, 1980). 4.3.2 Fe-based Catalysts

It was reported that all Fe-based catalysts (FeOOH, Fe203, pyrite, FeSO4, etc.) show similar yields of liquid when 0.5% Fe was added. On the other hand, FeOOH catalyst that was distributed on the surface of coal as fine particles via impregnation showed higher yield of liquid than other catalysts, as shown in Fig. 4.3 (Cugini et al., 1994). The sintering of the FeOOH nanometer particle catalyst still occurred in the sulfidation and resulted in a decrease of surface area even though its initial value was ca. 138 m2/g, resulting in lower catalytic activity. This indicates that the higher the distribution of the sulfides is, the higher the catalytic activity. It was proposed that natural pyrite was activated via the grinding in NEDOL process because of the lower activity of natural pyrite. The relationship between particle size of pyrite and the yield of liquid in the liquefaction of Wandoan coal is shown in Fig. 4.4 (Hirano et al., 1997). The yield of oil significantly increased when the average particle size was under 50 ~tm, and the yield increased from 40% for particle size of 100/~m, i.e., unground pyrite, to 58% for particle size of 0.86/.tm, which is over the objective of the NEDOL method. By XPS analysis, it was found that there is ca. 10% of SO~- on the surface of natural pyrite. In contrast, most of the surface of ground pyrite in air is present in the form of SO 2-. Thus, the grinding in the oil phase or the addition of sulfur after the oxidation will restore the catalytic activity of the pyrite (Hirano et al., 1997). The preparation of nanometer particle ~,-geitite (FeOOH) was developed by Kaneko et al. (1995). In the liquefaction of Yallourn coal, the yield of oil was only 40-43% in the

4.3 Catalysis in Coal Liquefaction

195

100

75-

50

~.

~D

25

+ o 2:

0 = 0 0
J

o =

+ ~ 0 0

< z

Pow~der [] CH2C12 soluble,

Y Impregnation

9 Heptane soluble

Fe203 5000 ppm, FeOOH 2500 ppm, Ammonium heptamolydate (AHM) 1500 ppm Fig. 4.3 Effect of several catalysts on liquefaction of Illinois #. 6 coal at 450 ~ and under 1000 psi of H2 for 1 h. [Reproduced with permission from Cugini, A.V. et al., Catal. Today, 19, 401 (1994)]

case using pyrite and a-FeOOH ground to a particle size of 0.5 pm. On the other hand, the yield of oil was 50% using ~,-FeOOH of 1.2 pm, and increased to 55% when ~,-FeOOH of 0.4 pm was used. The higher activity of 7'-FeOOH was attributed to the difference in crystallization, whereby ~,-FeOOH can be readily transformed to the active phase, i.e., pyrrhotite at temperatures under 250 ~ and growth of crystal at high temperatures was suppressed. In the analysis of XRD of variable temperature units in the sulfidation atmosphere for FeS2, ?'-FeOOH and limonite, it was found that ?'-FeOOH and limonite were completely transformed into pyrrhotite at 300 ~ whereas the peak of FeS2 was still present at 300 ~ and a temperature of 400 ~ was necessary for its complete disappearance. Further, the crystal size of sulfided ~,-FeOOH and limonite was c a . 10 nm at 250 ~ and merely grew to 20 nm even at 400 ~ In contrast, the crystal size originating from FeS2 was c a . 40 nm and second particles further formed due to cohesion of first particles. Consequently, the crystal size was 2-4 times greater than when ?'-FeOOH and limonite were used. Thus, it is important to form active pyrrhotite at lower temperatures and to avoid the growth of greater particles as much as possible (Suzuld and Ikenaga, 1998). In addition, Fe(CO)5 and the Fe(CO)5-sulfur system are also often used in the liquefaction of coal. It is important to elucidate the catalytic roles played by highly dispersed catalysts and sulfur which are added during coal liquefaction although these are not well defined. On the other hand, since the reactions involved in coal liquefaction include hydrocracking and hydrogenation by molecular hydrogen and donor solvents, a number of at-

196

4 Liquefactionof Coal 70

60-

50-

-.,,.
40 0.1 1.10 ll0 Average radius (~m) 100

Fig. 4.4 Relationshipbetween average radius of pyrite particle and oil yield in liquefaction of Wandoan coal at 450 ~ and under 75 kg/cm2of H2 for 1 h in a 5 L autoclave. [Reproducedwith permissionfrom Hirano, K. et al., J. Jpn. Inst. Energy, 75, 911 (1997)] tempts have been made to elucidate the mechanism of hydrogen transfer occurring during coal liquefaction in the presence of solvents (Billmers et al., 1986; Camaioni et al., 1993; Autrey et al., 1995; McMillen et al., 1985, 1991; Malhotra and McMillen, 1993; Murakata et al., 1993). A useful method for clarifying the mechanism of hydrogen transfer in coal liquefaction is to utilize isotopes such as deuterium and tritium tracers (Fu and B laustein, 1963; Franz and Camaioni, 1980; Brower, 1982; Schweighardt et al., 1976; Cronauer et al., 1982; Skowronski et al., 1984; King and Stock, 1982; Collin et al., 1980). Recently, it has been reported that tritium and carbon-14 tracer techniques were effective in quantitatively monitoring the hydrogen during coal liquefaction (Kabe et al., 1986a, 1987a, 1987b, 1990d, 1991 a, 1991 b; Montano et al., 1981; Ishihara et al., 1993, 1995). These reports showed that the hydrogen mobility of coal and coal-related compounds could be quantitatively analyzed using the hydrogen exchange reactions occurring between coal, the gas phase and the solvent, as well as by considering the hydrogen addition reactions. Recently, Godo et al. (1997a) reported a liquefaction of Taiheiyo coal in the presence of conventional and dispersed iron catalysts (pyrrhotite and Fe(CO)5) and in the presence of sulfur, in which a tritium and radioactive 35S dual-tracer method was used. As we know, tetralin is one of the most simple and convenient model compounds because it contains both an aromatic ring and a naphthene ring, and because it can serve as a hydrogen donor solvent. Before investigating the complex reaction with coal, the reaction of tetralin with tritiated hydrogen was performed in the absence of coal. Figs. 4.5 and 4.6 respectively show the products and tritium distributions in the exchange reaction between tetralin and tritiated hydrogen. Table 4.3 shows the amount of hydrogen exchanged between the gas phase and the solvent. Although naphthalene (NP), 1-methylindan (MI) and n-butylbenzene (BB) were produced in the absence of a catalyst and sulfur (Run 1), the yield of each product was very low, with the virtual absence of hydrogen exchange. When sulfur was added (Run 2), the yield of NP increased significantly (from 0.7 to 3.0%), with most of the added sulfur converted into hydrogen sulfide (Table 4.3). If all the added sulfur reacted with tetralin to produce hydrogen sulfide, it would correspond to the production of 2.7% of NP. These results indicate that sulfur promotes the dehydrogenation of tetralin to

4.3 Catalysis in Coal Liquefaction Run Feed Catalyst Sulfur

197

~:~:~:~1~
-lg ~iii!zi~ D Butylbenzene [-1 Methylindan

3 4 5 6

Pyrrhotite Pyrrhotite Fe(CO)5 Fe(CO)5 lg lg 2g 0 ' 92' 94' ~:iii~iii 96' 9'8 ' 100%

Fig. 4.5 Products in the exchange reaction between tetralin and tritiated hydrogen. [From Godo, M. et al., Energy Fuels, 11, 726 (1997)]

Run
1 2

Catalyst

Sulfur

lg

9 []

Solvent Gas

3 4 5 6

Pyrrhotite Pyrrhotite Fe(CO)5 Fe(CO)5 1g 1g 2g

Tritium equilibrium
I I I ' I '

20

40

60

80

100%

Fig. 4.6 Tritium distribution in the exchange reaction between tetralin and tritiated hydrogen. [From Godo, M. et al., Energy Fuels, 11,726 (1997)]

produce naphthalene. Further, the yield of MI also increased from 0.3 to 0.6%, and the yield of BB from 0.1 to 0.3%; the tritium distribution in the solvent increased from 0.6 to 7.7%. When pyrrhotite was added (Run 3), the yields of NP, MI and BB were respectively 2.0, 0.7 and 0.4 %, and the tritium distribution in the solvent was 59%. These values therefore increased significantly, but the amount of HzS generated was very small. When pyrrhotite and sulfur were added simultaneously (Run 4), the tritium distribution in the solvent amounted to 64%. This was very close to the equilibrium value of 82%, which was calculated on the assumption that the hydrogen atoms were completely and randomly dispersed between the gas phase and the solvent. It was assumed that the independent effects of the sulfur and pyrrhotite would considerably increase the amount of hydrogen exchange. When Fe(CO)5 and sulfur were used (Runs 5 and 6), the yields of MI and BB were not as high as that obtained with pyrrhotite and/or sulfur. The tritium distributions in the solvent in Runs 5 and 6 increased to 16% and 27%, respectively, significantly greater than in the absence of catalysts (Runs 1 and 2). However, these values were smaller than those ob-

198

4 Liquefactionof Coal Table 4.3 HydrogenTransfer between Gas Phase and Solvent Run 1 2 3 4 5 6 Catalyst m ~ Pyrrhotite Pyrrhotite Fe(CO)5 Fe(CO)5 Sulfur m 1g ~ 1g 1g 2g Amount of hydrogen Amountof H2S exchanged (g) generated (g) 0.013 0.156 2.032 2.428 0.275 0.410 0.000 0.800 0.121 1.011 0.012 1.041

[From Godo, M. et al., Energy Fuels, 11,726 (1997)] tained with pyrrhotite. In the case of Fe(CO)5, part of the added sulfur produced iron sulfide, with the remainder being converted into H2S (Table 4.3). After the reaction in the presence of pyrrhotite and Fe(CO)5, iron sulfide was recovered from the inside of the autoclave. Compared with the pyrrhotite reaction, the particle size of the iron sulfide recovered from the reaction with Fe(CO)5 and sulfur was larger and more uniform. Further, many large fragments had plain surfaces. It was assumed that these fragments had been deposited on the inside wall of the autoclave. Although Fe(CO)5 is one of the most highly dispersed iron catalysts used in direct coal liquefaction, the particle size of the iron sulfide recovered in the absence of coal was about 1.2 pm, which was larger than the particles of pyrrhotite (about 0.6/.tm). It appears that compared with pyrrhotite an increase in particle size results in a decrease in the amount of hydrogen exchanged. It was assumed that the formation of products and the hydrogen transfer from the tetralin proceeded via a tetralyl radical, which acted as an intermediate in the conversion and the hydrogen exchange of tetralin (Kabe et al., 1990d-e; Ishihara et al., 1995). When radicals generated in coal react with tetralin in the system, a tetralyl radical may be formed easily. However, if coal is not included in the system, a tetralyl radical is difficult to generate. In the system of tetralin and gaseous hydrogen, tetralin may collide with itself to give a tetralyl radical (Eq. (4.24)). The hydrogen exchange between tetralyl radicals and tritiated hydrogen or tritiated hydrogen sulfide can be assumed to proceed via Eq. (4.25) depending on the concentration of tetralyl radicals, which also controls the formation of methylindan in Eq. (4.26). The products were methylindan by isomerization, butylbenzene by hydrocracking and the main product, naphthalene by dehydrogenation. Decalin by disproportionation was not formed. This is consistent with a previously reported result (Hooper et al., 1979; Jan et al., 1984). Butylbenzene may be formed by the dissociation of tetralin with hydrogen at on as shown in Eq. (4.27). Concerning the hydrogen exchange reaction of tetralin, studies using deuterium have been reported extensively. In hydrogen exchange between tetralin-dl2 and hydrogen in coal at 400 ~ 1 h in a shaken autoclave system, protium was incorporated into H~ (66%), H~ (23%) and Hat (11%) positions in tetralin and that H~ absorption of the recovered naphthalene in ~H nuclear magnetic resonance (NMR) was approximately seven times as intense as H~ absorption (Skauronski et al., 1984). Collin and Wilson (1983) showed in their NMR measurement that, in the reaction of tetralin with deuterium and coal, the intensity of the signal at the H~ position was larger than that at the H~ position. These reports indicate that the 1-tetralyl radical appears to be a more important intermediate than the 2-tetralyl radical in the conversion and the hydrogen exchange of tetralin, although it has been proposed that 2-tetralyl radical may be an intermediate in the formation of methylindan (Franz and Camaioni, 1980). Therefore, the formation of the 1tetralyl radical in this system may be the rate-determining step for both the conversion of

4.3 Catalysisin Coal Liquefaction ~ HT or TSH T + H. or SH.


CH2 9

199 (4.24)

+ H.

(4.25) CH3

H9

(4.26)

+ H.

H9

(4.27)

+ SH9

(4.28)

+ H2S + SH9

(4.29) (4.30) (4.31)

-+- H2S

tetralin and the hydrogen exchange. It was previously reported that the rate of conversion of tetralin in the presence of H2S was nearly equal to that in the absence of H2S (Godo et al., 1997a). In contrast to the result with H2S, it was considered that sulfur promoted the formation of tetralyl radicals. The processes of radical formation could be represented as Eqs. (4.28)-(4.31), which is consistent with the fact that sulfur promotes naphthalene formation. Because sulfur promotes the formation of tetralyl radicals, the concentration of tetralyl radicals will increase. As a result, the formation of products was promoted by sulfur. When pyrrhotite was added, the products yields increased and the tritium distribution in the solvent increased significantly. Autrey et al. (1996) suggested that FeS catalysts are reduced by donor solvent. It was considered that pyrrhotite promoted not only the formation of the tetralyl radicals in Eq. (4.32) but also the dissociation of the tritiated hydrogen molecules in Eq. (4.33). The liquefaction of Taiheiyo coal with tritiated gaseous hydrogen was performed in the presence of catalysts and sulfur. Fig. 4.7 shows the distribution of the products from coal. Without the catalyst and sulfur (Run 7), the product yield ( = 100-residue) was 74% and the SRC was 71%. In the presence of sulfur (Run 8), although there was an increase in light fractions (such as light oil), the product yields decreased. This suggests that sulfur simultaneously promoted both the thermal decomposition and polycondensation of coal. In Cat. -Cat. + H-Cat. (4.32)

T2

2T-Cat.

(4.33)

200

4 Liquefaction of Coal

the presence of pyrrhotite or Fe(CO)5, the product yields in liquefaction increased to more than 80%; the use of sulfur also increased the product yields. When Fe(CO)5 and 2 g of sulfur were added (Run 12), the product yields reached 89%, which was higher than those obtained with pyrrhotite. In the presence of Fe(CO)5 and sulfur, the yield of preasphaltene decreased and the yields of oil and asphaltene increased in comparison with the use of pyrrhotite. This indicates that Fe(CO)5 and sulfur could be used to obtain products with a lighter composition. The result also indicated that the liquefaction activity of Fe(CO)5 was higher than that of pyrrhotite. Figure 4.8 shows the tritium distributions in Taiheiyo coal liquefaction. Compared with the absence of sulfur (Run 7), the addition of sulfur (Run 8) increased the tritium distribution in the solvent and decreased that in coal. The use of pyrrhotite only (Run 9) increased the distributions in both coal and the solvent. When pyrrhotite and sulfur were both added (Run 10), the tritium distribution in the solvent increased and that in coal decreased. It was believed that the sulfur promoted the hydrogen transfer from the gas phase to the solvent, and that the reaction mechanism was the same as that occurring in the absence of coal (described above). In contrast, when Fe(CO)5 and sulfur were used, the tritium distribution in the coal increased in spite of the presence of sulfur, suggesting that the catalyst derived from Fe(CO)5 and sulfur acted directly on the coal and increased both the rate of coal
Run Catalyst Sulfur

:o:.:***:.:,:o:.:.:.:.:.~
8 9 10 11 12
-

lg
-

i !ii !!iii i ii!


:i:::::i::!:i:i::::::::!:~

Pyrrhotite Pyrrhotite Fe(CO)5 Fe(CO)5

1g 1g 2g
' I ' I ' I ' I '

n THFIS VA BI-THFS im HI-BS IN HS F-! Light-oil ! ! Naphtha E! Gas

20

40

60

80

100 daf%

Fig. 4.7 Products from coal in Taiheiyo coal liquefaction using gaseous tritium. [From Godo, M. et al., Energy Fuels, 11,728 (1997)] Run Catalyst Sulfur

8 9 10 11 12

-Pyrrhotite Pyrrhotite Fe(CO)5 Fe(CO)5

lg
--

i I~

Coal Solvent Gas

1g 1g 2g

iiiiiiii iiiiiiiiiiiii iiiii!iii i iiiiiiiill /iiiiiiii!ii!i!iiiiil


o 'o4o6'o8o oo

Fig. 4.8 Tritium distribution in Taiheiyo coal liquefaction using gaseous tritium. [From Godo, M. et al., Energy Fuels, 11,728 (1997)]

4.3 Catalysisin Coal Liquefaction Table 4.4 HydrogenTransfer among Gas Phase, Solvent and Coal Amount of hydrogen transferred (g) Run 7 8 9 10 11 12 Catalyst --Pyrrhotite Pyrrhotite Fe(CO)5 Fe(CO)5 Sulfur -1g -1g 1g 2g From gas to solvent 0.229 0.301 0.320 0.510 0.187 0.213 From gas to coal 0.129 0.116 0.328 0.207 0.335 0.397 Amount of hydrogen added to coal (g) 0.264 0.335 0.330 0.348 0.354 0.415

201

[From Godo, M. et al., Energy Fuels, 11, 729 (1997)] conversion and the tritium transfer to coal. The amounts of hydrogen transferred from the gas to the solvent and from the gas to the coal, and the amount of hydrogen incorporated into the coal, are listed in Table 4.4. The amount of hydrogen incorporated into coal increased with the addition of catalysts and sulfur. In the absence of a catalyst (Runs 7 and 8), and in the presence of pyrrhotite (Runs 9 and 10), the amounts of hydrogen transferred from the gas to the coal in the presence of sulfur were less than those in the absence of sulfur. However, the amount of hydrogen incorporated into the coal in the presence of sulfur was greater than when sulfur was absent. It is therefore likely that sulfur promotes the hydrogen addition from the solvent to the coal. On the other hand, in the case of Fe(CO)5, the amounts of hydrogen transferred from the gas to the solvent were less than those when pyrrhotite was used, whereas the amounts of hydrogen transferred from the gas to the coal were greater; moreover, the amount of hydrogen incorporated into the coal increased significantly. This showed that Fe(CO)5 was effective in transferring the hydrogen from the gas to the coal, and suggests that the iron sulfide generated from Fe(CO)5 was dispersed successfully on the coal particles and was more effective in transferring the hydrogen from the gas to the coal than in transferring it from the gas to the solvent. In order to trace the behavior of added sulfur, the reactions were conducted using 35Slabeled sulfur. Fig. 4.9 shows the 35S distributions in the Taiheiyo coal liquefaction. In Run 4, using pyrrhotite and sulfur but no coal, 9% of the added sulfur was transferred into the catalyst after the reaction, corresponding to 5% of the sulfur atoms in the pyrrhotite. This shows that a sulfur e x c h a n g e reaction occurred b e t w e e n the added sulfur and pyrrhotite. In the presence of coal (Run 10), it is likely that a similar sulfur exchange would occur. In the presence of Fe(CO)5 and 1 g of sulfur (Runs 5 and 1 1), most of the 35S was distributed into the THFI fraction. The amount of sulfur used (lg) was not enough to produce H2S. In the presence of Fe(CO)5 and 2 g of sulfur (Runs 6 and 12), half of the 35S was distributed into the THFI fraction, and almost all the remaining 35S was distributed into H2S; the distribution into the solvent and the coal was negligible. The pyrrhotite catalyst is a nonstoichiometric iron sulfide which has a number of vacancies or defects on the catalyst surface. A possible mechanism for the sulfur exchange reaction is shown in Fig. 4.10. The added sulfur produced [35S]H2S, which was then assumed to dissociate into H and 35S-SH groups (SH group labeled by 35S) on the surface of the pyrrhotite catalyst. 32S-sulfur in pyrrhotite would generate 32S-SH, which would be in a state of equilibrium with [32S]H25. The H and SH groups are intermediates in the sulfur exchange reaction; this hydrogen atom may contribute to the promotion of the hydrogen trans-

202

4 Liquefaction of Coal Run Coal Catalyst Sulfur -10 + --+ 12 + Pyrrhotite 1 g Pyrrhotite 1 g Fe(CO)5 Fe(CO)5 Fe(CO)5 Fe(CO)5 1g 2g 1g 2g 0 20 40 60 80 100 % II [-1 I VA HzS Solvent Coal (THFS) Coal (THFI) + Cat.

Fig. 4.9 35S distribution in Taiheiyo coal liquefaction. [From Godo, M. et al., Energy Fuels, 11,729 (1997)]

fer into tetralyl radicals and coal radicals. This would increase the amount of hydrogen exchange and addition via the radical reaction mechanism.
[35S]H25 [35S]HzS

[35S]SH~

H, [
'~

[35515H. ~

([3'S]Fe'-xS., ~..--...._.._.__..._-

Fig. 4.10 Possible mechanism of sulfue exchange. [From Godo, M. et al., Energy Fuels, 11,729 (1997)]

4,3.3

Non Fe-based Catalysts

H-coal method, by which higher yield of gasoline cut can be obtained with one-stage process by using a Co-Mo/A1203 catalyst to be used for the petroleum desulfurizafion, can be thought about to be a promising process if the loss of the catalyst can be prevented (Comolli et al.,, 1982). Molybdenum is one of higher active metals and is expensively studied. Some advances in recent research are introduced. Sakanishi et al. (1996) reported a result obtained in coal liquefaction using Ni-Mo catalyst supported on a hollow carbon micro particles (Ketjen Black: KB). The example of the result is shown in Fig. 4.11. When the catalyst, which is 3wt% of the coal, was added to the coal, oil yield obtained at 440 ~ for 60 min reached 54%, indicating Ni-Mo/KB catalyst was more active than the commercial Ni-Mo/A1203 catalyst and the synthesized FeS2 catalyst. The surface area of KB is 1270 m2/g and greater than that of A1203, Thus the active metal can more readily disperse on the former than on the latter. Further, the NiMo/KB catalyst is also suitable to upgrading of the primary liquefied oil of the coal (Sakanishi et al., 1997). Zhang et al. (1994) compared a Co-Mo catalyst supported on an alumina carrier, the surface of which was modified with TiO2 and ZrO2, or carbon by the thermal cracking of

4.3 Catalysisin Coal Liquefaction 60 50 40


r~ "O
- -

203

O"

30 20 10
, IFI
,

.,..~

00

40 60 80 100 120 140 Reaction time (min) A Gas, 9 Oil, 9 AS, [] PA, 9 Residue

_..

