You are on page 1of 6

Ind. Eng. Chem. Res.

1992,31,1835-1840

1835

REVIEWS
Desulfurization of Petroleum Coke: A Review
Hassan Al-Haj-Ibrahim and Badie I. Morsi* Chemical and Petroleum Engineering Department, University of Pittsburgh, Pittsburgh, Pennsylvania 15261

A number of processes ranging from purely physical such as solvent extraction to thermal or chemical treatment using different agents including hydrocarbon gases, alkali metal compounds, and hydrogen have been proposed for petroleum coke desulfurization. The complexity of the sulfur-carbon bonds in coke, together with the extremely variable nature of the coke itself, have precluded a theoretical study with the emphasis being laid on empirical data as a consequence. The conclusion that has emerged from this review is that a thermal treatment using a suitable agent is essential for effective desulfurization at temperatures less than 1100 K.
Introduction Petroleum coke, or petcoke for short, is no longer a left-over by-product of the bottom-of-thebarrel refinery processes whose chief purpose is the production of other products. Petcoke has become a valuable product in its own right, and the demand for high-quality low-sulfur coke is increasing. However, more coke with high sulfur content is being produced, and means whereby such sulfur content is reduced to an acceptable level or eliminated altogether are called for, in particular with the ever-tightening restrictions on sulfur oxide emissions for environmental considerations. Petcoke can be produced from virgin crude residues by precipitation reactions of high molecular weight compounds, asphaltenes, and resins or from highly aromatic t a r or decanted oil stokes by condensation and polymerization of aromatic compounds. Delayed petcoke is produced by a semicontinuous process which can be carried through in one of the following ways: ultimate, oncethrough, or intermediate coking (1). Delayed petcoke has a much more uniform crystallinity than other types of petcoke since the delayed coking process allows the time needed for the coke crystals to orient themselves upon one another ( 2 ) .Green coke is made of petroleum pitches and residues by any of several coking processes, chief of which are delayed coking, contact coking, fluid coking, and flexicoking. Green delayed coke is a petcoke which has not been calcined for the removal of moisture and excess volatile matter (3). The pores in the green coke are filled, probably with a hardened residuum from the coker feed which distills off during calcination exposing the pores and typical lamellar structure of the calcined coke. The main types of green delayed petcoke include needle coke, honeycomb, sponge coke, and shot coke (3-8). The green coke has a low ash content between 0.3 and 0.5 wt % ,and ita fixed carbon content is generally between 83 and 90 w t % (3). A typical chemical composition of green petcoke as obtained by ultimate analpis showed that the t o t a l sulfur content was 1.29 wt % and the fixed carbon content was 91.8 w t % (9). The volatile matter is composed of heavy hydrocarbons deposited in the coke matrix. It is generally between 9 and 21 wt % and is mainly a function of the coking drum temperature. An increase of the drum temperature vaporizes off the lower-sulfurbearing oils from the coke leaving a more concentrated sulfur-bearing coke (7). Sulfur in Petcoke The sulfur content of the petcoke strongly depends on the nature of the coking feedstock (crude oil) and ita sulfur content. The sulfur content of the feedstocks seems to increase with increasing the concentration of asphaltenes and Conradson carbon content. For instance, higher sulfur contents were found in sponge coke (produced from high-resin asphaltene feedstocks) than in honeycomb coke (produced from low-resin asphaltene feedstocks) or needle coke (produced from highly-aromatic feedstocks). The sulfur content in petcoke varies widely (from less than 0.5% in gilsonite to more than 10%) mainly depending on the sulfur content of feedstock (10). Typically, sponge coke contains between 1% and 6%. Sponge coke containing 4 % is used for fuel whereas that of less than 4% sulfur content is used in anode manufacturing. Needle cokes for electrode manufacturing are required to have lese than 1%sulfur content (11). S u l f u r contents in sponge cokes have been correlated to the sulfur content and Conradson carbon residue of feedstock by Jacob (7). He found that the sulfur content of sponge coke increases almost proportionally to the feedstock sulfur contents and increases less strongly with increasing Conradson carbon residue. Coking temperature also affects the sulfur content of petcoke, though in less degree, mainly due to the vaporization and removal of the low sulfur containing volatile matter which result in a reduction of the total sulfur content in the coke (7). Most of the sulfur in petcoke exists as organic sulfur bound to the carbon matrix of the coke (12). Some sulfur but could also exist as sulfates and as pyritic sulfur (13)) these do not in general make up more than 0.02% of the t o t a l sulfur in coke (3). In at least one case, however, pyritic sulfur was reported to be as high as 0.4% (14). Free sulfur may occasionally be present (13). The structure of organic sulfur compounds in petcoke remains largely unknown, and no precise analytical

* Author to whom correspondence should be addressed.

0000-5005/92/ 2631-1035$03.00/0 0 1992 American Chemical Society

1836 Ind. Eng. Chem. Res., Vol. 31, No. 8,1992

methods exist today to determine exactly this structure (3,15). The organic sulfur compounds identified by the Bureau of Mines in four crude oils include thiols (alkyl, cyclic, and aromatic), sulfides (alkyl, cyclic, and alkyl cycloalkyl), disulfides (alkyl, cyclic, and aromatic), and thiophenes (16). The thiophenes are most prevalent in the heavy fractions of crude oils and hence in coke. Other sulfur compounds are less prevalent in coke. Work done by Sabott (17)indicates that sulfur may not be present in coke in the form of thiols or aromatic and aliphatic sulfides, and sulfides do not make in fact more than 0.003% of the total sulfur in petcoke (3). Sulfur may exist in the coke in many forms (3): (a) as thiophenes attached to the aromatic carbon skeleton; (b) attached to side chains of aromatic or naphthenic molecules; (c) between the aromatic sheets or on the surface of clustered molecules; (d) on the coke surface or in coke pores bound by capillary condensation, adsorption, or chemisorption. Desulfurization of Petcoke The desulfurization of petcoke involves the desorption of the sulfur present in the coke pores or on the coke surface, and the partition and removal of the sulfur attached to the aromatic carbon skeleton. For the removal of the sulfur in the first category a purely thermal treatment at temperatures less than 1100 K is generally sufficient. A more severe treatment and/or the use of chemicals is, however, necessary for the removal of the sulfur attached to the carbon skeleton, particularly in the case of the thiophenic sulfur which is much more stable than the other organic sulfur compounds and therefore much more difficult to remove. This makes it evident why the effective desulfurization of petcoke, involving as it does the rupture of the thiophenes, is not as simple a process as may be desired. The thiophenes do make up most of the sulfur present in the petcoke. They, on the other hand, are much less stable chemically than their aromatic isologs, and it is always possible to find compounds that react more readily with the thiophenes than with the aromatic or other compounds of the coke structure (18), a fact which can be made use of in desulfurization processes. Processes for Petcoke Desulfurization Although much experimental work has been done on the desulfurization of petcoke, there is still as yet no commercial process for desulfurization (19). One common feature of all, or most, published work is its exclusively empirical nature. This is presumably because a theoretical, thermodynamic and kinetic, study is not possible due to lack of essential data on the nature of S-C bonds, their free energy change (AC), and the manner and speed of sulfur replacement. In addition, some important factors such as the coke structure and porosity undergo significant changes during the course of desulfurization, a fact which makes the problem even more complicated (20). 1. Solvent Extraction. Solvent extraction would offer the simplest approach to desulfurization if it were possible to selectively dissolve the organic sulfur compounds present in the coke. As materials of similar chemical structure are more likely to be mutually soluble in one another, aromatic and similar compounds might be used as solvents. Experience with coal indicates that weak organic acids such as phenols and nitrobenzene are more effective than other organic solvents (16).These could be used to dissolve sulfides and disulfides and possibly some thiophenes as well.