Fig. 4.11 Effectof reaction time on liquefaction of Wyoming coal in the presence of Ni-Mo/KB catalyst at 440 ~ and under 13 MPa of H2 (Tetralin/Coal = 1.5, Catalyst/Coal -- 3 wt%). [Reproduced with permission from Sakanishi, K. et al., Energy Fuels, 10, 218 (1996)]

cyclohexane, with a commercial Co-Mo/A1203 catalyst in order to investigate the carrier effect. There is almost no difference among the catalysts, and the effect modified in the surface by TiO2 cannot be recognized. This was attributed to that catalytic activity in the coal liquefaction depended on surface area, and the micropore structure such as pore volume. Thus although the surface nature of the prepared catalysts was modified, the micropore structure remained still no change due to without the modification for the support. Tian et al. (1996) investigated the effect when Mg or Mo is added to the iron sulfide as the second element as well as the addition method. Catalytic activity was improved by adding Mo while there was no influence on the activity and the selectivity of the iron sulfide catalyst when adding Mg. Especially, the conversion and oil yield increased respectively 8% by impregnating Mo into the iron sulfide. The oil soluble or water soluble compounds as a high distributed catalyst precursor, which is considered to be effective in the coal liquefaction, such as metal carbonyl (Ikenaga et al., 1997; Warzinski and Bockrath, 1996; Zhang et al., 1997), metal naphtenate (Yoon et al., 1997) and ammonium tetrathiomolybdate (ATTM) (Burgess and Schobert, 1996; Song et al., 1997a, 1997b; Schobert et al., 1997; Schroeder et al., 1997) has been discussed. Ikenaga et al. (1997) found that in the presence of Ru3(CO)12 or Mo(CO)6 catalysts of high catalytic activity the molecular hydrogen activated on the catalyst was supplied to the radical directly and promptly even in the existence of the hydrogen donor solvent of the redundancy as well. On the other hand, when large quantities of resolution radicals and aromatic compound existed in the system of reaction, it was found that the hydrogen from tetralin mainly participates in the reaction than hydrogen activated on the catalyst. Moreover, it was reported that the re-hydrogenation of the naphthalene formed by a dehydrogenation of the tetralin hardly happened. A Mo(CO)6-H2S system catalyst was applicable in the solvent-free liquefaction of the bituminous coal. Though Mo(CO)6 discomposed and became Mo carbide, which is not of

204

4 Liquefactionof Coal

high activity in the liquefaction of coal without a source of sulfur, the decomposition of Mo(CO)6 was promoted, and MoS2 was formed, and coal liquefaction was promoted in the hydrogen sulfide atmosphere. Thus, a solvent-free liquefaction of coal in an autoclave is possible enough by using Mo(CO)6 and it is pointed out that hydrogen consumption speed at the early stages of liquefaction is an important factor in liquefaction of coal (Warzinski and Bockrath, 1996). It was also reported by Yoon et al. (1997) that a metal sulfide catalyst of the high dispersion was formed, and that the amounts of hydrogen consumption increased in order of Mo > Co > Fe, corresponding to the increase in the conversion and oil yield when adding three-fold sulfur of the amount of stoichiometry to transition metal naphthenate (Mo, Co and Fe etc.). Burgess et al. (1996) and Song et al. (1997a) impregnated coal to the water or the water/THF mixed solution of ATTM in order to form an active phase-MoS2 in situ on the surface of the coal, and examined about the catalytic activity. Moreover, the addition of water to the ATTM/coal system improved remarkably the conversion and oil yield when liquefaction reacted by 2-step (350 ~ minutes and 400 ~ minutes) (Song et al., 1997b). This is because the surface area of MoS2 treated with ATTM and water in 350 ~ became ca. 6 times in comparison with the surface area of MoS2 treated with only ATTM. On the other hand, Oshima et al. (1997) examined the influence of the shape of Mo catalysts on coal liquefaction. The influence of a catalyst form (powder or grain-shaped) on the internal diffuse in the pore depended on the kind of the solvent when a Co-Mo/A1203 catalyst was used. Further, no superior liquefaction yield was observed in the case impregnating Mo onto coal than that obtained in the case dissolving an oil soluble Mo catalyst into the solvent, suggesting that the Mo species supported on the coal in the former case was eluted into the solvent, and was then sulfided and catalyzed the reaction just as a super-micro-particle catalyst of MoS2. In order to develop practical processes of direct liquefaction of coal, it is important to estimate the rates of hydrogen transfer during the liquefaction. A demonstration scale process is now being developed in Japan. As the reactions in this process consist of both hydrocracking by gaseous hydrogen and hydrogen transfer from donor solvents, the understanding of these reactions is regarded as significant for the process design. Generally, coal liquefaction consists of processes in which coal is thermally hydrocracked and the produced radicals are hydrogenated by hydrogen atoms supplied from the surroundings. Fig. 4.12 shows the hydrogen transfer model in which coal is liquefied in a representative hydrogen donor solvent, tetralin, under pressurized hydrogen (Kabe, 1986). In this scheme, hydrogen atoms are supplied to coal directly by tetralin (donation) and/or gaseous hydrogen by the aid of the catalyst (spillover), or hydrogen atoms in coal are also used (shuttling). Liquefied products are hydrocracked by gaseous hydrogen to give lighter products on the catalyst. At the same time, naphthalene may be hydrogenated to tetralin on the catalyst. Many traditional approaches studying the product distribution under various experimental conditions have been made (Moritomi, et al, 1983; Guin et al., 1979; Tsai and Weller, 1979; Rottendorf and Wilson, 1979; Chow, 1981; Ueda et al., 1981) and a number of attempts have been made to elucidate the mechanisms of hydrogen transfer by the use of deuterium tracer, NMR and M.S. methods (Skowronski et al., 1984; Maekawa et al., 1980). However, a few investigations using radioisotope tracer method have been made. Poutsma et al. (1982b) performed the liquefaction of Illinois coal in a bench scale flow system and investigated the behavior of tetralin during the reaction by a laC tracer technique. They found that the grafting of tetralin to molecules derived from coal occurred to some ex-

4.3 Catalysis in Coal Liquefaction

205

3H-Donation / z - - ~

H-Shuttling 3 H H-Spillover

ng

Rehydrogenation Fig. 4.12 Hydrogentransfer model in coal liquefaction. [FromKabe. T., J. Jpn. Petrol. Inst., 39 345 (1986)]

tent. The 3H and 14C tracer methods allow a study of tritium incorporation by the addition and exchange into the coal products from the gas phase as well as from the solvent (Kabe et al., 1983a, 1983c, 1985). There are various interpretations for the effects of solvent and catalyst on coal liquefaction. One of them is that the catalysts are effective in regenerating donor solvents which are dehydrogenated during the liquefaction. From this point of view, Moritomi et al. (1983) studied the hydrogen transfer at the initial stage of liquefaction. Another interpretation that solvents are effective only to disperse coal and catalyst particles and to dissolve gaseous hydrogen and then catalysts act to hydrogenate coal and coal products by the use of gas phase or dissolved hydrogen (Ueda et al., 1981). In order to elucidate the behavior of hydrogen in coal liquefaction, the liquefaction of Taiheiyo coal by use of 3H and 14C double-labeled toluene solvent (p-3H-toluene and methyl-lac-toluene), and 3H and 14C double-labeled tetralin solvent (3H labeled tetralin and a small amount of 1-14C-naph thalene or ~4C labeled tetralin) has been studied (Kabe, 1986). Coal liquefaction was performed in a 350 ml autoclave containing 75 g of solvent, 25 g of coal and 0 or 5 g of Ni-Mo-A1203 catalyst and filled with hydrogen at an initial pressure of 0 or 5.9 MPa. It was heated at a heating rate of 10 ~ and held at 400 ~ during the reaction. The reaction mixtures in the autoclave were separated by filtration, distillation and extraction, as shown in Fig. 4.13. In this diagram, naphtha, light oil and SRC were the distilled fractions under 200, 204-350 and over 350 ~ respectively. As for the solvent extractions, HS, BS and PS represent hexane, benzene and pyridine soluble products, respectively; HIS, BIS and PIS represent the insoluble ones. SRC was separated into HS (oil). BS-HIS (asphaltenes) and PS-BIS (preasphaltenes). The separated gas, coal products and solvents were weighed and analyzed by GC. Naphthalene, decalin, methylindan and butylbenzene produced during coal liquefaction are thought to be converted from tetralin solvent, and the recovered yields of the solvent containing them ranged from 97 to 98 wt%. The specific activities of 3H and 14C in the reaction products were measured with a liquid scintillation counter. Colorless or light colored products were directly dissolved into the scintillator, while the colored liquid, solid and gas were oxidized to H20 and CO2 to avoid

206

4 Liquefaction of Coal

color quenching. Figure 4.13 shows the separation process of products; material balances of products, 3H and 14C were obtained in the flow of separation process as shown in Fig. 4.13. Specific radioactivities of 3H and ~4C of solvents and coal products in Fig. 4.13 suggest that 3H was transferred from tetralin to the reaction products, but, except tetralin, there was little transfer of 14C during the liquefaction. The 14C transferred to the liquefaction products was only 1.6% of the total ~4C, even in another experiment at 440 ~ thus it is assumed that incorporation of the solvent molecule to coal products scarcely occurs. The yield of hydrogenation of naphthalene can be evaluated using the amount of 14C transferred from naphthalene to tetralin. The amount of 14C which was contained in tetralin after the liquefaction with the catalyst at 400 ~ for 30 min was about 1.2% of total 14C, and the remaining ~4C was left in naphthalene. The re-hydrogenation of naphthalene to tetralin, therefore, was slight during the primary liquefaction step within 30 min, and the hydrogenation-dehydrogenation reactions between tetralin and naphthalene was also slight. When the liquefaction was conducted at 440 ~ and for 120 min, the amount of 14C transfer was much larger (18%) (Kabe et al., 1983a), suggesting faster re-hydrogenation of naphthalene to tetralin at 440 ~ than at 400 ~ The hydrogen addition and exchange reactions between tetralin and gaseous hydrogen without coal were also examined using 3H-labeled tetralin solvent and unlabeled gaseous hydrogen under the same liquefaction conditions. In the absence of coal, 12% of the tetralin was hydrogenated to decalin by gaseous hydrogen and 3% tetralin was converted into naphthalene at 400 ~ and 30 min. The observed 3H distribution, 84.4% in solvent and 15.6% in gas, agreed with the calculated value, 16.0% assuming the complete scrambling of
Coal (25.02 g) + 3H-Tetralin (75.18 g) + H2 (1.34 g) + Catalyst (5.08 g) + ~4C-Naphthalene (0.20 g)
I

3H: 63000 dpm/g E 14C:32500 dpm/g

Gas 50 kg/cm2 (9.36 g as H20) E total 142600 dpm 0 dpm


I I

I
Liquid Solid I Benzene Ext.

,
Light Oil (2.55 g) E 22300 1200 SRC (11.15 g) E 9300 1500

I
BIS I Pyridine Ext.

Naphtha (2.34 g) E 28500 0

Solvent (68.01 g) ] Tetralin (60.47 g) E 58200 800 Naphthalene (7.54 g) E 40100 299000

I
PS-BIS (0.48 g) 15700 3400

I PIS (15.45 g)
I

HS (4.44 g) I- 9900 L 2100

BS-HIS (4.88 g) r-" 9200 L 1100

PS-BIS (1.83 g) I-- 7700 L 500

Residue (6.34 g) 4400 E 300

Ash (4.03 g)

I Catalyst (5.08 g)

Catalyst, Ni-Mo-A1203; Initial H2, 5.9 MPa; Reaction temperature and time, 400 ~ and 30 min. Fig. 4.13 Balances of material and radioactivity of 3H and ~4C in Taiheiyo coal liquefaction at 400 ~ and under initial pressure of H2 of 5.9 MPa for 30 min in the presence of Ni-Mo-A1203 catalyst. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 346 (1986)]

4.3 Catalysis in Coal Liquefaction

207

hydrogen atoms between solvents and gaseous hydrogen. When the same experiment was conducted without a catalyst, neither hydrogen addition to a solvent from gas phase n o r 3H exchange was observed. Under the experimental conditions at 400 ~ and 30 min, therefore, the disproportionation of tetralin to naphthalene and decalin is less than 3% of tetralin in the presence of the catalyst and 0% in the absence of catalyst. Figure 4.14 shows the effects of reaction time and the catalyst on the product distributions for the experiments using 3H labeled tetralin and unlabeled gaseous hydrogen. The product distributions indicate that the catalyst enhances the hydrocracking of preasphaltenes to oil and asphaltenes. Fig. 4.15 shows the amounts of 3H transferred to the same products as in Fig. 4.14. The amount of tritium in gaseous hydrogen is significantly smaller than the equilibrium value, 13%, of tritium exchange among gaseous hydrogen, coal and solvent. Moreover, even in the presence of the catalyst, this value does not increase in the reaction, thus coal or coal liquids may inhibit the hydrogen exchange reaction between gaseous hydrogen and tetralin on the catalyst. Fig. 4.16 shows the concentrations of naphthalene produced during coal liquefactions in the presence and absence of the catalyst. In the presence of the catalyst, the concentrations of naphthalene formed in only tetralin or in the presence of coal products (HS and BS-HIS) are also plotted in Fig. 4.16. In the presence of coal without the catalyst, the concentration of naphthalene increased dramatically. When the catalyst was introduced, however, the concentration of naphthalene was always lower than that without catalyst. On the other hand, when the experiment was conducted without both gaseous hydrogen and catalyst, the amount of residue was the same as that in the presence of gaseous hydrogen without catalyst (Kabe et al., 1983c). The results suggest that the hydrogen atoms used in liquefaction are largely supplied by tetralin solvent in the primary liquefaction of coal in the presence or absence of the catalyst, but the supply of hydrogen
~., 1 0 Without catalyst 0 ~ With catalyst x x ~ Naphtha

lOOi

"~ so
~ 6o

x ~

~.~ 80 ~ ~,
O

zx Light oil Oil

4o)
-.~ 0

~, 60

40

~ 2o
i i 30 60 90 120 Nominal reaction time (min) 9 : Unconverted Coal, O : I D +Oil, I I I 30 60 90 Reaction time (min) ID : ~ + Asphaltene, : A + Naphtha i 120

300

~ : 9 + Preasphaltene, A : O + Light Oil,

Reaction temp. : 400 ~ Initial H2 pressure: 60 kg/cm 2 Fig. 4.14 Effect of reaction time on the product distribution for Taiheiyo coal liquefaction at 400 ~ and under initial pressure of H2 of 5.9 MPa. [From Kabe. T., J. Jpn. Petrol. Inst., 39 347 (1986)]

208 Without catalyst


1O0

With catalyst 10C o .... 9C I o ~

90

80
-,~

5
...

30

60 90 120 300 Reaction time (min) O 9Tetralin, | 9 SRC, [] 9Gas,

30

60 90 Reaction time (min) A 9Light oil

120

9Naphtha,

Fig. 4.15 Effect of reaction time on the tritium distribution for Taiheiyo coal liquefaction at 400 ~ and under initial pressure of H2 of 5.9 MPa. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 348 (1986)]

151
!

~=,~ l O ~

-S Z

[]

5,-

30

60 90 Reaction time (min)

120

E-l, In the presence of coal, catalyzed; II, In the presence of coal, uncatalyzed; O, In the presence of oil (HS); O, In the presence of asphaltene (BS-HIS); A, In the presence of solvent (tetralin); Catalyst, Ni-Mo-A1203; Initial H2, 5.9 MPa; Reaction temperature, 400 ~ Fig. 4.16 Effect of reaction time on production of naphthalene in tetralin solvent at 400 ~ and under initial pressure of H2 of 5.9 MPa in the presence of Ni-Mo-A1203 catalyst. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 348 (1986)]

4.3 Catalysisin Coal Liquefaction

209

atoms from tetralin is depressed in the hydrocracking steps of HS and BS-HIS as shown in Fig. 4.16. This means that the hydrogen molecules dissociated on the catalyst are used in the secondary hydrocracking of coal liquids. In addition, the concentrations of decalin remained constant throughout the reaction time, but the concentrations of methylindan and butylbenzene increased gradually, regardless of the presence or absence of the catalyst. This indicates that decalin is an impurity in tetralin; methylindan and butylbenzene are produced from tetralin during the reaction. Figure 4.17 shows the product distributions of the secondary hydrocracking of the primary coal liquids: HS, BS-HIS, PS-BIS and unseparated SRC, after the reaction for 120 min at 400 ~ in tetralin under pressurized H2 and in the presence of the catalyst. HS and BS-HIS mainly produce light oil and HS, respectively. It is shown that in the presence of BS-HIS, the hydrocracking of HS to light oil is somewhat depressed. PS-BIS is hydrocracked to produce BS. From another experiment in which the reaction time was varied, it was shown that the amounts of HS and light oil remained almost constant during the reaction of PS-BIS and, at the same time, PS-BIS decreased and BS-HIS was produced at a constant rate. From these results, it is assumed that the presence of PS-BIS depresses further hydrocracking of BS-HIS or HS (Kabe et al., 1984a). The components of the source SRC are shown in Fig. 4.17. Calculation can be made to elucidate what composition should be given when each component of SRC is hydrocracked independently and put together in proportion to the source composition. The calculated result is shown by SRCcalc and compared with the result of the hydrocracking of the unseparated SRC (source SRC). A comparison shows the following results: (1) The presence of PS-BIS in the system largely depressed the hydrocracking of BS-HIS to HS, and (2) the hydrocracking of PS-BIS takes precedence over that of BS-HIS. These results show that the heavier coal products are more adsorbable on the catalyst and have the priority to react there. 0 50
I I I I I I 9 I I I I

100

.s
BS BIS

I I

BS

HS

'iL
L N ill N

i
BIS

BIS

BS

i
HS

HS

SRC

IL I

BS

i
HS

L fll

SRCcalc

li i S~ SRC I BIS I
I BS BIS

BS

i I

L I

iI N
I I

HS

N, Naphtha; L, Light oil; HS,HS" BS, BS-HIS; BIS, PS-BIS; SRCcalc, Calculated from the data of HS, BS-HIS and BIS and these contents in source SRC. Fig. 4.17 Productdistribution in liquids from hydrocracking of coal. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 349 (1986)]

210

4 Liquefaction of Coal

Figure 4.18 shows the distributions of products and 3H after the liquefaction in the system of 3H-labeled gaseous hydrogen and unlabeled tetralin at 400 ~ for 30 min with or without a catalyst. In the presence of the catalyst, the amounts of 3H incorporation into liquid products are larger than in the absence of catalyst. These increases in incorporation are due to the hydrocracking of the liquid products using gaseous hydrogen and the hydrogen exchange between gaseous hydrogen and coal products. The effect of solvents during the liquefaction was also examined using toluene and naphthalene solvents in 3H-labeled gaseous hydrogen. The distributions of products and 3H are shown in Figs. 4.19 and 4.20. Toluene was used as the sample of a solvent which is less active as a hydrogen donor, and naphthalene was used as the sample which can change to hydrogen donor, namely tetralin. Comparing Figs. 4.19 and 4.20 with Fig. 4.18, toluene and naphthalene are shown to be inferior solvents for the liquefaction without catalyst, but they give good results with a catalyst. In the case of toluene solvent, the amount of 3H transferred to solvent is the least among the three solvents, but 3H transfer to coal products in the presence of a catalyst is the largest. In the case of naphthalene solvent, the amount of 3H transferred to solvent is larger compared with the case of tetralin solvent with the catalyst. It is concluded that, in toluene solvent, hydrogen atoms used for the coal liquefaction are supplied from the gas phase directly, but in naphthalene solvent, they are supplied from the gas phase partly through the solvent. It is assumed that tetralin produced by the re-hydrogenation of naphthalene is dehydrogenated immediately by coal and converts to 3H-containing naphthalene. The hydrogen transfer cycle of naphthalene --~ tetralin ~ naphthalene is very effective in the liquefaction in naphthalene solvent. That is why 3H content in naphthalene is high, compared with the case of tetralin solvent and also why the 3H contents in coal products are less than in the case of toluene solvent. These results agree with those of Moritomi et al. (1983) and other workers (Guin et al., 1979; Tsai and Weller, 1979; Rottendorf and Wilson, 1979; Chow, 1981)
Catalyst Ni-Mo-AI203 None 3H Ni-Mo-A1203 / Products 0 10 20 30 40 50 60 70 80 90 % 100

None

m
Residue: Solvent:

T L N ~

i
PS-BIS: ~ Naphtha: ~ ] N: Naphthalene. BS-HIS: ~ Gas: T: Tetralin. HS:

Light oil: ~ I

Fig. 4.18 Products and 3H distributions in coal liquefaction in tetralin solvent at 400 ~ under initial pressure of 3H labeled H2 of 5.9 MPa for 30 min. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 349 (1986)]

4.3 Catalysis in Coal Liquefaction Products 0 10 % 100

211

Catalyst

20

30

40

50

60

70

80

90

Ni-Mo-A1203 None 3H Ni-Mo-A1203 None Residue: ~ Light oil: ~ PS-BIS: ~ Naphtha: ~ BS-HIS: ~ Gas: ~ HS: ! ~ Solvent: [ I

Initial H2, 5.9 MPa; Reaction temperature and time, 400 ~ and 30 min. Fig. 4.19 Products and 3H distributions in coal liquefaction in toluene solvent under 3H-labeled H2. [From Kabe. T., J. Jpn. Petrol. Inst. 39, 350 (1986)] Catalyst Ni-Mo-A1203 None 3 H
1 I m

Products 0 10

20

30

40

50

60

70

80

90

% 100

Ni-Mo-A1203

None Residue: ~ Light oil: ~ Solvent: PS-BIS: ~ Naphtha: ~ BS-HIS" ~ Gas: ~ HS: Solvent: I I

I~1 N: Naphthalene, T: Tetralin.

Initial H2, 5.9 MPa; Reaction temperature and time, 400 ~ and 30 min. Fig. 4.20 Products and 3H distributions in coal liquefaction in naphthalene solvent under 3H-labeled H2. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 350 (1986)] The concentrations of 3H in coal products and solvents given are s h o w n in Fig. 4.21. In the case of n a p h t h a l e n e solvent, 3H concentration in tetralin is the largest in the solvent fractions and corresponds to the value which is given w h e n naphthalene is h y d r o g e n a t e d by two 3H labeled h y d r o g e n molecules. On the other hand, 3H concentrations in u n c h a n g e d solvents are small. This means that the h y d r o g e n e x c h a n g e b e t w e e n solvents and h y d r o g e n m o l e c u l e s is m i n o r and that the a m o u n t of the h y d r o g e n e x c h a n g e is less than a few percent of the solvent h y d r o g e n atoms even in the presence of the catalyst. H o w e v e r , 3H concen-

212

4 Liquefaction of Coal 25

20

15

,':2 Z

9 ~ rj
o

10

Residue

t PS-BIS

BS-HIS

t HS

Light oil

Tetralin

t Toluene Decalin

Naphthalene

Liquefied products

Solvents

Fig. 4.21 Tritium concentrations in liquefied products and solvents. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 350 (1986)]

trations in coal products are much larger and, in the case without catalyst, follows the order toluene > naphthalene > tetralin, indicating the reverse order of the hydrogen donating qualities of these solvents. The 3H concentrations in coal products follow the same order in the presence of the catalyst. It suggests that, in toluene solvent, the liquefaction proceeds with hydrogen spillovers from hydrogen adsorbed on the catalyst, but, in naphthalene solvent, coal is liquefied by hydrogen donation from tetralin converted from naphthalene. In the absence of a catalyst, 3H concentrations in heavier coal products are larger than that in lighter coal products. This agrees with Skowronski's results (1984). The reason for this is not clear but it seems that the heavier products have much more mobile hydrogen atoms such as in OH or NH, which exchange with hydrogen molecules even without a catalyst. From the data above, the amount of hydrogen transferred by addition or exchange reactions among gaseous hydrogen, solvent and coal products can be calculated. The calculated results are shown schematically in Fig. 4.22, in which the solid and dotted arrows indicate the directions of hydrogen addition and exchange among shown phases, respectively. The data presented in Fig. 4.22 illustrate that the catalyst slightly decreases the hydrogen addition from solvent to coal but it considerably increases hydrogen addition from gas phase to coal products. The latter increased with reaction time. The increase in hydrogen consumption is assumed to be equal to the amount used for the hydrocracking of the coal liquids, because the yields of the liquefaction, which are calculated from the amounts of unconverted coal, are almost the same in both cases with and without the catalyst. When coal and/or heavy coal liquids are present, however, the hydrogenation of solvents by gaseous hydrogen is rather slight even in the presence of the catalyst compared with the case of hydrogenation of solvents in the absence of coal. In other words, the hydrogen exchange between gaseous hydrogen and solvents is suppressed since solvents cannot easily be adsorbed on the catalyst in the presence of the heavier coal products. The hydrogen exchange between gas phase and coal products, on the contrary, proceeds considerably even in the absence of catalyst. This result means that there is much transferable hydrogen in the coal

4.3 Catalysis in Coal Liquefaction Gas phase Without 0.01 /,/'~1.22 , catalyst r . . . . 0.01 .... ~ Coal = Solvent 0.88 Gas phase 9' ~ 0"60/.,'100080 ',~ 0.04 p(~," 0.36 _ ~ Coal ~ Solvent 0.76 3H-Tetralin H2 Gas phase Gas phase

213

0"20j~,,'~1.20 0"24///9' ,' 1.32 yr . 0.40 yr '0.36 Coal = -~ Solvent C o a l . . . . . . . . ~ Solvent 0.76 0.04 Gas phase 9' 0.88~,,,~.60 0.84~',~0.04 <0.90/,.,,061050 ',~1.06 p,r 0.40 _ ~ ,r " > 0 . 3 6 _ ~ Coal ~ Solvent Coal ~ Solvent 0.68 >0.70 Tetralin Tritium Naphthalene Tritium Hydrogen weight (g) Coal (100 g) Gas phase

With catalyst

Solvent Gas phase

Reaction temperature and time, 400 ~ and 30 min. = Hydrogen exchange, -- Hydrogen addition Fig. 4.22 Diagram of hydrogen transfer in Taiheiyo coal liquefaction at 400 ~ for 30 min. [From Kabe. T., J. Jpn. Petrol. Inst., 39, 351 (1986)]

products which are easy to exchange with gaseous hydrogen but are difficult to exchange with hydrogen atoms in solvent molecules. The mechanism of hydrogen transfer during coal liquefaction was determined using 3H and 14C tracers and the following results were obtained. (1) In the process of primary liquefaction, coal is thermally decomposed and hydrogenated by the hydrogen atoms released from hydrogen donor solvent. (2) In the process of the hydrocracking of primary coal liquids, the coal liquids are hydrocracked on the catalyst by hydrogen atoms supplied mainly from gaseous hydrogen. (3) The hydrogen exchange between solvent and gaseous hydrogen during coal liquefaction is little even in the presence of a catalyst. (4) The adsorption of the heavier coal liquids are more easily on the catalyst surface, thus the hydrocracking of these heavier products take precedence over that of the lighter products. (5) In naphthalene solvent, only a small amount of coal is liquefied in the absence of catalyst because hydrogen transfer from both solvent and gaseous hydrogen is difficult. However, the liquefaction in naphthalene in the presence of the catalyst proceeds much faster, because of the hydrogen transfer cycle of naphthalene -+ tetralin -+ naphthalene due to the hydrogenation of naphthalene in the presence of a catalyst. (6) The distinction between hydrogen addition and exchange among gas phase, solvent and coal products can be made clear by these series of experimental methods using radioisotope tracers. In order to elucidate the nature of hydrogen transfer among gas, coal and solvent, 3H and 14C tracer experiments were also carded out using Datong coal (Kabe et al., 1990c). Fig. 4.23 shows the effect of liquefaction time on the conversion and product distribution for Datong coal at 400 ~ in the absence and presence of Ni-Mo/A1203 catalyst. Clearly, the initial liquefaction rate was rapid, and nearly one third of the coal was converted when the autoclave was held at 400 ~ for only a few minutes, and the initial rate did not depend on the catalyst. The conversion reached two thirds after 5 h with or without the catalyst.

214 30

4 Liquefaction of Coal

301

(b)
0

60 ~
r

~,~6 2 0

o 20

4o ~
,.o ~

~ 10x

~ 20 ~

.o .~

~ 10 ~

20 ~
0

~ O
1

-----------D0 2 3 4 Reaction time (h) 5 0 0 1 2 3 4 Reaction time (h) 5 0

Fig. 4.23 I-] Gas, X Naphtha, A Ligth oil, O HS, V HIS-BS, V BIS-THFS, 9 THFIS. Changes in product distributions with reaction time for Datong coal liquefaction at 400 ~ a) without, and (b) with a catalyst. [Reproduced with permission from Kabe. T. et el., Fuel Proces. Technol., 25, 48, Elsevier (1990)]

However, the catalyst changed the product distribution by accelerating the hydrocracking of preasphaltenes at first and then hydrogenate asphaltenes to oil. These results suggest that the catalyst was not able to hydrogenate the coal directly but that the heavier products in liquids are predominantly adsorbed on the catalyst surface and react with hydrogen. The same results were observed for Taiheiyo and Wandoan coals (Kabe et al., 1986a, 1987a). Besson et al. (1986) reported more detailed studies on the effects of Ni-Mo/A1203 and FeeO3 catalysts on coal liquefaction at 400 ~ in tetralin solvent and reached a similar conclusion, i.e., that the catalysts had little effect on the conversion of the coal into THF-soluble products but increased the amount of asphaltenes (toluene soluble products). The hydrogen distributions among gas phase, solvent and coal products are shown in Fig. 4.24. Without a catalyst, the hydrogen distribution does not change with reaction time. When a catalyst was added, hydrogen and coal products increased and that of gas decreased. In order to clarify the amount of hydrogen transferred to coal products, the amounts of hydrogen transferred from the gas phase and solvent to coal products were plotted with elapse of time in Fig. 4.25. Without the catalyst, the amounts of hydrogen initially transferred to 100 g of coal were 0.21 g from the gas phase and 0.42 g from the solvent, respectively. The hydrogen transfer from solvent to coal was the main reaction in the initial stage. The amount of hydrogen transferred gradually increased and reached 1.1 and 1.0 g from the gas phase and the solvent at 5 h, respectively. With the catalyst, the amounts of hydrogen transferred to coal products were 0.73 g from the gas phase and 0.28 g from the solvent at 0 h. This indicates that hydrogen transfer from gaseous hydrogen to coal products was rapid, even at the initial stage of the reaction when the catalyst was used. The amount of hydrogen transferred from gaseous hydrogen to coal products increased to 2.2 g at 5 h, however, that from the solvent did not increase so much and was only 0.69 g at 5 h. Since naphthalene produced by hydrogen transfer from tetralin to coal was re-hydrogenated to tetralin on the catalyst, the amount of hydrogen transferred from the solvent apparently did not increase. Total amounts of hydrogen transferred to coal were 2.1 and 2.9 g without and with the catalyst, respectively. Fig. 4.24 also shows tritium distributions among gas

4.3 Catalysis in Coal Liquefaction

215

phase, s o l v e n t and coal products. W i t h o u t the catalyst, the tritium distribution s h o w s that tritium transfers to coal p r o d u c t s in the initial stage, then tritium in s o l v e n t increases. This suggests that, as a first step, gas p h a s e tritium transfers to coal f o l l o w e d by the e x c h a n g e b e t w e e n coal and solvent. W h e n the c a t a l y s t w a s added, tritium w a s t r a n s f e r r e d to coal m o r e r a p i d l y at the initial stage o f the reaction. T r i t i u m w a s also t r a n s f e r r e d to s o l v e n t at the initial stage, in contrast to the case w i t h o u t the catalyst. A s s h o w n in Fig. 4.25, h y d r o Conversion (%) 32.2 44.9
I I

Conversion (%) 67.2


I

63.4
I

53.9 66.8
il I

75.0
I

80.4
I

80 A .o
.ID .,..~

L
-

(b) a A

'

= 60

.o ..a "~ 4 c~
.~

j
/

~ 40 _ 20

__

'

2 3 4 Reaction time (h)

2 3 4 Reaction time (h)

Fig. 4.24 Changes in balance of hydrogen and tritium in Datong coal liquefaction at 400 ~ (a) without and (b) with a catalyst. Hydrogen distribution: D gas, A solvent, C) coal; Tritium distribution: 9 gas, A solvent, 9 coal. [Reproduced with permission from Kabe. T. et el., Fuel Process. Technol., 25, 49 (1990)]

2
o c..) e~0

.,..~

~D o (

~__A____._--~

-------~-

01

Reaction time (h) Fig. 4.25 Changes in amounts of hydrogen addition with reaction time in Datong coal liquefaction at 400 ~ H added to coal from gas phese: C) without a catalyst, @ with a catalyst; H added through tetralin: A without a catalyst, A with a catalyst. [Reproduced with permission from Kabe. T. et el., Fuel Process. Technol., 25, 49, Elsevier (1990)]

216

4 Liquefaction of Coal

20 .o 10

(a) 300 ~ 300min 440 ~ 180min 230 ~ 120min 9

.~ 20

(b)
440 ~ (0-'- 300 min)

|
o

~
o

10-

E
I I I

20

40 60 Conversion (%)

80

100

20

40 60 Conversion (%)

80

100

Fig. 4.26 9 Residual coal, V Preasphaltene, V Asphaltene, O Oil, F-I Solvent Change in tritium concentrations with conversion of Datong coal liquefaction at 400 ~

(a) Without, and (b) with a catalyst [Reproduced with permissionfrom Kabe. T. et el., Fuel Process. Technol., 25, 50, Elsevier (1990)]

gen was transferred to coal from tetralin to produce naphthalene at the initial stage, and tritium was transferred to the solvent through hydrogenation of naphthalene on the catalyst. From Fig. 4.24 (b), after 5 h, the tritium distribution among gas phase, solvent and coal products approached the hydrogen distribution. This indicates that hydrogen exchange reaches the equilibrium among gas phase, solvent and coal products at 5 h when the catalyst is added. Fig. 4.26 shows the change in tritium concentrations of each liquefied product and the solvent with conversion. The experimental points in Fig. 4.26 correspond to those in Fig. 4.23 (400 ~ 0-5 h), unless otherwise noted. The points with special notation in Fig. 4.26 show the results at different temperatures. The two horizontal dotted lines in Fig. 4.26 indicate the expected mean equilibrium tritium concentrations for coal, representing the radioactivity in one gram of coal at the equilibrium in the hydrogen exchange reaction. If the equilibrium among coal, solvent and gas phase (lower line), and between coal and gas phase only (upper line), were established in the hydrogen exchange reaction, the equilibrium tritium concentrations for coal would be 5 X 103 and 19 X 103 dpm/g-coal, respectively. The expected concentration for the solvent under equilibrium conditions among the three components in the system is shown by an arrow on the fight-hand side of the graphs (10 X 103 dpm/g tetralin). Without the catalyst, tritium concentrations in coal products increased with the conversion of coal. Since tritium is transferred from the gas phase to the solvent via coal in the absence of the catalyst, the tritium concentration in the solvent began to increase after about 30 wt% of coal was converted. On the other hand, when the catalyst was added, a large number of hydrogen in coal components exchanged with gaseous hydrogen at the initial stage of reaction. Hydrogen exchange reached equilibrium among gas phase, solvent and coal products at the final stage of the reaction consistent with the results given in Fig. 4.24. Figure 4.27 shows the product distribution as a function of temperature at a reaction time of 2 h without the catalyst. In the temperature range 300-440 ~ the conversion of coal increases with increasing temperature. The hydrocracking to lighter products has little chance of occurring below 300 ~ Fig. 4.28 shows the amounts of hydrogen transferred and exchanged at 300, 350 and 400 ~ in the absence of catalyst. The solid and dotted arrows indicate the directions of hydrogen transfer and exchange among shown phases, respectively. Numbers with the arrows indicate the amounts (g) per coal (100 g) of the trans-

4.3 Catalysis in Coal Liquefaction

217

ferred or exchanged hydrogen for 2 h (Kabe et al., 1990c). Since the direct hydrogen exchange between gas phase and solvent hardly occurred in the absence of coal and catalyst, it was ignored in Fig. 4.28. The amounts of hydrogen transferred and exchanged increased with rise in temperature. The amount of hydrogen transferred from the gas phase to coal doubled with a rise from 300 to 350 ~ However, the amount of hydrogen transferred from solvent to coal remarkably increased with a rise from 350 to 400 ~ These results indicate that gaseous hydrogen is the main hydrogen donor in coal liquefaction at 350 ~ and that the capability of tetralin as a hydrogen donor appeared at 400 ~ Since hydrogen in tetralin transfers to coal more rapidly than gaseous hydrogen at 400 ~ only a slight amount of hydrogen transferred from gas phase to coal would increase with a rise from 350 to 400 ~ Figure 4.29 shows the relationship between the tritium concentrations of the coal products and the reaction time at 300 ~ In all cases at 300 ~ in the uncatalyzed experiments,
50 100

40
0 0

80
0 0 n~

30

60

0 ,d

20

4o

e~O .,..~

9 .,... ,.o

lO

20 m

0 200

300 Temperature (~

I 400

9 Residue, 9 Preasphaltene, A Asphaltene, O Oil Fig. 4.27 Effect of temperature on the fractional weights of products. [Reproduced with permission from Kabe. T. et el., Fuel Process. Technol., 25, 51, Elsevier (1990)] Reaction temperature 300 ~ Gas phase 0.25~0 , ~p'(" Coal 4 ~ .08g 0.12 g ~ 0.18 g 350 ~ Gas phase 0.50 g / ~ , " " / , , " 0.58g Jcp'!" Solvent Coal ~ 0.34 g ~ 0.27 g 400 ~ Gas phase 0.62 g / " /~,," 1.86 ~p'( Solvent Coal 1.48 g pSolvent 0.65 g

Fig. 4.28 Scheme of hydrogen transfer and exchange (g/100 g coal) between three phases in the absence of catalyst. <---: Hydrogen transfer, and <--- - - ---~: hydrogen exchange. [Reproduced with permission from Kabe. T. et el., Fuel Process. Technol., 25, 52, Elsevier (1990)]

218

4 Liquefaction of Coal

the tritium c o n c e n t r a t i o n s o f r e s i d u e s and p r e a s p h a l t e n e s w e r e h i g h e r than a s p h a l t e n e s and oils. U n l i k e the results o b t a i n e d at 4 0 0 ~ the c o n c e n t r a t i o n o f tritium in the products inc r e a s e d w i t h i n c r e a s i n g r e a c t i o n time until 4 h. A f t e r that, the p r o d u c t s a p p e a r e d to be saturated w i t h tritium. H o w e v e r , the tritium c o n c e n t r a t i o n s o f tetralin w e r e v e r y low, w h i c h indicates that the h y d r o g e n e x c h a n g e r e a c t i o n o f tetralin r e q u i r e s h i g h e r a c t i v a t i o n energy. T a b l e 4.5 s h o w s the tritium c o n c e n t r a t i o n s o f the products t o g e t h e r with the tritium and ma-

O
o

~7
2

8 L)
Q

4 Reaction time (h)

9 Residue, V THFS-BIS, V BS-HIS, O HS, [] Tetralin Fig. 4.29 Tritium concentrations in coal products and tetralin in uncatalyzed reaction of Datong coal at 300 ~ [Reproduced with permission from Kabe. T. et el., Fuel Process. Technol., 25, 52, Elsevier (1990)]

Table 4.5 Tritium Concentrations and Material Balances in Datong Coal Liquefactiona at 300 ~ for 6h Tritium concentration (dpm/g) Coal product Residue Preasphaltene Asphaltene Oil Light oil Naphtha Gas Solvent Tetralin Naphthalene Methylindan Decalin Gas -2767 3443 2144 1770 5256 5993 0 m 210 4116 210 210 Tritium balance (%) 7.47 6.65 0.31 0.26 0.09 0.09 0.07 0.00 2.68 2.23 0.44 0.00 0.01 89.83 Amount of product (g) 24.60 21.94 0.83 1.12 0.45 0.15 0.10 0.01 71.61 70.31 0.96 0.10 0.24 Recovered ratio (%) 85.30 3.23 4.35 1.75 0.80 0.39 0.04 93.18 1.34 0.14 0.34

alnitial conditions: Coal 25.05 g, Tetralin 75.01 g, and H2 1.25 g. [Reproduced with permission from Kabe. T. et el., Fuel Process. Technol., 25, 53, Elsevier (1990)]

4.4 HydrogenTransferReaction in Coal Liquefaction

219

terial balances in Datong coal liquefaction at 300 ~ and a reaction time of 6 h. From these data and the hydrogen content of the residue (4.6 wt%), the ratio of exchangeable hydrogen in the residue hydrogen can be calculated and the exchanged hydrogen at 6 h (approximately equilibrium value) amounts to 7.8 atom% of hydrogen in the residue. This value may suggest the ratio of the active hydrogen such as in the form of-OH and -NH to total hydrogen in coal. This shows which hydrogen in coal is exchanged. Since oil, asphaltenes and a part of preasphaltenes are dissolved in the solvent, these are expected to undergo more rapid hydrogen exchange than the insoluble carbonaceous materials. However, the results are different. This may suggest that there is more mobile hydrogen in residue. Further, at 300 ~ only the active hydrogen in coal products is exchanged by hydrogen. Although a detailed analysis of phenolic OH group of Datong coal has not been done, a comparable analysis of bituminous coals has been reported (Pestryakoc, 1986; Maekawa, 1975). Bituminous coals which have a chemical composition of C: 75-85% and H: 5.0-5.4 wt% contain 6-12 atom% of phenolic OH hydrogen for total hydrogen. Kotanigawa et al. (1979) concluded that the exchange reaction between deuterium gas and phenol took place rapidly at 350 ~ with ZnO-Fe203 catalyst. OH hydrogen of polycondensed aromatic phenolics is able to exchange at lower temperatures.

4.4 Hydrogen Transfer Reaction in Coal Liquefaction 4.4.1 Introduction


Since the reactions involved in coal liquefaction include hydrocracking and hydrogenation by donor solvents and molecular hydrogen, a number of attempts have been made to elucidate the mechanism of hydrogen transfer occurring during coal liquefaction in the presence of solvents. For this purpose, the use of various model compounds related to coal is very effective because the reactions with those compounds are much simpler than coal liquefaction and it is much easier to trace their behavior of hydrogen. Among these model compounds, tetralin is one of the most simple, interesting and convenient model compounds because it is cheap and easier to obtain, has an aromatic ring and a naphthene ring in its structure and can serve as a hydrogen donor solvent. The hydrogen transfer from tetralin to coal molecules, and from gas phase to tetralin especially has been extensively studied to understand the mechanism of coal liquefaction (Cronauer et al., 1978, 1979; Billmers et al., 1986; Skowronski et al., 1984). The mechanisms of hydrogen transfer from a donor solvent such as tetralin to various coal structures and their subsequent fragments remain largely unspecified, notwithstanding the efforts in this area over the years. Much discussion on donor solvent in coal liquefaction is based on the presumption that the principal mechanism involves thermal scission of weak bonds, followed by capping of the resulting free radicals by hydrogen atom abstraction from donor solvent (Kuhlmann et al., 1985), even though a number of researchers have pointed out alternative possibilities (Franz and Camaioni, 198 l a; Stein, 1982). It has also been suggested that hydride transfer (Brower, 1977) and concerted H2 transfer processes (Virk, 1979) play important roles in coal liquefaction. In addition, direct transfer of hydrogen atom from solvent derived radicals to substituted positions in aromatic tings as preliminary steps in depolymerization of coal structures has also been suggested to be important in coal liquefaction (McMillen et al., 1987). In the elucidation of the hydrogen transfer, hydrogen exchange is other form of hydrogen transfer and is also extensively studied to understand the mechanisms of various processes (Benjamin et al., 1982b, 1983; Davis and Garnett, 1975; Gamett and Kenyon,

220

4 Liquefactionof Coal

1971). It has been recognized that the exchangeability of a compound is strongly relative to its molecular structure, and further to its capacity as a hydrogen donor or acceptor. Based on this viewpoint, it is possible to estimate the structural feature of coals by the investigation of hydrogen exchange. In the studies, the isotope tracer method has been extensively utilized as an effective means as well as model compounds. For example, in coal liquefaction, this method makes it possible to determine the amount and the structural positions of hydrogen in coal reacting with hydrogen during liquefaction by labeling reactive sites with deuterium or tritium. Most of these investigations were performed using deuterium tracer. Several research groups have used deuterium to investigate the mechanism of coal hydrogenations and the reaction of coal-related model compounds such as tetralin since 1967 (Fu and Blaustein, 1967; Franz and Camaioni, 1980; Brower, 1982; Schweighardt et al., 1976; Cronauer et al., 1982; Skowronski et al., 1984; King and Stock, 1982; Collin and Wilson, 1983). In these researches, intensive effort has been made to obtain a better understanding of the coal hydroliquefaction mechanism. Such knowledge may lead to the improvement of coal utilization and coal hydroliquefaction efficiency by elucidating hydrogenation rates and mechanisms as well as the sites of hydrogen incorporation from the gas and solvent phases. However, because of the low solubility of coal in solvents and the lack of quantitative data from 2H-NMR, it is difficult in these studies to conduct quantitative analysis of hydrogen transfer among the gas phase, solvent and coal. The investigation of coal liquefaction mechanisms using radioactive tritium as a tracer has started in recent years. The representative investigations are the quantitative estimation of hydrogen mobility in the systems consisting of coal and donor solvent, coal and gas phase hydrogen, and coal and water, reported by Kabe and coworkers (Kabe, 1984, 1986, 1988; Kabe et al., 1983a-c, 1984, 1985a-d, 1986a-c, 1987a-c, 1989a-b, 1990a-e, 1991a-c; Ishihara et al., 1993, 1994, 1995; Yamamoto et al., 1987). The results have shown that tritium tracer methods have several distinct advantages over the deuterium tracer methods, especially for a complicated reaction system. These reports indicate that the hydrogen mobility of coal and coal-related compounds can be quantitatively analyzed using the hydrogen exchange reactions occurring between coal, the gas phase and the tetralin solvent, as well as by considering the hydrogen addition reactions.

4.4.2 Behavior of Hydrogen in Coal Liquefaction


The use of catalytic hydrotreatment is an important aspect of the process of Wandoan coal liquefaction now being developed in Japan. Although the reaction conditions and the nature of the products may present a number of problems in the development of the process, the fundamental reactions of coal liquefaction in the initial reaction stage are the thermal decompositions of coal structure under hydrogen atmosphere in the presence of a donor solvent (Curran et al., 1967; Neavel, 1976; Derbyshire and Whitehurst, 1981). When there are hydrogen atoms, which can stabilize coal radicals made by thermal decomposition, the coal is converted to coal-derived liquids. If no hydrogen is available, however, the free radicals recombine to form heavier products (Ohe et al., 1985). Therefore, an understanding of the hydrogen transfer mechanism during liquefaction is essential for process design. The liquefaction of Wandoan coal in 3H- and in 14C-labeled solvent was studied (Kabe et al., 1987b). It was reported that tetralin solvent provided the coal with hydrogen atoms, and that the presence of a catalyst decreased the addition of hydrogen atoms from solvent to coal and increased the addition from gaseous hydrogen to coal. Since gaseous hydrogen was not traced, however, the hydrogen transfer path from the gas phase to coal remains unclear.