Table I. Maximum Desulfurization of Petcoke by Extraction with Different Solvents solvent max desulf, % ref o-chlorophenol 20 14 19 17, 21, 35 pyridine aqua regia 14 17,21 phenol 14 17 furfural 14 17, 35 Chlorex 17 13 naphthalene 13 17 p-cresol 9 35 xylene 9 21, 35 benzene 7 17, 21, 33, 35, 39 nitrobenzene I 35 ethanolamine 6 17,21 toluene 5 17, 21, 35 acetone 2 33 CCl, 1 17, 21 1 17, 21, 33 cs2

Extractions with coke using a large variety of solvents were made in a Soxhlet extractor. No sulfur removal was reported when petroleum ether (21),dioxane (21),or hydrochloric acid (17,211 was used. Extractions made with other solvents led to some sulfur removal, but in no case was more than 20% sulfur removed (Table I). This indicates clearly that solvent extraction is not an effective method of desulfurization. However, the selectivity of solvent extraction would be enhanced if the coke macromolecule could be cleaved. Coke depolymerization can be effected by different methods including mild hydrogenation, oxidation, and prepyrolyais. The effectiveneas of these depolymerization techniques has not been investigated. Phillips and Chao (14) found that increasing the extraction temperature improved desulfurization from 11% at 290 K to 20% at 430 K. This can be explained by the fact that some coke depolymerization must have taken place as a result of the extraction temperature increase. 2. Thermal Desulfurization. By thermal desulfurization is meant the process whereby a fixed static bed of petcoke is heated under atmospheric pressure in an inert atmosphere to a specified temperature and then kept at that temperature for a specified period of time. This process was felt to be the most promising process for the desulfurization of petcoke, and can be the only one possible when other techniques prove to be difficult or inefficient as was found in at least one case with Syrian petcoke (12). Further, petcokes are normally calcined up to a temperature of about 1700 K and desulfurization could be an added asset if it can be shown to take place to a significant degree within this temperature range. The efficiency of desulfurization, however, is not only dependent on the maximum temperature to which the coke is subjected, but other factors affect it also including rate of heating, gas atmosphere, and in particular residence time at the maximum temperature. In order to neutralize the effect of this last factor, a constant value of 30 min residence time was assigned to it throughout and a comparison of the effects of temperature were made on this basis. Table I1 gives a summary of the maximum desulfurization achieved under these conditions at different maximum temperatures. An examination of t h i stable makes it evident that the process of thermal desulfurization can be divided into four stages: (a) The first is an initial phase of desulfurization (300-1100K), with the desorption of sulfur bound on the surface or in the pores, and the simultaneous cracking of side chains of aromatic molecules. The sulfur bound on the surface refers to the S-C bonds in the outer layers or clusters (pre walls, surfaces) that can be physically exposed to the fluid environments. The proceas of sulfur separation starta at 770-820 K, first increasing to reach a maximum

Ind. Eng. Chem. Res., Vol. 31, No. 8,1992 1837


Table 1 1 . Maximum Desulfurization as a Function of Temperature (6,12,13,22,26,27,48,66) temp, K mar desulf, I temp, K max desulf, 5'0 1100 26 1600 32 1300 28 1700 64 1400 30 ls00 91 1500 27 1900 96

at about lo00 K and then decreasing until it stops at 1270 K (22). The maximum amount of sulfur removed in this stage is about 25% (Table 11),which can be taken as a rough indication of the amount of non-thiophenic sulfur in coke. Experimental work indicates that no reaction occurs during this stage between metal Contaminants and sulfur gases, since 'no variation in the degree of desulfurization of different cokes with different amounts of metals was observed (22). (b) A second stage is that in which little or no desulfurization takes place (1100-1600K), particularly in the case of coke made from aromatic-type feedstock. Most of the sulfur removed is derived from the decomposition of the thermally-stable sulfur hydrocarbons bound in side chains. The desulfurization effect is significantly depressed by the metal-hydrocarbon compounds which react with the dissociated sulfur to form refractory sulfur (16). Metal-hydrocarbon compounds refer mainly to vanadium and nickle metals that are trapped in the porphyrinic structures. These metals, certainly, survive coking temperaturea (23). It should be also mentioned that the metals in the feedstock end up in the coke structure. At temperatures greater than or equal to 1500 K desulfurization is further inhibited by the formation of thermally-stable metal sulfides (24). Ash and metal contaminants seem to have no effect on desulfurization up to a temperature of 1500 K (22,24,25). At higher temperatures desulfurization is inhibited by the formation of thermally-stable sulfur compounds (13,22, 24). The inhibiting effect of ash is related to ita amount and composition (26). (c) Desulfurization is dramatically increased when coke is heated above 1600 K. The energy available at 1700 K is sufficiently high for the decomposition of sulfurhydrocarbon compounds of stabilities up to those of thiophene structure (22). Complete elimination of sulfur is not, however, likely even at this high temperature (25, 27). The removal of sulfur is connected with the creation of an organized phase detectable by X-ray methods (20). An organized phase refers to an extended two-dimensional and very short tridimensional arrangements of aromatic layers that resemble graphitic structures which are typical characteristics of needle cokes. (d) Further increase in temperature (>1800K) is not certain to lead to more desulfurization, but this depends also on the nature of coke. The degree of desulfurization is directly related to the total sulfur content of the coke. Gimaev and Syunyaev ( B )on , the other hand, found that for a particular type of coke the degree of desulfurization is affected only by the temperature and is independent of the initial amount of sulfur. The high temperature required for thermal desulfurization can affect some critical coke properties (21)and its effect is much more than the effect of residence time (29). The apparent density increases from 1.2-1.4to 1.6-2.0 g/cm3 on heating to 1600 K (15). Above 1600 K density changes depend on the initial sulfur content, decreasing with cokes of high sulfu content and increasing with cokes of low sulfur content. Real density increases with tem-

perature up to 1700 K, and then it decreases ( 3 , 1 5 , 3 0 ) . The decrease of real density above 1700 K called ''puffing" is a consequence of the porosity developed when the sulfur gets out of the coke matrix at temperatures above 1700 K. In applications above these temperatures, such as arc electric furnaces, this puffing effect can reach levels at which large pores are opened, and stresses are concentrated leading to breakage of graphitic electrodes. For this reason a dynamic puffing test is specified as a part of the characterization of fder cokes for these applications. The use of puffing inhibitors, such as Fe203,generally at 2 pph level can reduce the puffing effect to an acceptable level (23). Coke porosity increases with temperature and so does its reactivity (31). The surface area increases to a maximum at about lo00 K and then it decreases to a minimum at about 1300 K, beyond which there is a further increase in surface area (30). Reis (3) found that the electric resistivity of the petcoke was reduced with increasing temperature up to 1600 K. However, he found that higher temperatures increase the electric resistivity which reaches a maximum at 2300 K (3). Most desulfurization seems to occur within the first 30 min of the thermal treatment up to 1700 K. According to Akhmetov and Syunyaev (32) the highest rate of desulfurization is within the first 5 min, but this does not seem to apply in all cases as can be shown by the results of Brandt and Kapner (19). Increasing the residence time has no effect on the degree of desulfurization except in the range 1600-1800 K (13,19,22,25,27). It is not clear whether the elimination of sulfur is affected in any way by the inert gas used during the thermal treatment (33). Most workers used nitrogen or argon. Carbon-nitrogen complexes are not likely to be formed when petcoke is heated in nitrogen (12). The use of argon may inhibit desulfurization at lower temperatures, but more work is needed for more conclusive results. Thermal desulfurization under vacuum does not seem to be beneficial to desulfurization (20). Decreasing the mean diameter of coke particles may improve the desulfurization efficiency ( I , 32,341, but there is no consensus on this point (13). The use of a fluidized bed may improve the desulfurization efficiency (34). 3. Oxidative Desulfurization. The use of an oxidizing gas during the thermal treatment of petcoke may favor the removal of sulfur. This is related to the resulting weight loss of the coke and its improved porosity. Heating in COz inhibits desulfurization. In the experiments carried out by Hussein et al. (12) little desulfurization was observed below 1500 K (12%), but desulfurization was rapid at greater temperatures (67% at 1800 K). This was accompanied by lower coke yield. When petcoke was heated in oxygen, no sulfur was removed below a certain temperature (lo00K (33) or 1600 K (34)). Air can be used as an oxidizing gas. When coal was heated to 450 "C in air, a maximum of 20% of the organic sulfur was removed (16). No tests with coke are reported. Petcoke can be desulfurized by heating under steam. In the temperature range of 900-1300 K a maxi36) and mum desulfurization of 31% was obtained (35, similar results were reported for coal (6). More sulfur can be removed by raising the temperature to 1600-1800 K. This value reported for petcoke (35, 36) appears to be extremely high when compared with the desulfurization data using oxygen (34). The use of a mixture of air and steam seemed to inhibit desulfurization (35). Oxidizing agents might be used to oxidize organic thiols, sulfides, and disulfides to sulfonic acids which can be hydrolyzed in boiling water to eliminate H2S04. Nitric acid