4.4 Hydrogen Transfer Reaction in Coal Liquefaction

221

Here, the liquefaction b e h a v i o r of W a n d o a n coal u n d e r a 3H-labeled h y d r o g e n atmosphere was investigated to clarify the role of gaseous h y d r o g e n ; the effect of solvent was also d e t e r m i n e d using unlabeled solvents such as tetralin, n a p h t h a l e n e and decalin (Kabe et al., 1987a). To d e t e r m i n e the h y d r o g e n transfer path f r o m the gas phase, liquefaction e x p e r i m e n t s were carried out using 3H-labeled h y d r o g e n gas in unlabelled tetralin, n a p h t h a l e n e and decalin. The distributions of products and tritium are s h o w n in R u n s 1-8 in Tables 4.6 and 4.7. In these tables, the decalin fraction contains decalin, 1 - m e t h y l i n d a n and b u t y l b e n z e n e . T h e decalin is verified to be an i m p u r i t y contained in tetralin, and the other two substances s e e m to be converted from tetralin in the case of tetralin solvent. In tetralin solvent, the catalyst does not e n h a n c e liquefaction yields as calculated f r o m the a m o u n t of residue, but it increases the c o n s u m p t i o n of gaseous h y d r o g e n and the hydrocracking of asphaltene (Runs 1 and 2). The catalyst does not affect the formation of decalin, but it reduces n a p h t h a l e n e formation during coal liquefaction. T h e distribution and concentrations of 3H indicate that the rate of 3H transfer f r o m gas phase to coal products in the p r e s e n c e of a catalyst is h i g h e r than that in its a b s e n c e ( T a b l e 4.7, R u n s 1 and 2). Therefore, in tetralin solvent, the catalyst p r o m o t e d h y d r o g e n transfer f r o m the gas phase to Table 4.6 Product Distribution for Wandoan Coal Liquefaction Run No. Solvent Catalyst Products (wt%) Residue Preasphaltene Asphaltene Oil Light oil Naphtha Gas Solvent (wt%) Naphthalene Tetralin Decalin b
a

Tetralin -+ 26.6 22.7 27.2 12.9 8.7 1.0 1.0 13.8 84.7 1.4 25.0 22.4 21.2 17.1 12.1 1.1 1.0 7.9 90.6 1.4

Naphthalene -+ 70.3 12.0 3.5 6.6 5.9 0.7 1.0 99.4 0.1 0.5 30.8 13.7 24.5 22.0 5.9 1.9 1.2 93.8 5.7 0.5

Decalin -+ 51.4 12.6 13.7 11.4 6.7 3.3 0.8 3.4 5.5 91.1 33.5 12.1 21.7 18.9 10.8 1.9 1.0 0.9 2.6 96.6

7a 8a Tetralin -+

0.4 99.2 0.4

2.5 85.6 11.9

Without the coal; bDecalin fraction contains decalin, 1-methylindan, and butylbenzene [Reproduced with permission from Kabe. T. et el., Fuel, 66, 1327, Elsevier (1987)] Table 4.7 Tritium Distribution for Wandoan Coal Liquefaction Run No. Residue Preasphaltene Asphaltene Oil Light oil Naphtha Naphthalene Tetralin Decalin b Sum in solvent fraction Gas phase
1 2 3 4 5

Tritium (%) 2.4 1.9 2.0 0.9 0.9 0.3 1.0 7.1 0.3 8.4 83.4 7.2 6.2 5.6 5.1 1.3 1.3 1.7 22.9 0.4 25.0 48.3 11.3 1.3 0.4 0.6 0.9 0.3 4.4 0.2 0.2 4.8 80.4 9.2 3.7 5.6 5.8 3.2 2.7 21.9 10.2 0.4 32.5 37.2 3.7 0.8 1.0 0.8 1.1 1.6 0.1 0.2 2.6 2.9 88.2 14.9 5.7 8.2 8.3 2.6 2.5 0.1 0.2 6.5 6.8 75.6
n m

m 0.0 0.0 0.0 0.0 100.0

15.3 66.8 1.0 83.1 16.9

Without the coal; bDecalin fraction contains decalin, 1-rnethylindan, and butylbenzene. [Reproduced with permission from Kabe. T. et el., Fuel, 66, 1327, Elsevier (1987)]

222

4 Liquefactionof Coal

coal products. In naphthalene solvent, the degree of liquefaction was obviously low without catalyst (Run 3). However, liquefaction proceeds at a substantial rate in the presence of the catalyst (Run 4). This suggests that gaseous hydrogen is used for liquefaction in naphthalene solvent in the presence of a catalyst. On the other hand, when decalin is used as the solvent in the presence of a catalyst (Run 6), the amount, of residue is almost the same as for naphthalene solvent but the products are lighter. Without the catalyst (Run 5), liquefaction in decalin proceeds more extensively than in naphthalene. The amounts of tetralin and naphthalene derived from decalin show that liquefaction proceeds to a considerable extent accompanying the hydrogen donation from decalin in the absence of a catalyst. The last columns (Run 8) of Tables 4.6 and 4.7 show the distributions obtained experimentally without the coal and coal products. The experimental 3H distribution, 83% in the solvent and 17% in the gas phase, agreed with the calculated values based on the assumption of a complete scrambling of hydrogen atoms of the solvent and molecular hydrogen. These results show that if coal or coal products are not present, the hydrogenation of tetralin to decalin and the hydrogen exchange between solvent and molecular hydrogen will proceed rapidly in the presence of a catalyst. On the other hand, no hydrogenation of solvent or hydrogen exchange occurs in the absence of catalyst (Run 7). Figure 4.30 shows 3H concentrations in liquefied products and in the solvent. It shows that 3H concentrations in coal products produced in tetralin and decalin solvents are lower than those in naphthalene solvent in the absence of a catalyst. But in the presence of a catalyst, the amount of 3H from the gas phase incorporated into coal products is largest in decalin solvent except for light oil. It shows that decalin is a good hydrogen donor without a catalyst, but molecular hydrogen is a better hydrogen donor than decalin when a catalyst is present. With a catalyst, naphthalene has a similar action to tetralin. In Fig. 4.30, a fairly low concentration of 3H in each solvent shows that the hydrogen exchange between solvent and molecular hydrogen is small in the presence of coal and coal products. On the other hand, the 3H concentration in tetralin converted from naphthalene is high because tetralin molecules are formed by hydrogenation of naphthalene using molecular hydrogen. The value of the 3H concentration of tetralin converted from naphthalene solvent was equal to the value calculated under the assumption that four hydrogen atoms from the gas phase are added to one naphthalene molecule. The 3H concentration of decalin fraction converted from tetralin and that of naphthalene converted from tetralin were the same as the 3H concentration in tetralin solvent itself, in the presence of the catalyst. Comparing the uncatalyzed experiments in the three solvents shown in Table 4.6, the degree of liquefaction, determined from the amount of residue, decreases in the order tetralin > decalin > naphthalene. This order conforms to that of the hydrogen donating power of the solvents. However, the order is tetralin > naphthalene -- decalin in the presence of the catalyst. This shows that the hydrogen donating cycle, naphthalene ~ tetralin ---) naphthalene in naphthalene solvent, is as effective for coal liquefaction as hydrogenation in decalin solvent in the presence of a catalyst. Table 4.7 shows that the amounts of 3H in coal products in uncatalyzed experiments (Runs 1 and 5) are almost the same in both tetralin and decalin solvents. In Run 3 in naphthalene without catalyst, the 3H content in the residue is higher than that in tetralin or in decalin solvent. In the absence of a catalyst, Fig. 4.30 shows that the amount of 3H incorporated into the products increases in the order oil < asphaltenes < preasphaltenes < residue, irrespective of the solvent used. These results agreed with those obtained by Skowronski et al. (1984) in a coal-deuterium gas system. The amount of 3H transfer from gaseous hydro-

4.4 Hydrogen Transfer Reaction in Coal Liquefaction 25

223

20

_.9.

15

~ ~Z
_ _

10

Residue

PA

I I
O

;L
LO N T D

Fig. 4.30 Effects of solvents and catalyst on tritium counts of products for Wandoan coal liquefaction at 400 ~ for 30 min. PA: Preasphaltene; A: Asphaltene; O: Oil; LO: Light oil; N: Naphthalene; T: Tetralin; D: Decalin; 1-Methyl-indan and butylbenzene [Reproduced with permission from Kabe. T. et el., Fuel, 66, 1327, Elsevier (1987)]

gen to coal products increases substantially in the presence of a catalyst. Fig. 4.30 also shows that 3H concentration of coal products is almost the same as that in tetralin and naphthalene solvents in the presence of a catalyst, and it seems to show that, in naphthalene solvent, a fairly large part of the hydroliquefaction was conducted by tetralin which was produced from naphthalene. On the other hand, 3H concentrations in coal products produced in decalin solvent are higher than those in other solvents in the presence of a catalyst. This indicates that direct hydrogenation of coal by gaseous hydrogen and the hydrogen exchange between hydrogen molecules and coal components are enhanced in decalin solvent. This also suggests that it is energetically more favorable for liquefied products to react with hydrogen dissociated on the catalyst than to be hydrogenated by the decalin itself. From these results it is concluded that naphthalene behaves as a hydrogen carrier from the gas phase to coal by being hydrogenated to tetralin with the help of the catalyst. Here the route of hydrogen incorporation from solvents and gaseous hydrogen to coal products is discussed. Hydrogen incorporation during coal liquefaction involves two reactions, i.e., hydrogen addition to coal products and hydrogen exchange among the coal products, the solvent and the gas phase. To clarify the correlation of these reactions, the author and his coworkers attempted to calculate and differentiate these two kinds of hydrogen incorporation using the experimental data of Tables 4.6 and 4.7. The results calculated for Runs 1 to 6 are shown schematically in Fig. 4.31 in which, the solid arrows show the directions of the hydrogen addition and the dotted arrows represent those of hydrogen exchange between the phases shown. The numbers with the arrows indicate the amounts (g) per coal (100 g) of the added or exchanged hydrogen. Referring to Fig. 4.31, when the catalyst is not used, coal liquefaction proceeds by hydrogen addition mainly from the solvent, except for the case of naphthalene solvent, and the gaseous hydrogen can exchange only with coal products. The catalyst enhances direct hydrogen addition from the gas phase to coal products and decreases the amount of hydrogen

224

4 Liquefaction of Coal Solvent Tetralin Gas phase Without catalyst 0.04/,.,/0 .80 //!, 0.40 Coal 1.04 Gas phase With catalyst 0.89 .79 0.3 .15 //k, 0.40 '~ ~ Coal Solvent 0.67 9 Fig. 4.31 0.6 Coal Naphthalene Gas phase 0.1 6/,.,/0 .40 //',, 0.24 Coal 0.08 Gas phase
tt /

Decalin Gas phase ' 0.1 6~,,'/0.40 //!, 0.12 Coal , 1.00 Gas phase .12 Solvent 1.36 / 0.72 0.2 .04 -. 0.12 Coal Solvent 0.48

Solvent

Solvent

Solvent

1.01 0.6 0.24 0.58

:Hydrogen addition, -,. . . . . . :Hydrogen exchange, ( Hydrogen-g ) Coal- 100g

Scheme of hydrogen exchange and addition among three phases. [Reproduced with permission from Kabe. T. et el., Report of Special Project Reserch on Energy, 16(1988)]

donation from the solvent. The hydrocracking of liquefied products and hydrogenation of the solvent are promoted by the catalyst. Tetralin is known to be an effective hydrogen-donor solvent in coal hydrogenation processes. In such processes, it is important to study the thermal behavior of tetralin. Recently, much attention has been focused on the mechanism of the pyrolysis of tetralin in the absence and presence of coal (Cronauer et al., 1978, 1979; Hooper et al., 1979; Benjamin et al., 1979; Franz and Camaioni, 1980a,b; Penninger, 1982; McPherson et al., 1985; Vlieger et al., 1984; Poutsma et al., 1982b; Yen et al., 1976, and references cited therein). For example, Hooper et al. (1979) reported that tetralin did not disproportionate to naphthalene and decalin between 300 and 450 ~ Benjamin et al. (1979) showed that the formation of 1-methylindan might be due to the cleavage of the 1-8a bond of tetralin. Franz and Camaioni (1980a,b) reported that the isomerization of tetralin to 1-methylindan proceeds through 2-tetralyl radical derived from the corresponding perester. Penninger (1982) reported that the formation of methylindan from tetralin involves a bimolecular step in the reaction with gaseous hydrogen. McPherson et al. (1985) also inferred that the mechanism of tetralin isomerization must include a multimolecular step. On the other hand, the reactivities of hydrogen in coal and tetralin have been investigated using deuterated tetralin through the reaction with coal (Franz, 1979; Franz et al., 1981; King and Stock, 1982; Skowronski et al., 1984; Collin and Wilson, 1983; Wilson et al, 1982, 1984; Cronauer et al., 1982; Schweighardt et al., 1976; Brower, 1982 and references cited therein). Cronauer et al. (1982) showed that significant hydrogen/deuterium exchange occurred between coal and deuterated tetralin. Showronski et al. (1984) clarified the role of gaseous hydrogen and donor solvent in coal liquefaction using gaseous deuterium and deuterated tetralin. However, because of the low solubility of coal to solvents and the lack of quantitative data from 2H-NMR, it was difficult in these studies to conduct quantitative analysis of hydrogen transfer among the gas phase, solvent and coal. One group has already reported that tritium and 14C tracer methods were effective in trading the reaction pathways of hydrogenations among gas phase, solvent and coal (Kabe et al., 1986a, 1987a, b, 1989a). Further, it has shown that the hydrogen exchange reaction between coal and hydrogen molecules remarkably proceeded with an increase in temperature from 350 to 400 ~ (Kabe et al., 1990b). Here, the researchers were interested in the hydroaromatic structure of tetralin itself, which can give hydrogen to coal during liquefaction. Although a number of attempts have been

4.4 HydrogenTransfer Reaction in Coal Liquefaction

225

made to elucidate the reactivity of tetralin in the presence of coal, little is known about the behavior of tetralin itself, especially the hydrogen mobility in it under coal liquefaction conditions. The reaction of tetralin with tritiated hydrogen molecules to estimate the hydrogen mobility of tetralin quantitatively using a tritium tracer method is discussed below. The reaction of tetralin with tritiated hydrogen molecule was performed at 350-400 ~ and the results are shown in Fig. 4.32a (Kabe et al., 1991a). In this temperature range, the products were 1-methylindan by isomerization, naphthalene by dehydrogenation, and nbutylbenzene by hydrocracking; the main product was 1-methylindan. With increase in temperature, the concentration of tetralin decreased and the concentrations of the products increased. Decalin by disproportionation was not produced at 350--400 ~ This is consistent with Hooper's result, in which the disproportionation of tetralin to decalin and naphthalene does not occur in the absence of coal (Hooper et al., 1979). When the reaction temperature increased to 440 ~ the yields of 1-methylindan, naphthalene and n-butylbenzene increased to 11.6, 4.8 and 5.2%, respectively, and the increase in 1-methylindan was most remarkable. That the yield of 1-methylindan remarkably increased with a rise from 400 to 2 I (a) Reaction time 120 min

r O

~.
O .,..~

0 350 375 400 Reaction temperature (~ /[~

5 ~. 4 ~ 3 O
O

(b) Reaction temperature 400 ~

~21

r-i

I-1

i 120 240 360 Reaction time (min)

i 480

Fig. 4.32 Effectof reaction temperature and reaction time on product yields in the reaction of tetralin with gaseous hydrogen (Tetralin: 75 g). /~ Naphthalene; D Methylindan; V Butylbenzene [From Kabe. T. et al., Ind. Eng. Chem. Res., 30, 1756, (1991)]

226

4 Liquefaction of Coal

440 ~ is also consistent with Hooper's report. Fig. 4.32b shows the effect of reaction time on the yields of 1-methylindan, naphthalene, and n-butylbenzene. These products increased with time. When the reaction time was prolonged from 300 to 480 min, yields of products, especially 1-methylindan, remarkably increased. Tritium in the gas phase was introduced into tetralin; and the amount introduced was estimated by the hydrogen exchange ratio (Kabe et al., 1991a). Fig. 4.33 shows the change in both the hydrogen exchange ratio and the conversion of tetralin with reaction time and temperature. At 350 ~ the hydrogen exchange ratio was 0.3% even at 300 min. The conversion of tetralin was also rather low. At 375 and 400 ~ both the hydrogen exchange ratio (open symbol) and the conversion of tetralin (closed symbol) increased with time and temperature. Some relationship seems to exist between the hydrogen exchange and the conversion of tetralin. When the reaction time was prolonged to 480 min, the hydrogen exchange ratio and conversion of tetralin at 400 ~ were 13.5 and 10.1%, respectively. When the reaction temperature increased to 440 ~ the hydrogen exchange ratio and conversion of tetralin at 120 min were 14.3 and 21.7%, respectively. It was reported that, in coal hydrogenation, the hydrogen exchange ratio of coal increased remarkably with a rise from 350 to 400 ~ to reach more than 40% at 400 ~ 120 min (Kabe et al., 1990b). The present resuits indicate that simple tetralin itself is more difficult to exchange with hydrogen molecules in gas phase than coal. To estimate the relationship between the hydrogen exchange ratio and conversion of tetralin, the values of the amount of exchanged hydrogen per amount of converted tetralin (g-atom/mol) at each temperature were plotted against reaction time in Fig. 4.34. Values at 375 and 400 ~ increased with time and gave a linear relationship. However, the values at 375 ~ were larger than those at 400 ~ at each reaction time. This shows that the hydrogen exchange between tetralin and molecular hydrogen strongly depends on temperature, although the hydrogen exchange may occur at the time when tetralin converts (vide infra). Further, the values at 375 ~ for 300 min and 400 ~ for 300 min were 25.3 and 16.5 gatom/mol, respectively, more than 12 g-atom in 1 mol of tetralin. This shows that tritium is introduced into not only converted tetralin (products), but also into remaining tetralin. These results indicate that there is an intermediate such as tetralyl radical which converts to a product or returns to tetralin and further can exchange with hydrogen molecules in those routes as shown in Eq. (4.34). On the other hand, the value of AEH/ACT at 350 ~ was somewhat constant. The mechanism at 350 ~ may be different from that at 375 and 400
~

or

"

9 products

(4.34)

Table 4.8 shows effects of the amount of tetralin and hydrogen pressure on the hydrogen exchange ratio, the conversion of tetralin, and the product distribution at 400 ~ for 120 min. The hydrogen exchange ratio and the amount of converted tetralin were plotted against the value of hydrogen per tetralin (mol/mol) in Fig. 4.35. The plots of the hydrogen exchange ratio showed approximately a straight line and increased in proportion to hydrogen/tetralin values, while the plot of the amount of exchanged hydrogen showed some scatter. The amount of exchanged hydrogen showed a maximum, about 0.3 g at 30 or 50 g of tetralin, 60 kg/cm 2. The plots of the amount of converted tetralin also showed a sure straight line and increased with increase in hydrogen/tetralin. The amounts of products formed were plotted against hydrogen/tetralin in Fig. 4.36, 1methylindan showed a straight line and increased with increase in hydrogen/tetralin. On

4.4 Hydrogen Transfer Reaction in Coal Liquefaction

227

4-

o0o.o "~176176

120

240

360

Reaction time (min) Fig. 4.33 Effect of reaction temperature and reaction time on hydrogen exchange ratio and conversion of tetralin (tetralin: 75 g). Hydrogen exchange ratio: O 400 ~ E] 375 ~ 350 ~ Conversion of tetralin: 9 400 ~ 9 375 ~ 9 350 ~ [From Kabe. T. et al.,Ind. Eng. Chem. Res., 30, 1756 (1991)]

30

20
O

d~
[-, <

10
<

0 0

i 120

i 240

i 360

Reaction time (min) Fig. 4.34 Effect of reaction temperature and reaction time on ratio of the amount of exchanged hydrogen (AEH) to the amount of converted tetralin (ACT). O 400 ~ IS] 375 ~ 350 ~ [From Kabe. T. et al., Ind. Eng. Chem. Res., 30, 1757 (1991)]

the other hand, although 2.3 g of naphthalene was produced in the absence of hydrogen as shown in Table 4.8 and Fig. 4.36, the amount of naphthalene formed was inhibited by the presence of hydrogen and showed the tendency to decrease slightly with increase in a hydrogen/tetralin molar ratio. The amount of n-butylbenzene formed also increased with an increase in hydrogen/tetralin. These results mean that the hydrogen-exchange ratio, the amount of methylindan formed, and the amount of converted tetralin increase with increase in hydrogen pressure. It is important that the gaseous hydrogen participates in the isomerization which does not accompany the income and the outgo of hydrogen. This may be one

228

4 Liquefaction of Coal Table 4.8 Yields of Products and Hydrogen Exchange Ratio a

Amt of tetralin, g 15 30 30 30 50 75 75 75 75 d

Hydrogen press., kg/cm 2 20 20 45 60 60 20 40 60 0

Hydrogen/ tetralin, mol/mol 2.70 1.21 2.70 3.69 2.14 0.42 0.83 1.25 0.00

Yield of product, c % HER, b % 4.45 3.23 5.12 10.90 6.39 0.67 1.22 2.04 Conv of tetralin, % 18.22 7.08 10.33 10.88 4.96 2.15 2.52 2.91 3.01 NP 5.78 2.58 2.45 2.36 1.07 0.85 0.80 0.81 1.70 MI 11.51 3.88 6.07 6.16 2.66 1.15 1.33 1.44 1.24 BB 1.07 0.74 1.89 2.34 1.27 0.24 0.45 0.64 0.20

a Reaction temperature, 400 ~ Reaction time, 120 min. briER = H~,drogen exchange ratio, c Np, Naphthalene; MI, Methylindan; BB, Butylbenzene. dNitrogen atmosphere (1 kg/cm~). [From Kabe. T. et al., Ind. Eng. Chem. Res., 30, 1757, (1991)]

15

-4
10-3

~0

"~
~

- 1

E <

0 0

t 1

t 2

i 3

i 4

Hydrogen / Tetralin (mol/mol) Fig. 4.35 Effect of molar ratio of hydrogen to tetralin on hydrogen exchange ratio and the amount of converted tetralin at 400 ~ for 120 min. O Hydrogen exchange ratio; 9 Amount of converted tetralin; A Amount of exchanged hydrogen (x 10-~ g in scale of amount of converted tetralin). [From Kabe. T. et al., Ind. Eng. Chem. Res., 30, 1757 (1991)]

of the reactions that make it possible for hydrogen in tetralin to exchange with the hydrogen molecule. McPherson et al. (1985) suggested that the fact that coal suppressed the formation of 1-methylindan necessitates a mechanism in which at least one step in the isomerization is multimolecular. Penninger (1982) also showed by the kinetics derived from gasphase work that gaseous hydrogen participated in the isomerization. Further, it was shown that the enhancement of ring cracking becomes less significant with increasing concentration of the hydrocarbon and that the hydrogen-initiated ring cracking is gradually transferred into a hydrogen donor mechanism as the leading reaction scheme when the concentration of tetralin is increased. A similar phenomenon was observed in the system. In Figs. 4.35 and 4.36, the relationships between the hydrogen/tetralin molar ratio and the amounts of 1-methylindan and n-butylbenzene formed or the hydrogen exchange ratios approximately follow straight lines. This means that the increase in tetralin decreases the cracking products and the hydrogen exchange. Tetralin may be activated by collision with itself or a