1838 Ind. Eng. Chem. Res., Vol. 31, No. 8,1992


Table 1 1 1 . Maximum Desulfurization of Petcoke on Treatment with Hydrocarbon Gases gas used temp, K mas desulf, % methane 1140 33 methane 1200 90 butagas 1070 26 effluent gas recirculation 1770 90 butene-2 1370 92 ethylene 1370 93 propylene 1470 94 Table IV. Processes of Effective Desulfurization (>84%) process temp, K desulf, % heating in sulfur-bearing gas 1810 84 heating with effluent gas recirculation 1770 90 heating in steam 1770 87 thermal treatment 1670-1770 89-96 simple HDS 1570 92 treatment with unsaturated hydrocarbon 1370-1470 92-94 gases preoxidation + treatment with methane 1200 90 treatment with Na2C03or K2C03 1170 93 HDS at high pressure 1070 a7 treatment with NaOH 870-1030 98 HDS or coke preoxidized at -610 K 1030 87 HDS with NaOH impregnation 1020 86 treatment with KOH 920 97 HDS in a fluidized bed 870 88

ref
33 19 39 57,58 17 17 17

was used for this purpose by Phillips and Chao (14) and and a maximum desulfurization of Syunyaev et al. (37), 31% was achieved (14). This was accompanied by an increase in oxygen and nitrogen and a decrease in carbon content (37). 4. Desulfurization in an Atmosphere of SulfurBearing Gas. Exposing the coke to H2Sin the temperature range 1000-1200 K increases the sulfur content of the coke. H2Sreacts with the incandescent carbon to form At more stable carbonaulfur compounds (19,35,38,39). elevated temperatures, however, sulfur-bearing gases decompose to form free sulfur or react to form free sulfur and thus induce desulfurization, with the preservation of the critical coke properties. Eighty-four percent sulfur removal was obtained by Brandt and Kapner at 1600 K (19). 5. Treatment in Hydrocarbon Gases. Most laboratory work indicates that the removal of sulfur can be induced at elevated temperatures by fluidizing the hot petcoke with low molecular weight hydrocarbon gases. These decompose to liberate hydrogen, which assists desulfurization, while the carbon is deposited in the coke Table pores increasing thereby the coke density (19,40). I11 gives a summary of maximum desulfurization obtained when different hydrocarbon gases were used. 6. Treatment with Alkali Metal Compounds. Coke may be desulfurized by treatment with an alkali metal compound. More than 98% sulfur can be removed using NaOH (38,41,42). Other compounds may also be used Different techwith varying degrees of success (41,42). niques were suggested in order to improve the desulfurization efficiency including preactivating the coke by or steam (44) or exposing the coke fluidizing with air (43) A major disadvantage sample to microwave radiation (44). of this process is the consumption of large amounts of alkali compounds (38). The presence of the alkali metal compounds in the treated coke is also indicated (38). 7. Hydrodesulfurization (HDS). Hydrodesulfurization is a common process of desulfurization in the oil industry, but it has not yet found application in the treatment of petcoke. The hydrodesulfurization reaction involves a number of steps. For the initial reaction, the diffusion of hydrogen into the pores of the granules appears to be the rate controlling step. As the reaction progresses, the rate-controlling step shifts to the reaction of hydrogen with sulfur compounds (46). A t 900 K it may be inferred that the chemical surface reaction and not the physical process is rate controlling (35).A t higher temperatures (>loo0 K) the reaction of Hfi with coke particles occurs more rapidly leading to the formation of new stable C-S compounds (the reverse reaction). In simple HDS petcoke is heated in a fixed bed under an atmosphere of hydrogen. Experimental work indicates that the rate of sulfur removal is greatest at about 1000 K, where about 50% of the sulfur is removed, but higher temperatures (1600 K) are required for the removal of the more refractory sulfur. At temperatures above 1300 K and below 1500 K sulfur removal is not apparent (20). The maximum desulfurization at the optimum temperature can