4.4 HydrogenTransfer Reaction in Coal Liquefaction

229

i
O O

V I I I 1 2 3 Hydrogen/Tetralin (mol/mol)

Fig. 4.36 Effectof molarratio of hydrogento tetralin on amounts of products at 400 ~ for 120 min. A Naphthalene; D Methylindan; V Butylbenzene. [From Kabe. T. et al., Ind. Eng. Chem. Res., 30, 1757 (1991)]

hydrogen molecule to produce one intermediate such as the tetralyl radical in Eq. (4.34). However, if the intermediate would be quenched by tetralin to form original tetralin, the conversion of tetralin and hydrogen exchange would be inhibited. The reaction of the intermediate with molecular hydrogen leads to hydrogen exchange in both the conversion and the reproduction of tetralin. Because molecular hydrogen promotes the conversion of tetralin, it should not quench the intermediate at least more rapidly than tetralin. The reactions of tetralin with tritiated hydrogen molecule in the presence of coal were investigated, and results are shown in Table 4.9. Although coal did not largely affect the formation of 1-methylindan, n-butylbenzene, or decalin at 300-400 ~ the amount of naphthalene formed remarkably increased, especially with a rise from 350 to 400 ~ compared with that in the absence of coal. It is suggested that the interaction between coal and tetralin is enhanced in the range of 350-400 ~ This is also observed in the variations in radioactivity in coal and tetralin (Kabe et al., 1991a). The radioactivity in tetralin remarkably increased with a rise from 350 to 400 ~ while that in coal did not change greatly. In the reaction of coal with tritiated molecular hydrogen without solvent, the radioactivity in coal increased remarkably in the range 350-400 ~ (Kabe et al., 1990b). When tetralin was added, it was assumed that tritium initially introduced into coal would be rapidly transferred to tetralin within this temperature range. On the other hand, the hydrogen exchange ratio of tetralin was much larger than that in the absence of coal as shown in Table 4.9. This indicates that coal promotes the hydrogen exchange reaction between molecular hydrogen and tetralin to introduce tritium into tetralin. King and Stock (1982) reported that the hydrogen exchange between coal and tetralin-&2 and naphthalene-d8 was readily reversible at 400 ~ and that the reactions were initiated by single-bond homolysis and by molecule-induced homolysis. Further, McMillen et al. (1987) reported that the hydrogen transfer from donor solvent to coal model compound proceeded by the radical hydrogen transfer mechanism. Billmers et al. (1986) also reported that hydrogen migration between 9,10-dihydro positions in anthracene structures was consistent with a free radical mechanism. In these reports, the hydrogen transfer processes are reversible and the hydrogen exchange reaction can occur through these radical mechanisms. In our systems, since coal

230

4 Liquefaction of Coal Table 4.9 Tritium Distribution in the Presence of Coal a Product distrib, b wt% TL 97.52 93.40 81.47 76.93 NP 2.50 6.47 16.72 19.55 MI 0.00 0.06 1.13 2.19 BB 0.00 0.06 0.62 1.18 DL 0.00 0.02 0.06 0.15 HER, % 0.05 0.60 3.60 10.10

a Tetralin, 75 g; Coal, 25 g; H2, 60 kg/cm2; Amount of hydrogen in gas phase, 1.24 g; Amount of hydrogen in tetralin, 6.82 g. bTL, Tetralin; NP, Naphthalene; MI, Methylindan; BB, Butylbenzene; DL, Decalin. [From Kabe. T. et al., Ind. Eng. Chem. Res., 30, 1758 (1991)]

generates radicals more easily than tetralin, gaseous hydrogen must be more easily activated on coal surface than tetralin. Tritium transferred into coal would readily exchange with hydrogen in tetralin through a radical mechanism. In the presence of coal, it has been reported that hydrogen in naphthalene, which was initially added, exchanges with gaseous hydrogen (Kabe et al., 1987a). Naphthalene formed from tetralin was isolated. In spite of the release of hydrogen, it contained tritium from the gas phase in the absence and presence of coal (Kabe et al., 1991 a). Two pieces of data in the absence and presence of coal were chosen because the hydrogen exchange ratios were very similar to each other. The amount of hydrogen exchanged per mole of naphthalene in the absence of coal was 1.85 g. Even in the presence of coal where a large amount of hydrogen was released, 1.00 g of hydrogen in naphthalene was exchangeable with gaseous hydrogen. It is suggested that when tetralin changes to naphthalene, hydrogen in tetralin will become very mobile and be able to exchange with molecular hydrogen. Skowronski et al. (1984) reported that, in the hydrogen exchange between tetralin-d~2 and hydrogen in coal at 400 ~ for 1 h in a shaken autoclave system, protium was incorporated into Ha (66%), H~ (23%), and Har (11%) positions in tetralin and that the H~ absorption of the recovered naphthalene in IH NMR was approximately 7 times as intense as the H~ absorption. Collin and Wilson (1983) showed from their insensitive nucleus enhancement by polarization transfer (INEPT) and gated spin echo (GASPE) NMR study that, in the reaction of tetralin with deuterium and coal, the mixture of tetralin consists of molecules that were nondeuterated and monodeuterated at Ha and/or H~ positions while no evidence was found for any molecules that were dideuterated at Ha and/or H~ positions. In their NMR measurement, the intensity of the signal at the H,~ position was larger than that at the H~ position (Collin and Wilson, 1983). Franz and Camaioni (1980), in a set of pyrolysis experiments with peresters of 1-tetralyl, 2-tetralyl, l-indanylmethyl, and 2-indanylmethyl, concluded that as pyrolysis of the 1-tetralyl perester gave no detectable amount of methylindan, formation of 1-methylindan was primarily through the 2-tetralyl radical. These reports represent that the 1-tetralyl radical appears to be a more important intermediate than the 2tetralyl radical in exchange with coal and that, in noncatalylic system without coal, the 2tetralyl radical as well as the 1-tetralyl radical would become important. In the hydrogen exchange of naphthalene resulting from the dehydrogenation of tetralin, the 1-tetralyl radical would be also important in exchange with coal. In our system with coal, however, the 2-tetralyl radical may be formed competitively with the 1-tetralyl radical to lead to the hydrogen exchange at the p-position of naphthalene since 1-methylindan was produced as a main product.

4.4 HydrogenTransfer Reaction in Coal Liquefaction

231

4.4.3 Effect of Coal Rank


The effectiveness of the tritium and 14C tracer techniques in tracing the reaction pathways of hydrogen atoms in coal liquefaction and quantitative information related to the mobility of hydrogen in coals have been discussed in the preceding sections (Kabe et al., 1983a, 1986a, 1987a-b, 1989a, 1990d). It was shown that the hydrogen exchange reaction between coal and gaseous hydrogen proceeds even at 300 ~ in Datong coal liquefaction, which made it possible to compare the hydrogen exchange reactions of three kinds of coals with different ranks in an extended temperature range. Below, the hydrogen exchange reactions of Datong coal as a bituminous coal, Wandoan coal as a subbituminous coal, and Morwell brown coal as a brown coal with tritiated gaseous hydrogen were investigated in the temperature range of 200-400 ~ and the hydrogen mobility of coal under coal liquefaction conditions are estimated in detail (Kabe et al., 199 lb). Datong, Wandoan, and Morwell coals were liquefied at 300-400 ~ for 120 min and the results are shown in Fig. 4.37. The yields of SRC (coal products) increased with temperature. The rate of liquefaction decreased in the order Morwell > Wandoan > Datong, which shows that coals with higher carbon content are more difficult to liquefy. Fig. 4.38 shows the variations in the tritium concentrations of residue and tetralin with temperature. At 200-230 ~ the tritium concentrations were very low and hydrogen exchange hardly occurred. At 300 ~ tritium was introduced to residue, indicating that hydrogen exchange reaction between coal (residue and SRC and gaseous hydrogen occurred at this temperature (vide infra). The tritium concentration of residue increased with temperature. However, the tritium concentration of tetralin remained very low below 350 ~ and it was much smaller than that of residue throughout the entire temperature range. The hydrogen exchange ratio is plotted against reaction temperature in Fig. 4.39. Although the liquefaction scarcely proceeded at 300 ~ the hydrogen exchange reaction of coal occurred. With a rise from 350 to 400 ~ the hydrogen exchange ratio increased remarkably and nearly 50% of 100 80
O

~- 60 ,~ 40 20

300

350 Reaction temp erature (~

400

Fig. 4.37 Effectof reaction temperature on yields of residue and SRC in liquefaction of several coals for 120, min. O, A: Datong; ~,/k: Wandoan; O, A: Morwell 0, ~, O: Residue; A,/k, A: SRC [From Kabe. T. et al., Energy Fuels, 5, 460 (1991)]

232

4 Liquefaction of Coal

- -

X
"0 0

10

Datong Wandoan Morwell

Residue Q ~ 0

Tetralin II t-A I--!

5
0 0

200

300 Reaction temperature (~

400

Fig. 4.38 Effect of reaction temperature on tritium concentrations in residue and tetralin. [From Kabe. T. et al., Energy Fuels, 5, 460 (1991)]

hydrogen in Datong coal or Morwell coal exchanged. For Wandoan coal, it was somewhat small. Since the hydrogen exchange between coal and gaseous hydrogen proceeded rapidly at the temperatures required for significant coal liquefaction, the exchange seems to be related to thermally produced radicals. The change of yields of residue and SRC with reaction time is plotted in Fig. 4.40. Liquefactions of Datong, Wandoan, and Morwell coals were performed at 400, 400, and 350 ~ respectively. When Morwell coal was liquefied at 400 ~ yields became extremely large and it was difficult to obtain complete material and tritium balance. To make the yield of Morwell coal similar to those of Datong and Wandoan, liquefaction of Morwell coal was performed at 350 ~ At 30 min, the yields of residues of Datong, Wandoan and
50
_

Reaction time: 2hr Datong 9 Wandoan Morwell O

.,,.~

40

~ 30
~= 20

~Z

10
A

--O
200

..
300 Reaction temperature (~ 400

Fig. 4.39 Effect of reaction temperature on the hydrogen exchange ratio of coal with gaseous hydrogen. [From Kabe. T. et al., Energy Fuels, 5, 461 (1991)]

4.4 Hydrogen Transfer Reaction in Coal Liquefaction

233

Morwell coals were 57, 25, and 50%, and the yields of SRC were 42, 61, and 42%, respectively. Although the extent of liquefaction does not necessarily follow the rank or the carbon content of coals (Yarzab, 1980), Datong coal, which is the highest rank among three, was the most difficult to liquefy even at 400 ~ In contrast to these observations, the hydrogen exchange reaction showed different results. The hydrogen exchange ratio is plotted against reaction time in Fig. 4.41. The hydrogen exchange ratio increased with time. After 300 min, the hydrogen exchange ratio of Datong coal was over 50%. Even at 350 ~ the HER of Morwell coal approached nearly 50%. On the other hand, the HER of Wandoan coal was small, 30% even after 300 min. These results showed that even though Datong coal was the most difficult to liquefy, hydrogen in Datong coal were the most mobile among the three coals. Since Datong coal has the highest rank, it can be presumed to have the most polycondensed structure. Since the radicals generated in liquefaction can be stabilized in aromatic molecules, they may promote the hydrogen exchange reaction rather than the hydrocracking reaction. As coal was liquefied, tritium in the gas phase was introduced into the coal. However, the amount of hydrogen exchanged in coal seems to differ depending on the kind of coal structure. Fig. 4.42 shows changes in the tritium distribution during coal liquefaction. Because in the noncatalytic system the amount of hydrogen added into coal was very small and the hydrogen distribution was nearly constant between the initial and final stages, it was approximated by straight lines. In Fig. 4.42, horizontal dotted and solid lines represent the hydrogen distributions of coal (Morwell and Datong 13% ; Wandoan 17%) and solvent (Morwell and Datong 74%; Wandoan 70%) among three phases, respectively. The arrow in Fig. 4.42 represents the hydrogen distribution of the gas phase (13%) among the three phases. When the hydrogen exchange reaction approaches equilibrium among the three phases, the tritium distribution in each phase will approach the hydrogen distribution in the phase. When Datong coal was used, the hydrogen exchange between gas phase and coal occurred at the initial stage of the reaction, then tritium was transferred from coal to solvent. With Morwell and Wandoan coals, the rate of tritium transfer from gas phase to coal and solvent was approximately equal. In Morwell coal, tri100

80
O

7K-

~- 60

40 >, 20

Reaction time (hr) Fig. 4.40 Changes in yields of residue and SRC with reaction time. O, A: Datong (400 ~ ~ , / k : Wandoan (400 ~ O, A: Morwell (350 ~ O, ~, O: Residue; A,/k A: SRC [From Kabe. T. et al., Energy Fuels, 5, 461 (1991)]

234

4 Liquefaction of Coal 60 50
O

40
cD ~x0

.~ 30
~

20 10
0

~Jlv

Wandoan ~ 400 ~ Morwell


I I I I

O 350 ~
I

Reaction time (h) Fig. 4.41 Changes in the hydrogen exchange ratio of coal with gaseous hydrogen with reaction time. [From Kabe. T. et al., Energy Fuels, 5, 461 (1991)]
100

80

9 ~.. 60
. ,...,

"~ 40
E

~" 20

Reaction time (h) Fig. 4.42 Change in the tritium distributions during coal liquefaction. Datong (400 ~ O, II, A; Wandoan (400 ~ (D, ill,/k; Morwell (300 ~ O, IS], A Coal: O, ~, O; Gas phase: II, [], ~; Solvent: A,/t,, A Upper solid lines, hydrogen distribution of solvent among three phases; dotted lines, that of coal; arrow, that of gas phase. [From Kabe. T. et al., Energy Fuels, 5, 462 (1991)]

tium introduced into coal transferred to solvent very slowly, while in Wandoan coal the tritium transfer from coal to solvent was very fast. The reason for this result is not yet understood. The reactivity of hydrogen in coal decreased in the order Datong ~ Morwell Wandoan, which is consistent with the result from HER in Fig. 4.41. Since tetralin, which has aromatic and naphthene rings in its structure, can be regarded as a model of one type of structure in coal, the hydrogen exchange reaction of tetralin with gaseous hydrogen was also investigated. The hydrogen exchange ratio of tetralin was 0.2% at 350 ~ and increased with increase in temperature. However, the exchange ratio was below 1% even at 400 ~ and the tritium concentration of tetralin was about one tenth that of coal. As shown

4.4 Hydrogen Transfer Reaction in Coal Liquefaction

235

90

O
70
O m

~ 5o ~= 3o 1
m

A--

T
2

I
4

I
6

I
8

Reaction time (h) Fig. 4.43 Change in yields of residue and SRC with reaction times at 300 ~ Datong: O, A; Wandoan: ~, A; Morwell: O, A Residue: O, ~, O; SRC: A,/k, A. [From Kabe. T. et al., Energy Fuels, 5, 462 (1991)] (Kabe et al., 1991b)

in Fig. 4.38, a substantial amount of tritium can be introduced into tetralin in the presence of coal. However, this result shows that tetralin could not be tritiated in the absence of coal. These results indicate that the exchange reaction of hydrogen in tetralin requires types of radicals produced from coal which are not produced from the thermolysis of neat tetralin. It seems that radicals produced in coal react easily not only with gaseous hydrogen but also with hydrogen in tetralin to cause the hydrogen exchange. Since it was clarified that the hydrogen exchange reaction proceeded even at 300 ~ coal liquefaction was further investigated at 300 ~ and the results are shown in Fig. 4.43. Yields of residue and SRC did not change with the elapse of time and Datong coal was hardly liquefied at 300 ~ Wandoan and Morwell coals were liquefied to give SRC in 20 and 25 wt% yields, respectively. However, these values did not change after 240-360 min, indicating that hydrocracking reactions proceed slowly at 300 ~ The change in tritium concentration with time at 300 ~ is shown in Fig. 4.44. The tritium concentration of each of the three coals approached low constant values below 3000 dpm/g, and that of Datong coal was the highest among the three. Tritium transfers to coal through both hydrogen addition and exchange reaction. In order to estimate the hydrogen exchange ratio, the amount of tritium transferred by hydrogen addition must be subtracted. The hydrogen exchange ratio at 300 ~ is plotted against reaction time in Fig. 4.45. After the amount of hydrogen added was subtracted, the hydrogen exchange ratio increased in the order of D a t o n g Wandoan ~ Morwell. The largest amount of tritium was transferred to Datong coal; however, since the amount of hydrogen added to Datong coal was larger than that added to Wandoan and Morwell coals, the hydrogen exchange ratio of Datong coal became small. The hydrogen exchange ratio for Datong, Wandoan, and Morwell approached constant values, 4.5, 5.0 and 7.8%, respectively. The HER for Morwell coal was the largest and therefore the hydrogen exchange at 300 ~ may be related to the exchange of hydrogen in functional groups such as -OH and -NH. Although a detailed analysis of such active hydrogen has not been done, the comparable analysis of bituminous coals has been reported (Pestryakov, 1986; Maekawa, 1975). Bituminous coals which have a chemical composition

236

4 Liquefaction o(Coal

)<

Residue SRC

Datong 9 A

Wandoan 9 A

Morwell 0 A

3
O .,..~

Reaction time (h) Fig. 4.44 Change in tritium concentrations with reaction time at 300 ~ [From Kabe. T. et al., Energy Fuels, 5, 462 (1991)]

Reaction temperature: 300 ~


~"

.2
*~ 6 .
C9 cD

Datong Wandoan Morwell

9 9 2 / / / ~

O ~ fib

;;m

4 Reaction time (h)

Fig. 4.45 Effect of reaction time on the hydrogen exchange ratio of coal with gaseous hydrogen at 300 ~ [From Kabe. T. et al., Energy Fuels, 5, 462 (1991)]

of C 75-85% and H 5.0-5.4wt% contain 6-12 atom% of phenolic OH hydrogen for total hydrogen. Yokoyama et al. (1967) reported that high-rank coals, which have a chemical composition of C 75-84% and H 5.8-6.4% (daD, contain 3-9 atom% of phenolic (OH) hydrogen and carboxylic acid (COOH) hydrogen for total hydrogen, while low-rank coals which have a chemical composition of C 61-70% and H 5.3-6.0% (daf), contain 12-14 atom% of those. Kotanigawa et al. (1979) reported that the exchange reaction between deuterium gas and aromatic hydrogen in phenol took place rapidly at 350 ~ with ZnOFe203 catalyst and that no such exchange reaction occurred in the absence of catalyst. They did not refer to the exchange reaction between deuterium gas and hydrogen of the hydroxy group in phenol. The reaction of phenol with tritiated gaseous hydrogen was carried out at 340 ~ for 2 h in the absence of a catalyst; 8.8% of the hydrogen in phenol underwent tritium exchange.

4.4 HydrogenTransferReaction in Coal Liquefaction

237

Since it can be assumed that only hydrogen in the hydroxy group in phenol is exchangeable, this indicates that 53% of the hydrogen in the hydroxy group in phenol exchanged with gaseous hydrogen at 340 ~ for 2 h. Further, the reaction of aniline with tritiated gaseous hydrogen was also performed at 300 ~ for 2 h in the absence of a catalyst; 13.7% of the hydrogen in aniline underwent tritium exchange. Since it can be assumed that only the hydrogen in the amino group in aniline is exchangeable, this indicates that 48% of the hydrogen in the amino group in aniline exchanged with gaseous hydrogen at 300 ~ for 2 h. These results support the suggestion that, in the reaction of coal with gaseous hydrogen, OH and NH hydrogen in polycondensed aromatic compounds were exchanged at lower temperatures. 4.4.4 Effect of Solvent

Solvents play an important role in coal liquefaction because they can be used as hydrogen donors and dissolve some portion of the coal (Whitehurst et al., 1980). Solvents with naphthene tings, such as tetralin, mainly serve as donor solvents, while those with two or more aromatic tings, such as methylnaphthalene, dissolve a larger amount of coal than those with aliphatic structures because coal mainly consists of condensed aromatic structures. Naphthenic solvents, such as decalin, may have poorer ability to donate hydrogen or to dissolve coal than tetralin or methylnaphthalene. Since the liquefaction includes hydrogenation and hydrocracking of coal, with hydrogen in the gas phase and solvent, a number of attempts have been made to clarify the hydrogen transfer mechanism in coal liquefaction in the presence of solvents (Billmers et al., 1986; McMillen et al., 1985; Murakata et al., 1993). Billmers et al. (1986) suggested a free radical mechanism following kinetic experiments in model reactions of coal liquefaction. McMillen et al. (1985) reported the importance of solvent radicals in the hydrogen transfer reaction between the coal model and solvent. In these researches, tetralin was used as a hydrogen donor solvent, and decalin (Murakata et al., 1993) and methylnaphthalene (Oga et al., 1985; Sato et al., 1992), which have poor ability as hydrogen donors, were used as solvents. A more useful method to trace hydrogen transfer mechanisms in coal liquefaction is to utilize isotopes, such as deuterium and tritium tracers. A deuterium tracer was effective in tracing reactive sites in coal and coal model compounds; however, there are few examples which enable quantitative analysis of hydrogen mobility in coal because of the poor solubility of coal products and the difficulty of quantification of the deuterium tracer (Fu and Blaustein, 1967; Franz and Camaioni, 1981b; Brower, 1982; Schweighardt et al., 1976; Cronauer et al., 1982; Wilson et al., 1984; Collin and Wilson, 1983; Skoweonski et al., 1984). Further, hydrogen transfer mechanisms in the presence of various solvents have not yet been sufficiently clarified using the deuterium tracer. For example, Benjamin et al. (1982) studied the hydrogen exchange reaction of a group of aromatic compounds in recycled solvents with diphenylmethane-d2 (Ph2CD2), deuterated pyrene or D2 gas under liquefaction conditions, assuming that reactivity toward hydrogen exchange is related to hydrogen shuttling. They reported that methyl substituted aromatics, such as methylnaphthalene and toluene, underwent extensive exchange reactions, while non-substituted aromatics, such as naphthalene, biphenyl ether, showed little observable exchange with three deuterated reagents. They concluded that the methyl substituted aromatic and hydroaromatic compounds in the recycled solvent make the most important contribution to hydrogen shuttling and hydrogen transfer. However, the detailed mechanisms and the position of the hydrogen exchange were not discussed. Recently, it has been reported that tritium and 14C tracer techniques are effective in tracing quantitatively the hydrogen in coal liquefaction (Kabe et

238

4 Liquefaction of Coal

al., 1987b, 199 la-b, Ishihara et al., 1993). In these works, it was shown that quantitative analysis of hydrogen mobility in coal can be determined by hydrogen exchange reactions between the coal, gas phase and solvent, as well as by hydrogen addition. In order to investigate the effect of the kind of solvent on hydrogen transfer between the coal, gas phase and solvent, the reaction of tetralin, decalin and 1-methylnaphthalene with tritiated gaseous hydrogen was investigated in the absence and presence of coal to estimate hydrogen mobility
10

~
O

0
280

, -300

,
320

, , A ...
340 360 380 400

.
420

Temperature (~ Fig. 4.46 Effect of temperature on the yields of products from solvents in the absence of coal. Products from tetralin: O Naphthalene; 9 n-butylbenzene; 9 l-methylindan Products from decalin: A Naphthtalene; I, Tetralin Products from 1-methylnaphthalene: [] Naphthalene [From Kabe. T. et al., Prepr., ACS Div. Petrol. Chem. (1994)]

30

~O
O .,,.,

20

10

0 280

300

320

340

360

380

400

420

Temperature (~ Fig. 4.47 Effect of temperature on the hydrogen exchange ratio of solvents in the presence and the absence of coal. In the presence of coal: O Tetralin; A Decalin; V-I 1-Methylnaphthalene In the absence of coal: 9 Tetralin; I, Decalin; 9 1-Methylnaphthalene [From Kabe. T. et al., Prepr., ACS Div. Petrol. Chem. (1994)]

4.4 HydrogenTransfer Reaction in Coal Liquefaction

239

of the solvent and coal quantitatively (Ishihara et al., 1995). Before examining complicated reactions with coal, solvent reactions with tritiated hydrogen in the absence of coal were performed under the conditions of 300-400 ~ and 5.9 MPa. Although tritium was introduced to the solvents by hydrogen addition and hydrogen exchange reactions, most of the tritium was introduced through hydrogen exchange. The product yields and HER of solvents are plotted against temperature in Figs. 4.45 and 4.46, respectively. In the reaction of tetralin, the products were naphthalene (NP) by dehydrogenation, n-butylbenzene (BB) by hydrocracking and the main product. 1-methylindane (MI) by isomerization. These products remarkably increased with a rise from 375 to 400 ~ and the yields of NP, BB and MI reached 0.8, 0.7 and 1.5%, respectively. Decalin was not formed by disproportionation. This is consistent with a previously reported result (Hooper et al., 1979). Tetralin and naphthalene were formed from decalin above 375 ~ Although the yields of these products increased with temperature, the values were lower than 0.4%, even at 400 ~ Naphthalene and very small amounts of unidentified products were formed from 1-methylnaphthalene. The yield of naphthalene increased remarkably with temperature and reached about 9% at 400 ~ indicating that 1-methylnaphthalene was easy to decompose above 350 ~ In the absence of coal, tritium was introduced into solvents over 350 ~ Tritium balances in the reaction of solvents with tritiated hydrogen, and the amount of hydrogen exchanged, are listed in Table 4.10. HERs are also plotted against temperature in Fig. 4.47. Although HERs increased with temperature, HERs of tetralin, decalin and methylnaphthelene were only 2.0, 1.5 and 3.1%, respectively, even at 400 ~ The results in the absence of coal indicate that the simple solvent by itself is difficult to exchange with hydrogen molecules in the gas phase under the conditions generally used for coal liquefaction, probably because it is difficult to form radicals without coal. The reaction of Wandoan coal with tritiated gaseous hydrogen was performed in the presence of tetralin, decalin or 1-methylnaphthalene. The effect of temperature on the conversion of coal is shown in Fig. 4.48. The conversion of coal, which was calculated from the difference between weights of the reacted coal and its tetrahydrofuran insoluble fraction, decreased in the order tetralin > methylnaphthalene > decalin, and those at 400 ~ were 87, 54 and 45%, respectively. The main product yields from solvents are plotted against temperature in Fig. 4.49. In these reactions, tetralin was converted to naphthalene by donating hydrogen to coal. Very little decalin was formed by hydrogen addition from either the coal or gas phase to tetralin. Although the formation of tetralin and naphthalene Table 4.10 TritiumDistribution and Hydrogen Exchange after Reaction of Solvent with Tritiated Gaseous Hydrogen in the Absence of Coala Temperature (~ 350 350 350 375 375 400 400 400
a

Rgas

Rsolvent

Solvent Tetralin Decalin 1-Methylnaphthalene Tetralin Decalin Tetralin Decalin 1-Methylnaphthalene

(dpm) 992771 990712 992387 962020 954558 907118 900963 890775

(dpm) 7229 9288 7613 37980 45442 92882 99037 109225

Amount of hydrogen exchangedb (g)


9 . 7 6 X 10 -3 1.23 X 10 -2 1.04 X 10 -2 5.29 X 10 -2 6.24 X 10 -2

1.37 X 10-1 1.44 X 10-1 1.66 X 10-1

Reaction time, 120 min. Total radioactivities were normalized o n 10 6 ; bTritium recovery, 100_5% [Reproduced with permission from Ishihara, A. et al., Fuel, 74, 64, Elsevier, (1995)]

from decalin was observed, the yields were very small, as shown in Fig. 4.49. The amount of hydrogen addition from tetralin to coal at 300 ~ was 0.13 g, similar to that from decalin

240

4 Liquefaction of Coal 100 80


O O O .,..~

60

~. 40
O

200 , I i I i I , I , I i l i

280

300

320

340

360

380

400

420

Temperature (~ Fig. 4.48 Effect of temperature on the conversion of coal. O Tetralin; A Decalin; 7q 1-Methylnaphthalene [Reproduced with permission from Ishihara, A. et al., Fuel, 74, 66, Elsevier (1995)]

to coal at 400 ~ i.e., 0.11 g. Coal conversion with tetralin at 300 ~ was 35%, which was close to that with decalin at 400 ~ i.e., 45%, indicating that such an a m o u n t of hydrogen f r o m solvent to coal can c o n v e r t m o r e than one third of W a n d o a n coal. A significant a m o u n t of naphthalene was f o r m e d from 1-methylnaphthalene. The yield of naphthalene f r o m 1 - m e t h y l n a p h t h a l e n e r e m a r k a b l y i n c r e a s e d with t e m p e r a t u r e and r e a c h e d 17% at 400 ~ The fact that the conversion of coal in m e t h y l n a p h t h a l e n e was higher than that in decalin may be due to the difference in the solubility of coal in m e t h y l n a p h t h a l e n e and decalin, or the addition of methyl radicals formed from m e t h y l n a p h t h a l e n e to coal radicals. 30

es 20

~,

10

, ~ , -

280

300

320

340

360

380

400

420

Temperature (~ Fig. 4.49 Effect of temperature on the yields of products from solvents in the presence of coal. Products from tetralin: O Naphthalene; 9 Decalin Products from decalin: A Naphthalene; A Tetralin Product from 1-methylnapthhalene: E-]naphthalene [Reproduced with permission from Ishihara, A. et al., Fuel, 74, 66, Elsevier (1995)]

4.4 Hydrogen Transfer Reaction in Coal Liquefaction

241

H y d r o g e n e x c h a n g e b e t w e e n h y d r o g e n in s o l v e n t and tritiated g a s e o u s h y d r o g e n in the p r e s e n c e o f coal was estimated. T r i t i u m d i s t r i b u t i o n s and the a m o u n t s o f h y d r o g e n exc h a n g e d in s o l v e n t are listed in T a b l e s 4.11 and 4.12. A l t h o u g h tritium was i n t r o d u c e d into s o l v e n t by h y d r o g e n addition and h y d r o g e n e x c h a n g e reactions, m o s t o f the tritium w a s int r o d u c e d t h r o u g h h y d r o g e n e x c h a n g e . H E R s are p l o t t e d against t e m p e r a t u r e in Fig. 4.47. H E R s of tetralin and d e c a l i n in the p r e s e n c e o f coal i n c r e a s e d g r a d u a l l y with an i n c r e a s e in t e m p e r a t u r e , and r e a c h e d 8.1 and 3.5%, r e s p e c t i v e l y , at 4 0 0 ~ HER of methylnaphthalene r e m a r k a b l y i n c r e a s e d with a rise f r o m 350 to 4 0 0 ~ This result m a y c o r r e s p o n d to the r e m a r k a b l e d e c o m p o s i t i o n o f m e t h y l n a p h t h a l e n e in this t e m p e r a t u r e range. As s h o w n in T a b l e 4.12, the a m o u n t o f h y d r o g e n r e q u i r e d f r o m d e c o m p o s i t i o n o f m e t h y l n a p h t h a l e n e to n a p h t h a l e n e and m e t h a n e was l a r g e r than that p r o v i d e d f r o m the gas phase. T h e r e f o r e , it c a n be c o n s i d e r e d t h a t the h y d r o g e n r e q u i r e d to f o r m n a p h t h a l e n e a n d m e t h a n e f r o m m e t h y l n a p h t h a l e n e was m a i n l y p r o v i d e d by coal. A m a n o et al. (1972) s u g g e s t e d f r o m kin e t i c e x p e r i m e n t s that, u n d e r c o n d i t i o n s o v e r 6 0 0 ~ a n d 5 m o l H2 p e r m o l t o l u e n e , Table 4.11 Tritium Distribution after Reaction of Gaseous Hydrogen in the Presence of Solventsa

Temperature
(~ 300 Solvent Tetralin Decalin 1-Methylnaphthalene Tetralin Decalin 1-Methylnaphthalene Tetralin Decalin 1-Methylnaphthalene

Rgas

Rcoal

Rsolvent

(dpm) 937202 906860 897886 727018 788071 842397 576002 603196 458413

(dpm) 62798 90492 79247 159213 190152 112959 181586 225842 143554

(dpm) 27581 2648 22868 113768 21777 44644 242412 170962 398033

350

400

a Reaction

time, 120 min. Total radioactivities were normalized to 106 dpm. Tritium recovery, 100 + _ _5% [Reproduced with permission frorfi Ishihara, A. et al., Fuel, 74, 66, Elsevier (1995)] Table 4.12 Hydrogen Transfers among Coal, Gas Phase and Solvent" Amount of hydrogen added Amount of hydrogen exchanged Solvent (g) 3.56 X 3.45 X 3.11 X 1.89 X 6.26 X 6.47 X 5.09 X 3.34 X 1.06 X
10 -2 10 -3

Temperature (~ 300

Solvent Tetralin Decalin Methylnaphthalene Tetralin Decalin Methylnaphthalene Tetralin Decalin Methylnaphthalene

Gas to coal (g) 0 0 0 0 0 0 4.04 X 10- 2 5.96 X 10 -2 6.38 X 10- 2

Solvent to coal (g)


1.27 X 10 - l b

Coal (g) 1.17 X 1.21 X 1.39 X 4.54 X 3.18 X 2.29 X 8.50 X 7.16 X 1.38 X 10 -1 10 -1 10-1
10 -1

350

400

9.78 X 3.96 X 3.41 X 6.53 X 5.88 X 5.26 X 1.11 X 1.96 X

10-5c 10- 5d 10-1 b 10- 3c 10- 2d 10-1 b 10 -lc 10-1 d

10- 2 10-1 10- 2 10- 2 10-1 10 -1 10~

10-1 10-1 10-1 10-1 10~

aCoa130

g: Solvent 75 g; 5.9 MPa. Initial amounts of hydrogen in the gas phase were 1.21, 1.18 and 1.22 g in the cases of tetralin, decalin and methylnaphthalene, respectively. Initial amounts of hydrogen in solvent were 6.82, 9.78 and 5.28 g for tetralin, decalin and methylnaphthalene, respectively. bAmount of hydrogen added with formation of naphthalene. cAmount of hydrogen added with formation of naphthalene and tetralin. din the case of methylnaphthalene, hydrogen addition from coal to solvent occurred to from naphthalene and methane. [Reproduced with permission from Ishihara, A. et al., Fuel, 74, 67, Elsevier (1995)]

242

4 Liquefactionof Coal

demethylation of toluene proceeded via radical chain mechanism, and that the reaction of toluene with the hydrogen atom to form benzene and a methyl radical was the rate-determining step. It has also been reported that, in hydrocracking of toluene at 455-490 ~ and 0.14 MPa, dealkylation of toluene proceeds via the radical chain mechanism (Gonikberg and Nikitenkov, 1954). A similar mechanism has been reported for demethylation of methylnaphthalene (Sato et al., 1993). Ogata et al. (1983) reported that the hydrogen atom which was formed by the addition of hydrogen sulfide, promotes dealkylation of methylnaphthalene. Further, Ogo et al. (1985) showed that dealkylation of methylnaphthalene proceeds by a radical chain mechanism, where radicals or hydrogen atoms are formed secondarily or by thermolysis of biphenyl. Taking into account these reports, the radical chain mechanism can be assumed in the case of methylnaphthalene. Possible mechanisms of hydrogen exchange and dealkylation with methylnaphthalene are shown in Eqs. (4.35)-(4.43) Hydrogen exchange reactions of coal proceed through the route shown in Eqs. (4.35)-(4.37). The reactions in Eqs. (4.35)-(4.37) are common among tetralin, decalin and methylnaphthalene. The tritium radical formed will react with methylnaphthalene to form tritiated naphthalene and a methyl radical, as shown in Eq. (4.38). The methyl radical reacts with another methylnaphthalene to form methane and a naphthylmethyl radical, which may react with tritiated coal to form tritiated methylnaphthalene and a coal radical, as shown in Eqs. (4.39) and (4.40). Dealkylation of methylnaphthalene can be explained by Eqs. (4.38)-(4.40) reasonably. In hydrogen exchange of methylnaphthalene through these routes, however, the HER does not exceed 4% of hydrogen in methylnaphthalene when it is calculated on the basis of conversion of methylnaphthalene, Under conditions where methylnaphthalene decomposes, methylnaphthalene will easily form a naphthylmethyl radical through the reaction with the coal and tritium radical as shown in Eqs. (4.41) and (4.42). A naphthylmethyl radical formed in such reactions can be tritiated through the routes shown in Eqs. (4.40) and (4.43). The remarkable decomposition of methylnaphthalene and the high HER value of methylnaphthalene at 400 ~ can be explained by considering the radical chain mechanism described above. If it is assumed that hydrogen exchange proceeds through the free radical mechanism, the relative ease of the formation of radicals would be related to the strength of a bond. For example, the bond dissociation energies of C-H in the benzene and benzyl positions of toluene are 469 and 356 kJ mol-', respectively. The latter hydrogen will be easier to exchange than the former. Concerning the hydrogen exchange reaction of tetralin, extensive studies using deuterium have been reported. As mentioned in section 4.4.2, it was proposed that the 1-tetralyl radical appears to be a more important intermediate than the 2-tetralyl radical in exchange with coal. In the hydrogen exchange of naphthalene resulting from the dehydrogenation of tetralin, the 1-tetralyl radical would also be important in exchange with coal. Possible mechanisms of hydrogen exchange of tetralin in the presence of coal are shown in Eqs. (4.44)-(4.47). 1-Tetralyl radicals are formed by the reaction of tetralin with a coal radical or a tritium atom, as shown in Eqs. (4.44) and (4.45). The formed tetralyl radical will react with a tritiated hydrogen molecule or a tritium atom in tritiated coal to form tritiated tetralin, as shown in Eqs. (4.46) and (4.47). Although tetralin may easily form a tetralyl radical, its lifetime is shorter than that of a naphthylmethyl radical, because the formation of naphthalene from a tetralyl radical, as shown in Eq. (4.48), occurs more easily in the presence of coal than the formation of tritiated tetralin, as shown in Eqs. (4.46) and (4.47). Therefore, the HER of tetralin in the presence of coal did not increase with a rise from 350 to 400 ~ as much as that of methylnaphthalene. Decalin does not form such radicals as the naphthylmethyl or tetralyl radical, which are stabilized by aromatic rings. Therefore, the remarkable increase

4.4 Hydrogen TransferReactionin CoalLiquefaction Coal-H Coal, 4- T2 Coal, + CH3 T9 Coal, 4- H,

243

(4.35) (4.36) (4.37)

Coal-T 4- To Coal-T T CH3

T 4- CH3" CH3 @ CH2, @ CH3 + Coal-T @ CH2, + CH3 + CH2, 4T2


"

(4.38)

CH24- CH3. @ CHzT + Coal. (4.40) 4- CH4 (4.39)

Coal-H

(4.41)

CH2, T.
' ~ +

T-H

(4.42)

CH2T
~ " 1 + T.

(4.43)

+ Coal,

'

+ Coal-H

(4.44)

+ T.

~ T

+ T-H

(4.45)

+ T-H

+ H-

(4.46)

244

4 Liquefaction of Coal

T ~ ] + Coal-T " ~ ~ ] + Coal. (4.47)

~ ~ ]

Coal.

Coal-H

(4.48)

in tetralin or decalin HER with a rise from 350 to 400 ~ did not occur. Hydrogen exchange of hydrogen in coal with tritiated gaseous hydrogen was also estimated. Since the HERs of solvents markedly increased in the presence of coal, it can be assumed that tritium in the gas phase may transfer to a solvent through coal. Based on this assumption, the HER of coal in the presence of solvent was estimated and plotted against temperature in Fig. 4.50. The reaction of coal with tritiated gaseous hydrogen in the absence of a solvent was also performed, for comparison with that in the presence of a solvent. Tritium balance and the amount of hydrogen exchanged are listed in Table 4.13. The HER of coal in the absence of a solvent was also plotted in Fig. 4.50. The HER of coal in inactive decalin was very similar to that in the absence of a solvent. Since tetralin can easily form a tetralyl radical under the same conditions, the HER of coal in tetralin was slightly higher than that of decalin. However, tetralin did not promote to an extreme degree the hydrogen exchange reaction among the gas phase, coal and solvent for the reasons described above. In contrast, the value of methylnaphthalene at 400 ~ deviated largely from the other curves. Hydrogen exchange in the presence of methylnaphthalene can be regarded as a reaction proceeding through the radical chain mechanism described in Eqs. (4.35)-(4.43). Different from the cases of tetralin and decalin are the reactions of the naphthylmethyl radical in Eqs. (4.40) and (4.43). Especially, the remarkable deviation of methylnaphthalene at 400 ~ in Fig. 4.50 may be related to the reaction of Eq. (4.43), in which tritium can be incorporated from the gas phase to methylnaphthalene directly.
Table 4.13 Tritium Distribution and Hydrogen Transfer after Reaction of Coal with Tritiated Gaseous Hydrogen in the Presence of Solvents" Temperature (~ 260 300 360 385 410
a

Rgas (dpm) 993261 896987 775033 705166 650260

Rcoal (dpm) 6378 103013 224967 294834 349740

Amount of hydrogen exchanged (g) 1.03 1.84 3.35 5.02 7.81 X X X X X 10- 2 10- ~ 10-1 10- ~ 10- ~

Amount of hydrogen added (g) 0.0000 0.0000 0.1006 0.1176 0.0516

Reaction time, 120 min. Total radioactivities were normalized to 106 dpm. Tritium recovery, 100 ___5%. Initial amount of hydrogen in gas phase, 1.6 g. Initial amount of hydrogen in coal. 1.86 g. [Reproduced with permission from Ishihara, A. et al., Fuel, 74, 65, Elsevier (1995)]

4.4 Hydrogen Transfer Reaction in Coal Liquefaction


100

245

o
0 ~0

80

"~ 60

N 40 .~ 2o
0 200
I , I

300 400 Temperature (~

500

Fig. 4.50 Effect of temperature on the hydrogen exchange ratio of coal. O Tetralin; A Decalin; E] 1-Methylnaphthalene; 9 Solvent free [Reproduced with permission from Kabe. T. et al., Prepr. A CS Div Petrol. Chem. (1994)]

4.4.5

Behavior of Representative Compounds

Solvents play an important role in coal liquefaction by acting as hydrogen donors or by directly dissolving a portion of the coal (Whitehurst et al., 1980), providing important thermal radical initiation pathways (Ruchardt et al., 1997). Tetralin, serves mainly as a donor solvent. Solvents with two or more aromatic rings dissolve a larger amount of coal than solvents with aliphatic structures because coal consists of condensed aromatic structures. Naphthenic solvent, such as decalin, may have poorer ability to donate hydrogen or to dissolve coal than tetralin or naphthalene. Since the liquefaction includes hydrogenation and hydrocracking of coal with hydrogen in the gas phase and solvent, a number of attempts have been made to clarify the hydrogen transfer mechanism in coal liquefaction in the presence of solvents (Billmers et al., 1986; Malhotra and McMillen, 1993; Murakata et al., 1993). Billmers et al. (1986) suggested a free radical mechanism following kinetic experiments in model reactions of coal liquefaction. Malhotra and McMillen (1993) reported the importance of solvent radicals in the hydrogen transfer reaction between the coal model and solvent. In these studies, tetralin was used as the hydrogen donor solvent, and decalin and methylnaphthalene were used as solvents which have poor ability as hydrogen donors (Ogo et al., 1995; Sato et al., 1992). As noted in Section 4.3.2, the sulfur or pyrite in coal correlates with an increase in coal conversion (Abdel-Baset et al., 1978; Montano and Granoff, 1980; Godo et al., 1997a). The pyrite is rapidly transformed into pyrrhotite, and H2S is produced from the reduction of pyrite under coal liquefaction conditions (Montano et al., 1981; Keisch et al., 1971; Harris et al., 1979). It has been suggested that HzS generated from pyrite is the catalyst for coal liquefaction (Lambert, 1982; Stenberg et al., 1983). On the other hand, H 2 0 is a major product in coal thermolysis and a large amount of H 2 0 is generated under rather mild coal liquefaction conditions. Since the reactions in coal liquefaction proceed with hydrocracking and hydrogenation by molecular hydrogen and donor solvents, the hydrogen transfer mechanism may be extremely influenced by the presence of HzS and H20. However, the role of HzS and H 2 0 regarding the hydrogen transfer in coal liquefaction is not well defined

246

4 Liquefactionof Coal

because the complex nature of coal and its derived products prevent a thorough understanding of the reaction mechanism. Here, the effect of H2S and H20 on the hydrogen exchange in the liquefaction using a model compound tetralin is discussed (Godo et al., 1997b, 1998c). Tetralin is one of the most interesting and convenient model compounds because it has an aromatic ring and a naphthene ring in its structure and can serve as an effective hydrogen donor solvent. The reactions of tetralin with tritiated hydrogen were performed under the conditions of 3 5 0 4 0 0 ~ in the presence of H2S. The products yields are plotted against the reaction time in Fig. 4.51. The reaction products from tetralin were 1-methylindan (MI) by isomerization, naphthalene (NP) by dehydrogenation and n-butylbenzene (BB) by hydrocracking. They increased monotonically with reaction time. The yields of MI, NP and BB in the presence of H2S at 400 ~ for 300 min were 2.3, 0.7 and 0.9%, respectively. Decalin was not formed under by disproportionation these conditions. The results in the absence of H2S were also plotted in Fig. 4.51. The yields of MI, NP and BB at 400 ~ for 300 rain were 2.1, 0.8 and 1.1%, respectively. The amount of each product in the presence of H2S was close to that in the absence of HaS. In the reaction of tetralin with tritiated hydrogen in the presence of H2S, the conversion and hydrogen exchange ratio of tetralin were plotted against reaction time in Figs. 4.52 and 4.53, and compared with those in the reaction in the absence of H2S. As shown in Fig. 4.52, the conversions of tetralin in the presence and absence of H2S at 400 ~ for 300 min were 4.1 and 3.9 %, respectively. These values were very close to each other. As shown in Fig. 4.53, the hydrogen exchange ratios of tetralin in the presence and absence of H2S increased gradually with time and reached 40.4 and 4.6%, respectively, at 400 ~ for 300 min. The hydrogen exchange ratio in the presence of HaS was about 10 times higher than that in the absence of H2S at 375 ~ and 400 ~ and reached 71% at 400 ~ for 600 min.