be explained by the fact that the net rate of desulfurization is the result of two competing processes, namely the removal of the labile sulfur by hydrogen and the thermal fixation of the sulfur by the reverse reaction which produces a more stable form of sulfur. The reverse reaction becomes more rapid a t temperatures in the range 1000-1200 K. At elevated temperatures (1500-1800 K) the presence of H2Swould induce sulfur removal, a fact which helps to explain the degree of desulfurization obtainable at 1600 K (92%). The efficiency of HDS may be improved by increasing the hydrogen flow rate to an optimum value (39)or by increasing the pressure. The use of fluidized bed desulfurization leads to greater desulfurization at reduced flow rates. This is probably due to the more efficient contact between hydrogen and coke (22). Cooling the coke at different intervals during HDS improved the level of desulfurization, but this was effective 48). only at certain critical temperatures (1000 K) (47, Similarly the effect of heating the coke in an inert atmosphere prior to introducing the hydrogen depended largely on the temperature at which the hydrogen was introduced, with the optimum temperature at about 720 K (49).When the coke is heated in an inert atmosphere, cooled, and then heated in hydrogen, the level of desulfurization is either decreased (33)or greatly enhanced (50) in comparison with simple HDS depending on the pretreatment temperature. Preoxidizing the coke at moderate temperatures may render it more amenable to HDS (511, but the advantage of this treatment is offset by the increased weight loas. The optimum temperature for maximum desulfurization and minimum weight loss was found to be in the range 570-620 K (48,51-53). The presence of steam in HDS can have variable effects depending on the temperature and/or Impregnating the coke with steam flow rate (33,42,45). an aqueous solution of NaOH prior to H2S enhances desulfurization (54). Miyoshi et al. found that alkali desulfurization proceeded smoothly under hydrogen when the coke was exposed to microwave radiation (45).

Conclusions Table IV gives a summary of the most effective desulfurization processes (with sulfur removal greater than 84%). These are grouped into four classes according to the maximum temperature required for effective desulfurization. At temperatures >1700 K, effective desulfurization can be achieved by purely thermal treatment. Steam, sulfur-bearing gases, or effluent gas recirculation can also be used, but with little advantage. In the temperature range 1400-1600 K, effective desulfurization can be achieved by simple HDS or on