3
0

2
~D .,..~

/ / /

~
0

-.-~~
I ~ I , I ,

200

400 600 Reaction time (min)

800

Fig. 4.51 Effectof reaction time on product yields at 400 ~ In the presence of H2S: O: Naphthalene; A: n-Butylbenzene; I1: 1-Methylindan. Reaction in the presence of H20: (1: Naphthalene; A: n-Butylbenzene; []: 1-Methylindan. In the absence of H2S and H20: C): Naphthalene; A: n-Butylbenzene; [S]: 1-Methylindan. [From Godo. M. et al., Energy Fuels, 11,472 (1997)]

4.4 Hydrogen Transfer Reaction in Coal Liquefaction 10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

247

8 99

,"

,i"

/,

// 6

0
t/

/"

.2
D"

4
/

/"

/
/"

/
s.,l I

/"

o
2

,,0/'/,"
/ 9

0 ......

:'

'

"

'

'

''

'

'

'

'

'

'

200

400

600

800

Reaction time (rain) Fig. 4.52 Effect of reaction time on the conversion of tetralin. Reaction in the'presence of HzS: h-: 350 ~ I1:375 ~ A:400 ~ Reaction in the presence of H20: []: 375 ~ A: 400 ~ ~:425 ~ Reaction in the absence of H2S and HzO: []: 375 ~ A: 400 ~ 0:425 ~ [From Godo. M. et al., E n e r g y F u e l s , 11,472 (1997)]

10

....

........

......................... 0s '
/ t it

100

8
O

80

t~

6O

~D

= ~D
O

~)

,/,,'"

//

40

//

" ./" /

--'" .-~

+---'-"

20

0 '

"0

" 200 400 600 Reaction time (rain)

'

0 800

Fig. 4.53 Effect of reaction time on the hydrogen exchange ratio. Reaction in the presence of H2S: +: 350 ~ I1:375 ~ A: 400 ~ Reaction in the presence of H20: []: 375 ~ 400 ~ ~: 425 ~ Reaction in the absence of H2S and H20: []: 375 ~ A: 400 ~ (3:425 ~ [From Godo. M. et al., E n e r g y F u e l s , 11, 472 (1997)] F i r s t - o r d e r plots o f these data for the c o n v e r s i o n and h y d r o g e n e x c h a n g e are s h o w n in Fig. 4.54 (a, b). All plots are a p p r o x i m a t e d straight lines. T h e result indicates that these reactions c o u l d be treated as first-order reactions. T h e rate c o n s t a n t s o f c o n v e r s i o n and hyd r o g e n e x c h a n g e w e r e d e t e r m i n e d f r o m the slopes o f first-order plots. A s s h o w n in T a b l e 4.14, the rate constants o f tetralin c o n v e r s i o n in the p r e s e n c e and a b s e n c e o f H2S at 375 ~

248

4 Liquefactionof Coal 0.00 --0.02 --0.04 X ,~ -0.06


|

',,,,
(a)

-0.08 --0.10 --0.12 ' 2()0 ' 4()0 ' 6()0 ' Reaction time (min) 0.0 800

0.0

-0.5 X
|

--0.5

-0.10 (b) . . . . 2(}0 . 4()0

--1.0

--0.15

600

- 1.5 800

Reaction time (min) Fig. 4.54 First-orderplots of conversion of tetralin and hydrogen exchange ratio. a) Conversion of tetralin, b) Hydrogen exchange ratio. Reaction in the presence of HzS:+: 350 ~ I1:375 ~ A: 400 ~ Reaction in the presence of H20: ~: 375 ~ 400 ~ ~: 425 ~ Reaction in the absence of H2S and H20: F-l: 375 ~ A: 400 ~ 0:425 ~ [From Godo. M. et al., Energy Fuels, 11,473 (1997)]

and 400 ~ were not significantly different. On the other hand, as shown in Table 4.15, the rate constants of hydrogen exchange reactions in the presence of H2S were about 12-15 times higher than those in the absence of H2S. From the Arrhenius plots, activation energies for the tetralin conversion in the absence and presence of H2S were determind. They were 3 3 _ 1 and 35__+1 kcal/mol, respectively. These values were very close to each other, suggesting that the conversion of tetralin in the absence and presence of HzS proceeds via on the same reaction mechanism, and that the presence of H2S did not affect the activation energies. Similarly, activation energies for hydrogen exchange reactions were determind. Activation energies of the hydrogen exchange in the absence and presence of HiS were 30___ 1 and 34___1 kcal/mol, respectively. These values were slightly different from each other, suggesting that the hydrogen exchange of tetralin in the absence and presence of H2S proceed via different mechanisms. The reactions of tetralin with tritiated hydrogen were performed under the conditions of

4.4 HydrogenTransfer Reaction in Coal Liquefaction Table 4.14 Rate Constants of Tetralin Conversion (min-1) Reaction 350 Tetralin Tritiated hydrogen HzS Tetralin Tritiated hydrogen H20 Tetralin Tritiated hydrogen -1.45 10-5 Temperature (~ 375 3.37 X 10-5 400 1.19 X 10-4 425

249

2.32 X 10-5

5.48 X 10-5

1.69 10-4

3.85 10-5

1.15 X 10-4

2.40 X 10-4

[From Godo. M. et al., Energy Fuels, 11,473 (1997)] Table 4.15 Rate Constants of Hydrogen Exchange Reaction (min-~) Reaction 350 Tetralin Tritiated hydrogen H2S Tetralin Tritiated hydrogen H20 Tetralin Tritiated hydrogen -2.70 X 10-4 Temperature (~ 375 9.00 X 10-4 400 1.97 X 10-3 425

2.62 10-5

4.85 10-5

1.02 X 10-4

6.10 X 10- 5

1.65 10- 4

3.33 10- 4

[From Godo. M. et al., Energy Fuels, 11,473 (1997)] 375-425 ~ in the presence of H20. The product yields are also plotted against the reaction time in Fig. 4.51. The yields of MI, NP and BB at 400 ~ for 300 min in the presence of H20 were 0.9, 0.2 and 0.4%, respectively. These values were less than half those in the absence of H20. Under every condition in the absence and presence of H20, the ratios among MI, NP and BB did not differ significantly. In the reaction of tetralin with tritiated hydrogen in the presence of H20, the conversion and hydrogen exchange ratio of tetralin were also plotted against reaction time in Figs. 4.52 and 4.53. The conversions of tetralin and hydrogen exchange ratio in the presence of H20 at 400 ~ for 300 min were 1.7 and 1.3%, respectively. The hydrogen exchange ratio at 375-425 ~ in the presence of H20 was about one third of that in the absence of H20. First-order plots of these data for the c o n v e r s i o n and h y d r o g e n e x c h a n g e are also shown in Fig. 4.54. As shown in Tables 4.14 and 4.15, both the rate constants of the tetralin conversion and hydrogen exchange in the presence of H20 were smaller than those in the absence of H20 at 375-425 ~ Similarly, from the Arrhenius plots of the rate constants in the presence of H20 shown in Table 4.14, the activation energy of tetralin conversion in the presence of H20 was determined and was 35_+ 1 kcal/mol. This value was very close to that in the absence of H20, suggesting that the conversion of tetralin in the presence of H20 proceeded via the same reaction mechanism as that in the absence of H20 depended. Similarly, the activation energy of the hydrogen exchange in the presence of H20 was determined from Table 4.15 and was 2 4 _ + 1 kcal/mol. In contrast to the results from the conversion of tetralin, this value was different from that in the absence of H20. This re-

250

4 Liquefaction of Coal

suit shows that the hydrogen exchange of tetralin in the presence and absence of H 2 0 proceeds via different mechanisms. The reactions of tetralin, decalin and naphthalene with tritiated hydrogen were performed at 400 ~ in the presence of H2S. Although tritium was introduced to the solvents by hydrogen addition and hydrogen exchange reactions, most of the tritium was introduced through hydrogen exchange. The product yields from solvents in the presence and absence of H2S are plotted against reaction time in Figs. 4.55 and 4.56. In the reaction with decalin, tetralin by dehydrogenation and some unidentified products were formed. The yield of
0.5

0.4

.=
m

0.3

0.2

.,..~

0.1

J
O

0.0 0 100 200 300 400 Reaction time (min) Fig. 4.55 Effect of reaction time on the yield of tetralin in the reaction of decalin at 400 ~ [Reproduced with permission from Godo. M. et al., Fuel, 77, 949, Elsevier.(1998)] H2: O" HE/H2S"

100

200

300

400

Reaction time (min) Fig. 4.56 Effect of reaction time on the yield of tetralin in the reaction of naphthalene at 400 ~ [Reproduced with permission from Godo. M. et al., Fuel, 77, 949, Elsevier. (1998)] H:: O; H2/H2S: 9

4.4 HydrogenTransfer Reaction in Coal Liquefaction

251

4
r~

= ~D >
O O

= O
~D > O

o 0

~ 100 200 Reaction time (min) 300 400

Fig. 4.57 Effectof reaction time on the conversion of solvents at 400 ~ H2: O tetralin; A decalin; D naphthalene Hz]HzS: 9 tetralin; 9 decalin; 9 naphthalene [Reproduced with permission from Godo. M. et al., Fuel, 77, 950, Elsevier (1998)] tetralin from decalin in the presence of HzS at 400 ~ for 300 min was 0.2%, and that in the absence of HzS was also 0.2%. In the reaction with naphthalene, tetralin by hydrogenation was formed and the yields of tetralin in the presence and absence of HzS at 400 ~ for 300 min were 3.1 and 0.7%, respectively. HzS only affected the hydrogenation of naphthalene. In the reaction of solvents with tritiated hydrogen in the presence and absence of HzS, the conversions and hydrogen exchange rations (HERs) of solvents were plotted against reaction time in Figs. 4.57 and 4.58, and compared with those in the reaction in the absence of HzS. As shown in Fig. 4.57, the conversions of tetralin, decalin and naphthalene in the presence of HzS at 400 ~ for 300 min were 4.1, 1.0 and 3.1%, and those in the absence of HzS at 400 ~ for 300 min were 3.9, 0.9 and 0.7%, respectively. These values of tetralin and decalin were very close to each other, suggesting that HzS did not participate in the rate-determining step of the conversions of tetralin and decalin to produce the products. However, the conversion of naphthalene in the presence of HzS was about four times higher than that in the absence of HzS. This shows that the hydrogenation of naphthalene is promoted remarkably by HzS. As shown in Fig. 4.58, HERs for tetralin, decalin and naphthalene in the presence and absence of HzS gradually increased with time. The HERs for tetralin reached 40.4 and 4.6%, at 400 ~ for 300 min, respectively, in the presence and absence of HzS. HERs for decalin reached 11.7 and 1.5%, respectively. The hydrogen exchange ratios of tetralin and decalin in the presence of HzS were about 10 times higher than those in the absence of H2S. However, HERs for naphthalene in the presence and absence of HzS were 22.9 and 7.7%, at 400 ~ for 300 min, respectively. HERs for naphthalene in the presence of HzS was about three times higher than that in the absence of HzS. The extent of the promotion effect of HzS on the hydrogen exchange in naphthalene was nearly equal to that of the conversion of naphthalene, and was less than those of the hydrogen exchange in tetralin and decalin. As a result, in the presence for HzS, the HER for tetralin was the highest among the three solvents. Under coal liquefaction conditions, it is well known that HzS is produced by the re-

252

4 Liquefaction of Coal 50

40
O .,..~

30

20

10

B~------'---'r-""'-5"-'-'~, i 0 100 200 Reaction time (min)

i 300 400

Fig. 4.58 Effect of reaction time on the hydrogen exchange ratio of solvents at 400 ~ H2: O tetralin; A decalin; I-1 naphthalene Hz/H2S: 9 tetralin; 9 decalin; 9 naphthalene [Reproduced with permission from Godo. M. et al., Fuel, 77, 950, Elsevier (1998)]

duction of pyrite in coal. It is suggested that compared with decalin and naphthalene the hydrogen mobility of tetralin is the highest under coal liquefaction conditions9 The rate constants of conversion and hydrogen exchange in the presence and absence of H2S were determined from the slopes of first-order plots and summarized in Tables 4.16 and 4.17. In these tables, k2/kl shows the ratio of the rate constants in the presence of H2S
Table 4.16 Rate Constants for Conversion of Tetralin and Formation of Decalin and Naphthalene (min-~) Reaction Tetralin Absence of HzS: kl
P r e s e n c e of H2S: k2

Solvent Decalin 3.08 X 10 -5 3.39 X 10 -5 1.1 Naphthalene 2.44 X 10 -5 1.06 X 10 -4 4.3

1.15 X 10 -4 1.19 X 10 -4 1.0

k2/kl

[Reproduced with permission from Godo. M. et al., Fuel, 77, 951, Elsevier (1998)]

Table 4.17 Rate Constants of Hydrogen Exchange Reaction at 400 ~ (min-1) Reaction Tetralin Absence of H2S" kl Presence of H2S" k2
kz]kl

Solvent Decalin 5.07 X 10 -5 4.37 X 10 -4 8.7 Naphthalene 2.66 X 10 -4 8.31 X 10 -4 3.1

1.65 X 10 -4 1.97 X 10 -3 12

[Reproduced with permission from Godo. M. et al., Fuel, 77, 951, Elsevier (1998)]

4.4 Hydrogen Transfer Reaction in Coal Liquefaction 800000 = "~ 600000 Ca)

253

400000
o

200000
.....

,~
Feed 0 200 400 600 800 Reaction time (min)

800000

"~ 600000
= o

~
Feec~ 0
,

(b)

400000
o

200000

r ir"~~~ll

200 400 600 Reaction time (min)

800

800000

(c)

"~ 600000
=

400000
o

200000
".~, [--.

0 Feed 0 200 400 600 Reaction time (min) 800

Fig. 4.59 Effect of reaction time on the tritium concentration in the presence of H/S. Reaction Temperature: a) 350 ~ b) 375 ~ c) 400 ~ O tritiated hydrogen; A HzS; [-] tetralin. [From Godo. M. et al., Energy Fuels, 11,474 (1997)]

and in the absence of H2S. This value represents the extent of the promotion effect of H2S on the conversion and hydrogen exchange reaction. The promotion effects of H2S on the hydrogen exchange in solvents increased in the order of naphthalene < decalin < tetralin.

254

4 Liquefaction of Coal

800000
600000

(a)

"----0--

400000
O

200000 o

,~
Feed 0

/
, II- I , I ~ ,

100

200

300

400

Reaction time (min)

800000

~.

(b)

~,

600000

O-

400000
O

~ 200000
o~

Feed 0

100 200 300 Reaction time (min)

400

800000
600000
O

oQ.,,.,..

(c)

400000
O

E 200000
.... j

0
Feed 0 100 200 300 Reaction time (min)

400

Fig. 4.60 Changes in the tritium concentration with reaction time in the presence of H20. Reaction Temperature: a) 375 ~ b) 400 ~ c) 425 ~ O tritiated hydrogen; A H2S; F-] tetralin. [From Godo. M. et al., Energy Fuels, 11,475 (1997)]

Further, k2/kl of the conversion of naphthalene is close to that of the hydrogen exchange. Changes in the tritium distribution in the presence of H2S among gaseous hydrogen, solvents and H2S at 400 ~ with reaction time are shown in Fig. 4.59a,b,c. In these figures,

4.4 HydrogenTransfer Reaction in Coal Liquefaction

255

solid horizontal lines show the equilibrium value which was calculated on the assumption that hydrogen atoms were completely scrambled between gaseous hydrogen, solvents and HzS. The tritium initially included in gaseous hydrogen decreased monotonically with the passage of time. On the other hand, the tritium concentration in solvents increased monotonically with time. The tritium concentration in HzS of the each solvents system increased rapidly, showed the maximum value beyond the calculated equilibrium value at 0 min and reached a very close value to that in gaseous hydrogen, and decreased with the same tendency as gaseous hydrogen molecules. This result shows that the hydrogen exchange reaction between gaseous hydrogen and H2S proceeded rapidly at the initial reaction stage and reached equilibrium between the two phases. At 300 min, the tritium distribution among gaseous hydrogen, solvents and HzS approached the calculated equilibrium value. Variation in the tritium distribution in the presence of H20 among gaseous hydrogen, tetralin and HzO at 375, 400 and 425 ~ with reaction time are shown in Fig. 4.60 (a-c). Since the rate of hydrogen exchange between gaseous hydrogen and tetralin in the presence of H20 was far smaller than that in the presence of HzS, the decreases of tritium in gaseous hydrogen and the increases of tritium in tetralin solvent were slower than those in the presence of HzS. The tritium concentration in H20 at every temperature increased gradually and was close to that in gaseous hydrogen at 300 min. However, the increasing rate of tritium concentration in H20 was obviously slower than that in HzS. It was assumed that the hydrogen exchange for tetralin in the absence of H2S proceeded via a tetralyl radical formed by unimolecular dissociation, which also acted as an intermediate in the conversion of tetralin, as shown in Eqs. (4.24) and (4.25) (Skowronski et al., 1984; Kabe et al., 1991a; Ishihara et al., 1995; Godo et al., 1997b, c). If the tetralyl radical is quenched by hydrogen radical (H.), which is formed from original tetralin by the thermal dissociation, the hydrogen exchange of tetralin will not proceed. Ttie reaction of the tetralyl radical with tritiated hydrogen leads to hydrogen exchange in the absence of H2S. The tritium distributions of HzS in Fig. 4.59 (a-c) show that the hydrogen exchange reaction between tritiated gaseous hydrogen and HzS proceeds rapidly under the coal liquefaction conditions. A possible mechanism of the hydrogen exchange reaction between gaseous hydrogen and H2S is shown in Eqs. (4.50) and (4.51). In Eq. (4.49), HzS produces H. and HS. radicals by the thermal dissociation under these reaction conditions. The formed H. will react with a tritiated hydrogen molecule and produce a tritium radical, as shown in Eq. (4.50). The reaction of a tritium radical (T.) with HS- leads to hydrogen exchange between tritiated gaseous hydrogen and HzS as shown in Eq. (4.51). The HER between tetralin and tritiated hydrogen in the presence of HzS was about ten times higher than that in the absence of HzS. In the case of hydrogen exchange of tetralin in the presence of HzS, H- produced from tetralin via Eq. (4.24) or from H2S via Eq. (4.49) changes into T- through Eqs. (4.50) or (4.52). Eq. (4.52) is considered to be much faster than Eq. (4.50), because the bond dissociation energy of hydrogen sulfide (370 kJ/mol) is much smaller than that of hydrogen (432 kJ/mol). Stenberg et al. (1983) reported the hydrogen donor ability of HzS by the stoichiometry of the biphenyl conversion and the favorably low bond dissociation energies of HzS compared with H2. The bond energy of H2 is greater than that of most C-H bonds whereas that for H2S is not. It is assumed that the concentration of T. increases remarkably in the presence of HzS. The most likely and energetically feasible early initiation step is the addition of T. into tetralin, promoting the formation of a hydrotetralyl radical and the hydrogen exchange in an aromatic ring, as shown in Eq.(4.53). The hydrogen exchange reaction as shown in Eqs. (4.45) and (4.46) may also occur in the presence of HzS. In the reaction of tetralin, decalin was not formed by hydro-

236

4 Liquefaction of Coal

genation. It is assumed that the hydrogenation of a hydrotetralyl radical leading to decalin does not proceed. It is also assumed that H2S plays an important role as a promoter of the hydrogen exchange between tritiated gaseous hydrogen and tetralin. The main product in the conversion of decalin was tetralin by dehydrogenation. The conversion in decalin was not affected by H2S. However, the hydrogen exchange was remarkably promoted in the presence of H2S. We assumed that the formations of tetralin and the hydrogen exchange in decalin proceeded with a radical reaction mechanism. Decalin may collide with tritium radical and produced radicals as shown in Eq. (4.54). Tetralin would be produced from decalyl radical through the intermediate which is formed by dehydrogenation. The reaction of the decalyl radical with tritiated hydrogen leads to hydrogen exchange as shown in Eq. (4.55). In the presence of H2S, tritium radical is produced more easily through Eq. (4.52) as the hydrogen exchange reaction of tetralin system.
H2S

HS, + + + H, HS, H,
H2 +

H,
T9

(4.49) (4.50) (4.51)

HT T, HTS

HTS
H2S

T,
T

(4.52)

Ho

(4.53)

+ To
H

- ~

+ HT

(4.54)

+ HT

. ~ H +

Ho

(4.55)

Both conversion and hydrogen exchange reaction in naphthalene were promoted about four times in the presence of H2S. It was suggested that the rate-determining steps of the conversion and hydrogen exchange reaction were the same and that H2S promoted the rate determining step. The product in the conversion of naphthalene was tetralin by hydrogenation. Tetralin would be produced by the multi-steps of hydrogenation, as shown in Eq. (4.55). Tritium was introduced to the tetralin by hydrogenation. However, the amount of tritium added was very small in comparison with the amount of tritium exchanged. In the presence of H2S, atritium radical is easy to generated and therefore the tritiated naphthalyl radical is also easy to be generated. The tritiated naphthalyl radical reacts with hydrogen atom to lead to dihydronaphthalene, which converts into tetralin. Further, HS. also abstracts the hydrogen atom in the tritiated naphthalyl radical reversibly to produce tritiated naphthalene, promoting the hydrogen exchange. Therefore, it is assumed that the common rate-determining step of the conversion and the hydrogen exchange of naphthalene is the formation of the naphthalyl radical. This mechanism would explain the same degree of the promotion effect of H2S on the conversion and hydrogen exchange. The hydrogen addition and exchange reactions between tetralin and gaseous hydrogen, and the formation of naphthalene and 1-methylindan from tetralin are considered to proceed via a radical reaction mechanism (Kabe et al., 1990d; Ishihara et al., 1995; Godo et al.,

4.4 HydrogenTransfer Reaction in Coal Liquefaction T ~ + To . ~ + H.

257

(4.56)

IHo
+HT H HH T 1997b), where a tetralyl radical was an intermediate in the hydrogen exchange and conversion of tetralin. Therefore, the formation of the tetralyl radical in this system was assumed to be the rate-determining step for both the hydrogen exchange and the conversion of tetralin. In the system of tetralin and gaseous hydrogen, tetralin may collide with not only itself but also tritiated hydrogen. It is possible that tritiated hydrogen affects the step of the formation of a tetralyl radical. If tritiated hydrogen affects the formation of the tetralyl radical, the hydrogen exchange ratio (HER) and conversion of tetralin will change by the partial pressure of gaseous hydrogen. As we know, it is difficult to control the reaction pressure in the reactions using the autoclave because the effective reaction pressure changes depending on reaction temperature. Further, an autoclave experiment takes longer than ca. 30 minutes to heat the system to 400 ~ the effective temperature for coal liquefaction. Hydrogen transfer and conversion of tetralin occur while the autoclave is heated to the set temperature. This may not be a major problem for comparative experiments but would be unsatisfactory for kinetic studies, especially in the short reaction time. In contrast, a flow type reaction system can control the reaction pressure strictly and is suitable for the short reaction time. Here, the hydrogen exchange reactions between tetralin and gaseous hydrogen were investigated in a flow type reactor to estimate the effect of reaction pressure and the kinetics in the short reaction time (Godo et al., 1997c). The reactions of tetralin with tritiated hydrogen were performed under the conditions 400-450 ~ and 2.5-9.8 MPa for 25-420 sec. The yields of products at 425 ~ are plotted against the reaction time in Fig. 4.61. In this temperature range, the reaction products from tetralin were 1-methylindan by isomerization as shown in Fig. 4.61a), naphthalene by dehydrogenation as shown in Fig. 4.61b), and nbutylbenzene by hydrocracking as shown in Fig. 4.61c); the main product was 1-methylindan. Decalin was not formed under these conditions. This is consistent with the previous result in the reaction using the autoclave (Godo et al., 1997b), in which the disproportionation of tetralin to decalin does not occur. The products increased monotonically with residence time, and the yields of methylindan, naphthalene and butylbenzene at 425 ~ for 450 reached 0.13, 0.05 and 0.06%, respectively. As shown in Fig. 4.61 a), the plots of methylindan at 2.5, 4.9 and 9.8 MPa showed approximately the same straight line and were not influenced by the reaction pressure. A similar result was also obtained for naphthalene as shown in Fig. 4.6 l b). The results in the reaction using the autoclave reported are also plotted in Fig. 4.61. Those reactions were conducted under the conditions at initial pressure 5.9 MPa for 0-300 min. In these figures, the ratios between the reaction time and the yields of products in the reactions using the autoclave are the same as those in the reactions using the flow reactor. So the relationship between the reaction time and the product yield can be

238

4 Liquefaction of Coal

Reaction time 10 -4 (s) 0.20


a

1
!

2
|

3
|

4
|

20.0

0.15

15.0

.=,
0.10

[]
10.0

-~ 0.05
~D

5.0 _
: I :

~D

0.00 t

0.0

~, 0.06
~D

6.0

-fi 0.04

[]

4.0 "fi

0.02

2.0 __. .,.9, i , I , 0.0

0.00

c
~0.06

[]

[]
6.0
N

0.04

4.0 .~
m

0.02

2.0

0 ~D .,..q

0.00 ( 0

0.0 500

100

200

300

400

Residence time (s) 0 2.5 MPa; Fig. 4.61 A 4.9 MPa; [] 9.8 MPa; 9 autoclave

Effect of residence time on the yield of products at 425 ~ a) methylindan; b) naphthalene; c) butylbenzene [From Godo. M. et aI.,AIChE, 43, 3108 (1997)]

compared within the same figure. The yields of methylindan and naphthalene in the reaction using the autoclave were slightly less than those of the flow reactor. However, these values were not affected significantly by the type of reactor. In the reaction using the autoclave, although the initial pressure was 5.9 MPa, the effective reaction pressure was about 14.7 MPa at 425 ~ by thermal expansion of gas phase. The result shows that the yields of methylindan and naphthalene are not affected considerably by the reaction pressure in the range from 2.5-14.7 MPa. In contrast, the yields of butylbenzene in the reaction in the

4.4

Hydrogen Transfer Reaction in Coal Liquefaction

259

flow reactor increased with rise of reaction pressure, and the formation of butylbenzene depended on the reaction pressure. The yield of butylbenzene in the reaction using autoclave also increased monotonically with reaction time because the reaction pressure of autoclave was nearly equal when the reaction temperatures were the same, and were close to that in the reaction of a flow reactor at 4.9 MPa. The amount of tritium introduced into tetralin from gas phase is represented by the hydrogen exchange ratio (HER). Fig. 4.62 shows the change in HER with reaction time. The HER of tetralin increased gradually with lapse of time and reached 0.11% at 425 ~ for 450 s. However, at 400 ~ the hydrogen exchange ratio was only about 0.04 % for 520 s. Further, the plots of HER could be shown with the same line in the pressure range from 2.5 MPa to 9.8 MPa. In addition, HER in the reaction using the autoclave was almost the same as that in the flow reactor. The result is similar to that of the conversion of tetralin into methylindan and naphthalene.
Reaction time 10 -4 (s) 0 0.15] , 1 i , 2 i , 3 i , 4 t 5 , 15.0

.o

.0.10

D
10.0

..a

Ca)

ca0

0.00

5.o ~

0.00

.t

'

0.0 500

100 O 2.5 MPa;

200

300

400 9 autoclave

Residence time (s) A 4.9 MPa; [] 9.8 MPa;

Fig. 4.62

Effect of residence time on the hydrogen exchange ratio of tetralin at 425 ~ [From Godo. M. et al.,AIChE, 43, 3108 (1997)]

First-order plots of these data for the formation of methylindan and naphthalene, and hydrogen exchange are shown in Fig. 4.63. All plots approximately fit a linear relationship, indicating that these reactions could be treated as first-order reactions. The formation rate constants of methylindan from tetralin could be determined from the slopes of first-order plots. These constants are shown in Table 4.18. In the same way, the formation rate constants of naphthalene, and the rate constants of hydrogen exchange were obtained and listed in Table 4.18. From the Arrhenius plots, activation energies of the tetralin conversions into methylindan and naphthalene, and hydrogen exchange between tetralin and tritiated hydrogen were determined and were 3 2 _ 2, 33 --_+2 and 33 +__2 kcal/mol, respectively. Activation energies in the reaction using the autoclave are also determined and are listed in Table 4.19. These values were very close to each other, suggesting that the conversion of tetralin into methylindan and naphthalene, and hydrogen exchange reaction of tetralin proceeds via the same reaction mechanism.

260 0.000

--0.001
x

,7,, --0.002 ,1 --0.003

OA

0.0000 --0.0002 -0.0004


,,,,,.t

--0.0006 -0.0008 --0.000


,-~ ,

--0.001 X
!

--0.002 .d --0.003
(c)

--0.004

200 400 Residence time (s)

600

Fig. 4.63 First-order plot of conversion of tetralin and HER. (Godo et al., 1997c) a) conversion of tetralin into methylindan; b) conversion of tetralin into naphthalene; c) HER. [From Godo. M. et al., A1ChE, 43, 3109 (1997)] Table 4.18 Rate Constants in the Reaction of Tetralin with Tritiated Hydrogen Temperature (~ Conversion of tetralin Into methylindan (Xl0 -6 sec -l) 400 425 450 1.6 + 0.3 0.5 -+- 0.1 1.1 _ 0.2 Into naphthalene ( x l 0 -6 sec -1) 3.7 + 0.7 1.0 + 0.2 2.7 --+ 0.5 Hydrogen exchange of tetralin (x10 -6 sec -1) 8.3 + 1.7 2.8 0.6 6.5 + 1.3

[From Godo. M. et al.,AIChE, 43, 3109 (1997)]

4.4 HydrogenTransfer Reaction in Coal Liquefaction

261

Table 4.19 Summaryof Yields from Various Direct Coal Liquefaction Processes Involving Hydrogenation Process Coal used Scale of operation (t/d) Hydrogenation temperature (~ Pressure conditions (MPa) Yields (%coal) Heterogas Hydrocarbon gas Naphtha Mid-distillate Heavy oil Residues (ash and unconverted coal) Hydrogen consumption wt% coal feed Reference SRC II Illinois 2.5 370 18 11.7 6.4 8.4 6.4 40.0 33.0 3 H-coal ExxonEDS Kohloel Illinois 250 450 20 12.5 10.5 13.3 19.7 31.0 37.1 4 BCL NEDOL Tanitoharum 150 455 17 Illinois Bituminous Morwell 200 200 50 410 460-470 450 15 30 15 12.0 5 14 10 11 48 4 11 21 16 36 22 16 6 5

Newman Newman Neavel 1985 1985 1961

Langhoff NEDO/NBCL Wasaka 1982 1994 1999b

It is assumed that the formation of methylindan and naphthalene, and the hydrogen transfer from the tetralin, proceeded via a tetralyl radical, which acted as an intermediate in the conversion and the hydrogen exchange of tetralin (Kabe et al., 1990d; Ishihara et al., 1995; Godo et al., 1997b). In the system of tetralin and gaseous hydrogen, the conversion into methylindan and naphthalene, and hydrogen exchange reaction of tetralin were not changed by the reaction pressure, indicating that gaseous hydrogen does not affect the formation of a tetralyl radical at least in the present reaction system using a flow reactor. Tetralin may be activated by collisions and become a tetralyl radical and a hydrogen atom according to Eq. (4.24). Some of the hydrogen atoms produced from tetralin would react with tritiated hydrogen molecular, leading to the formation of the tritium atom, as shown in Eq. (4.50). If the tetralyl radical is quenched by tetralin to form original tetralin, the conversion of tetralin and hydrogen exchange would be inhibited. The reaction of the tetralyl radical with the tritium atom leads to hydrogen exchange in the reproduction of tetralin, as shown in Eq. (4.57). The hydrogen exchange between tetralyl radicals and tritiated hydrogen (Eq. (4.46)) can proceed via radical hydrogen transfer reaction depending on the concentration of tetralyl radicals, which also controls the formation of methylindan and naphT @ + .T 9 H @ (4.57)

+ H.

(4.58)

thalene in Eqs. (4.26) and (4.58). Therefore, the formation of the tetralyl radical by unimolecular scission in this system may be the rate-determining step for both the conversion of tetralin into methylindan and naphthalene, and the hydrogen exchange. At 400 ~ the conversion and hydrogen exchange ratio of tetralin were very low. If coal is included in the

262

Liquefaction of Coal

system, a tetralyl radical may be formed easily. However, if coal which forms a radical is not included in the system, a tetralyl radical is difficult to form. Then the unimolecular scission of tetralin barely occurs at 400 ~ Hooper et al. (1979) reported that butylbenzene was formed by the thermal dissociation of tetralin, as shown in Eq. (4.59). While the plot of the yields of butylbenzene showed some scatter, the effect of the reaction pressure was recognized at every temperature. An alternative route to form butylbenzene appears in Eq. (4.27), where tetralin reacts with hydrogen atom.

4.5 Process of Coal Liquefaction


Coal liquefaction processes generally have as the principal objective the manufacture of transport fuels, and many have the ambitious target of producing premium fuels that can be directly substituted for the pump grades currently obtained from petroleum. The same philosophy of direct substitution can be adopted for the petrochemical industry using feedstocks for the massive aromatics/olefins plants currently producing most of the precursors for today' s synthetic materials. Methods for the direct liquefaction of coal have been developed in a number of countries and the processes in the United States, Germany and Japan are described here. 4.5.1 C o a l L i q u e f a c t i o n P r o c e s s e s in the U S A

A. Solvent Refined Coal (SRC-II) This process was originally developed by the Pittsburg and Midway Coal Company for the production of pure carbons from coal. It was later adapted as a method for desulfurizing the coals that were available from vast reserves in the US Appalachian coal basin. The American Clean Air Act of 1977 limited the sulfur content in the Appalachians to 2-5% sulfur. The SRC product is a pitch-like product and needs further hydrotreatment to make useful distillates. In the latest version of the SRC process being developed by the Catalytic Corporation (Fossil Energy, DOE/PC/50041-(79)) at their Wilsonville pilot plant, the SRC primary product is being hydrocracked to produce low boiling distillates. A line diagram of this process is shown in Fig. 4.64. Briefly, pulverized coal, slurried with recycle oil, is
Hydrogen

-~
High pressure hydro-extraction Solvent recycle Hydrogen Light solvent recovery

I
Critical solvent de-ashing

Ash concentrate

Coal

Slurry preparation

I
I

I
Catalytic hydrocracking J ebullating bed
,qm

I
Clean SRC solution

I
Product solvent separation

Gases

I I

Distillate product

Recycle solvent

Fig. 4.64 Two-stage SRC coal liquefaction. [Reproduced with permission from Davies, G.O. et al., Critical Report on Applied Chemistry Volume 9 Chemicals from Coal: New Developments, 102, Backwell Sci. Pub., (1985)]

4.5 Processof Coal Liquefaction

263

passed with high pressure hydrogen to a digestor where the coal is almost totally dissolved. After separation of the gas for recycle the digest is cleaned by critical solvent de-ashing. After de-ashing the SRC material is hydrocracked in an ebullating bed of a catalyst. A typical set of product yields from Illinois # 6 coal is given in Table 4.19. In the process, a portion of ash containing pyrite (FeS2) as a catalyst with recycle oil was cycled.

B. H-coal Process This process was developed by the Hydrocarbon Research Incorporation from the H-oil technology used commercially for the beneficiation of heavy petroleum residues. The heart of the process, illustrated in Fig. 4.65, is the ebullating bed hydrocracker where a slurry of coal in recycle oil is catalytically reacted with high pressure hydrogen. The high active CoMo particle catalyst prepared was used. The up-flow of the slurry expands the catalyst bed, while the flow of hydrogen induces a washing-machine mixing action. The net result is free movement of the catalyst pellets that prevents plugging by ash and carbon deposits, produces isothermal reaction conditions and facilitates the addition and removal of catalyst in the reactor. The hydrocracked products pass to a separation system where light distillates are recovered as product and the heavier products are split by vacuum distillation into heavy oil for recycle and an ash-laden pitch that can be used as a fuel for power generation. The process was developed at the Trenton Laboratories with units processing up to 3 ~ t of coal per day (Comolli et al., 1978). Data from these units were used to design a 250 t/d plant that was built at Catlettsburg, Kentucky. This large plant has been operating since 1981 and as the work has been sponsored by the US government. The results have been reported in a number of US Department of Energy documents (Fossil Energy, DOE/ET 10143-19/37). A yield summary from a typical run at Catlettsburg using a Kentucky coal is given in Table 4.19.

Reactor Coal (f "

Naphtha __~ Hydrotreating and r reforming

E:

Gasoline chemicals

::g

/t

Recycle I [I

tube "1 i
~ I

Naphtha ? ~ Hydrotreating , and ::::::::.~ Mid-distillates reforming distillates Sulphur(36 kg) Ammonia(9 kg) High-Btu-Gas (3200cu.ft.)

Hydr~ manufacture

I | ' ~ sHY~a~ n

Lt

Hydrogen

Jj

Fig. 4.65 H-coalprocess. [Reproducedwith permissionfrom Davies, G.O. et al., Critical Report on Applied Chemistry Volume 9 Chemicals from Coal: New Developments, 105, Backwell Sci. Pub. (1985)]

264

4 Liquefaction of Coal

C. Exxon EDS Process This direct coal liquefaction process has been piloted by the American Exxon Company at its Baytown Refinery in Texas (Exxon EDS, 1981). The stages of the process are shown in Fig. 4.66. Coal is slurried with hydrogenated recycle oil. The coal-oil slurry is passed via a preheater to a simple reactor at 410 ~ together with high pressure hydrogen (150 bar). In later versions of the Exxon process (Neavel, 1961) a proportion of the ash residue is recycled as it is known that coal ash can catalyze the hydrogenation reactions taking place. Hydrogenation of the pyrolyzing coal is effected by direct transfer from the hydrogen donor solvent and by shuttle transfer from molecular hydrogen present. Typical yields from the process are given in Table 4.19.
Catalytic [ Solvent i hydrogenation _~ Slurry ~_~ preparation H2

I Coal
preparation

Solvent

7"
Distillation I Heavy bottoms slurry _1 ~[ Flexicoking -I Ash residue

Gas - Liquid products

H2 ~

Liquefaction ~

H20 Air

~ Fuel gas

Fig. 4.66 Exxon EDS process. [Reproduced with permission from Davies, G.O. et al., Critical Report on Applied Chemistry Volume 9 Chemicals from Coal: New Developments, 111, Backwell Sci. Pub. (1985)]

4.5.2

C o a l L i q u e f a c t i o n P r o c e s s e s in G e r m a n y

German coal liquefaction processing is being developed by Ruhrkihle Oel und Gas Gmbh, and is based on the original liquefaction work (IG process) by Bergius and Billiviller (1918) and Pier (1929). The process (New IG process) has been described by Peters (1978) and Romey (1981) and has reached a large pilot plant (250t/d) scale of operation at the Bottrop plant near Essen (Langhoff et al., 1982). The process is illustrated in Fig. 4.67. Briefly, coal is slurried in heavy recycle oil and is passed with high pressure hydrogen (30 MPa) and a cheap throw-away catalyst (Fe-sulfur catalyst) to a simple tubular reactor where, at high temperature (460 ~ in the presence of high pressure hydrogen, the coal is hydropyrolyzed to produce distillates. The distillates are fractionated in a series of pressure let-down vessels and a vacuum distillation is used finally to separate the unreacted coal and ash from the heavy oil needed for recycle. Typical yields from the Bottrop plant are given in Table 4.19. The plant at Bottrop has been operating since November 1981 and it was reported by Langhoff (1982) that more than 1000 h of operation had been achieved by March 1982. The upgrading of the Kohloel syncrude is being investigated by VEBA OEL in bench-scale units. The quality of the various distillate products is similar to those obtained from SRC and H-coal syncrudes.

4.5 Process of Coal Liquefaction Coal + catalyst


Gas
,..._

265

"-

I [ txl ~00 bar ~75 ~


i

Light oil ~ Middle oil ~ Heavy oil

Preheater

, "-~--'Zt Reactors I ]

[ I

Vacuum L istillati~

r
Hydrogen Ash ~ Recycle oil I Gasification I I

1
Heavy residues+ash

Fig. 4.67 German coal liquefaction process. [Reproduced with permission from Davies, G.O. et al., Critical Report on Applied Chemistry Volume 9 Chemicals from Coal: New Developments, 113, Backwell Sci. Pub. (1985)]

4.5.3

C o a l L i q u e f a c t i o n P r o c e s s e s in J a p a n

The Agency of Industrial Science and Technology (AIST) of the former Ministry of International Trade and Industry (MITI) of Japan started the Sunshine Project in 1974 to develop new energy technologies (Yoshida, 1997; Office of New Sunshine Project, 1995). The Moonlight Project was started in 1978 to develop energy conservation technologies. Both projects successfully maintained Research and Development (R&D) schedules for energy subject areas with the close cooperation of industry, government and academic organizations. These projects have been steadily providing effective results for basic technology, practical applications and application to peripheral fields. The New Energy and Industrial Technology Development Organization (NEDO) has been established to develop coal liquefaction technology. NEDO started basic research for coal liquefaction development in 1980 and subsequently developed the NEDOL coal liquefaction process (Wasaka, 1999a; Wasaka and Ibaragi, 2000). The economics study of the NEDOL coal liquefaction plant, for commercial scale-plants, was conducted after a 150 t/d pilot plant operation had been completed. For coal liquefaction technology, R&D is being carried out on brown coal and bituminous coal liquefaction technology. These are technologies for liquefying coal and manufacturing transportation fuel that can be used in place of petroleum. A. Brown Coal Liquefaction Technology Brown coal liquefaction technology is intended realize the liquefaction of brown coal of Victoria, Australia, with the object of utilizing it as an energy resource. Research and Development was started in 1981 with a 50 t/d pilot plant in Victoria province. Fig. 4.68 shows the process. The initial goals of the project, 50% liquid fuel yield (actual result was 52%) and long-term operation with duration of 1000 hours (actual result was 1700 hours) were attained, and the research program was successfully finished in FY 1990. Based on

266

4 Liquefaction of Coal

Primaryhydrogenation H2
Raw brown coal Catalyst de-watering Removed water Reactor

Slurry )

I Recycle gas compressor [ Recycle gas [ ~ Off gas ] [purificationI

~- Purgegas (Fuel gas)

Separator ~ o r

(Lightoil ) ATM,fractionator I
Middleoil

Recycle solvent

Ballmill [_..... I Slurry making I


CLB recycle feed pump

Secondary hydrogenation Recycle gas

I rePrSs:iU~g1

~c~

Lightoil )

~ Solvent recycle from ATM fractionator Hydrotreatedde-ashedoil (HDAO) recycle

SepTat~

C~

Settler Fixedbed reactor I~ Sludge Fractionator CLB run down I

Fig. 4.68

Brown coal liquefaction (BCL) process. [Reproduced with permission from Shimasaki. K. et Jpn Inst. Energy, 78, 809 (1999)]

al., J.

the analysis of the results obtained by the operation of the pilot plant, the conceptual design of a 6000 t/d demonstration plant and an economic evaluation of the process were completed by the end of FY 1993 (NEDO/NBCL, 1994). Typical yields are given in Table 4.19. B. Bituminous Coal Liquefaction Technology As for bituminous coal liquefaction technology, the R&D program of the NEDOL process started in 1984. Fig. 4.69 shows a flow diagram of the NEDOL process (Wasaka, 1999b; Wasaka et al., 2003). The NEDOL process liquefies coal by using a Fe-based catalyst and hydrotreated solvent under relatively mild reaction conditions of 430--460 ~ and hydrogen pressure of 15-20 MPa. The initial goal for yields of light and medium oil such as gasoline and diesel oil (boiling range = C4-350 ~ is 50% or higher. Moreover, this process is applicable to a wide range of coal types from low-rank bituminous coal to low-rank subbituminous coal. The design of a 150 t/d pilot plant was started in 1988, and construction was started in 1991. Construction work was completed at the end of June 1996. Test runs were completed during 1997 to 1998 (Makino and Ueda, 1996). The actual liquid fuel yield was 58% and long-term operation with duration of 1920 hours was attained, and the research program was successfully finished in autumn, 1998. Typical yields are given in Table 4.19. 4.5.4 Present Status and Future o f Direct C o a l L i q u e f a c t i o n

As described above, various direct liquefaction processes are being developed and large commercial coal liquefaction plants of a capacity of 10,000 t/day have reached the conceptual design stage. Many direct coal liquefaction processes have been shown to be technically viable in plants processing up to 150-250 t/day. The thermal efficiency of conversion has proved to be high (--65%) showing a considerable advantage over gasification synthesis routes. Laboratory tests have shown that the distillates produced can be used as substitutes for the petroleum factions used for making synthetics. The processing required is complex, involving high temperatures, high pressures and catalysts, and the conversion consumes 4-7 % w/w of hydrogen and is consequently expensive. Although cheap coal is available

4.5 Processof CoalLiquefaction Coalpreparation


.............................. ! . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

267

Liquefaction

Catalyst I Coal ~

Hydr~

Distillation Atmospheric Separators~ _ ,., tower ~V~ ~--~

Reactors ~ L Pulverizer I Letdown -4-~l-~JPreheater valve


,

9 Gas --Naphtha - Gas oil

Vacuumtower I
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . j

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Hydr~ ~ara solvent i ~ Stripper i " ] ~ !

~ tors

"~-~ IHydr~ l-_C..'), Recycle solvent


j

Residue - Naphtha
.

~ ~

T ,,l,:~! Reactor Preheater

9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Solvent hydrogenation Fig.4.69 A flowdiagramof the NEDOLprocess. [Reproduced withpermission fromWasaka.S. et al.,J. Jpn. Inst. Energy, 78, 802 (1999)]
in many countries the overall cost of producing light distillates from coal is still considerably higher than their production from petroleum. There are numerous problems to be solved when considering future commercial coal liquefaction processes. First, there is the matter of hydrogen production. Much hydrogen is consumed in any coal liquefaction process corresponding to the production scale and the nature of the liquefied product. The issue of hydrogen manufacture in the process remains open to debate. Most of the methods currently under consideration involve the partial oxidization of the coal or use of residues after liquefaction. Among the conversion processes, the liquefaction process is most heavily influenced by the nature of the coal. However, there exists controversy regarding the chemical and macromolecular structure of coal despite recent advances, as described in Chapter 2. Thus, the exact roles of catalyst and solvent in coal liquefaction remain ambiguous. Design in the coal liquefaction process is based mostly on experience.

Você também pode gostar