Ind. Eng. Chem. Res., Vol. 31, NO.8, 1992 1839 treatment with unsaturated hydrocarbon gases. In the temperature range of 1170-1200 K, effective desulfurization can be achieved either by treatment with some alkali metal compounds (Na2C03, K2C03) or by preoxidation and subsequent treatment with methane. At temperatures lower than 1100 K, two processes are possible: (a) HDS using high pressures, NaOH impregnation, preoxidation, or fluidization; (b) Treatment with NaOH or KOH. Given the fact that the consumption of great amounts of alkali metal compounds is required for effective desulfurization and that preoxidation reduces the coke yield, it becomes evident that HDS in a fluidized bed offers the most attractive desulfurization process. This process is expected to be more effective with cokes of high sulfur content and high ash and heavy metals content, as at temperatures less than 1100 K the adverse effects of high ash and heavy metals content on desulfurization is largely avoided. The process would result in an increased porosity and surface area and improved reactivity. These properties are dependent on the sulfur concentration. The greater the amount of sulfur removed, the greater the structural changes expected in the coke treated. The coke density is expected to increase as a result of the treatment, and there will be some weight loss (<lo%).
(19) Brandt, H. H.; Kapner, R. S. Desulfurization of Petroleum Coke. Light Met. 1984,883-887. (20) Gillot, J.; Lux,B.; Cornuault, P.; Du Chaffaut, F. Changement de Structure Lors De La Desulfuration De Coke De Petrole. Carbon 1968,6,389-395. (21) Mahmoud, B. H.; Ayad, S.; Ezz, S. Y. Desulfurization of Pe455-462. troleum Coke. Fuel 1968,47(6), (22) Vrbanovic, Z . Thermal Desulfurization of Petroleum Coke. 167-176. High Temp.-High Pressures 1981,13, (23) Rodriguez, J. Needle Coke Production From Venezuelan Oil Residues. M.S. Thesis, University of Pittsburgh, 1989. (24) Vrbanovic, Z.Possibility of Using High Temperature Treatment of Petroleum Coke to Desulfurize It. Nafta (Zagreb) 1978,29(2),80-93. (25) Vrbanovic, Z. Thermal Desulfurization of Petroleum Coke. 107-112. High Temp.-High Pressures 1983,15, (26) Syunyaev, Z.I.; Abyzgildin, Y. M. Effect of Ash Components on Petroleum Coke Desulfurization. Khim. Seraorg. Soedin., Soderzh. Neftyakh Nefteprod. 1972,9,440-444. (27) Vrbanovic, Z. Thermal Desulfurization of Petroleum Coke. Bienn. Conf. Carbon 1981,16th,190-191. (28) Gimaev, R. H.; Syunyaev, Z. I. Effect of Temperature on the Equilibrium Sulfur Content of Petroleum Coke. Khim. Seraorg. Soedin., Soderzh. Neftyakh Nefteprod. 1972,9,445-451. (29) Skoczkowski,K.Effect of Temperature and Calcination Time On the Quality of Anthracite and Petroleum Coke. Koks, Smola, Gaz 1990,35 (4),97-99. (30) Barillon, E. Evolution Thermique De La Texture Poreuse Des Cokes De Petrole. Carbon 1967,5, 167-171. (31) Barillon, E. Modification Thermique De La Texture De Cokea De Petrole Lors Dune Desulfuration Thermique. J. Chim. Phys. Phys.-Chim. Biol. 1968,428-432. (32) Akhmetov, M. M.; Syunyaev, Z. I. Thermal Desulfurization of Petroleum Cokes. Neftepererab. Neftekhim. 1969,4, 9-12. (33) Schafer, W. C. Removal of Sulfur From Petroleum Coke by Pyrolysis. Q. Colo. Sch. Mines 1952,47(3),27-37. (34) E l - T a d , S.Z . ;Morsi, B. M. Thermal Treatment of Egyptian Petroleum Coke in an Atmospheric Oxygen. J. Mines, Met. Fuels 1979,27(6),185-191. (35) Parmar, B. S.; Tollefson, E. L. Desulfurization of Oil Sands Coke. Can. J. Chem. Eng. 1977,55, 185-191. (36) Medovikova, N.; Kaftanov, S. V.; Fedoseev, S. D.; Rusinovskaya, N. N.; Razina, G. N. Desulfurization of Petroleum Coke in Superheated Water Vapor and Preparation of Water Gas. Tr. Mosk. Khim.-Technol. Zmt. 1970,66,7-10. (37) Syunyaev, Z.I.; Gimaev, R. N.; Abyzgildin, Y. M.; Malyatova, G. P.; Zaitseva, S. G. Reactions of Organosulfur Compounds in Petroleum Coke. Khim. Seraorg. Soedin., Soderzh. Nef381-386. tyakh Nefteprod. 1968,8, (38) George, Z.M.; Schneider, L. G.; Tollefson, E. L. Desulfurization of A Fluid Coke Similar to the Athabasca Oil Sands Coke. Fuel 1978,57,497-501. (39) George, Z .M. Hydrodesulfurization of Coke From Athabasca Tar Sands Operation. Znd. Eng. Chem. Prod. Res. Dev. 1975, 14,298-303. (40) El-Ansary, A. F.; Em,S. Y. Desulfurization of Petroleum Coke by Butagas. Fuel 1973,52, 66-70. (41) George, Z.M.; Parmar, B. S.; Tollefson, E. L. Desulfurization of High Sulfur Cokes From Processing of Oil Sands Bitumen. Pac. Chem. Eng. Congr. [Proc.] 1977,2,1299-1305. (42) Lukaaiewicz, S.J.; Johnson, G. C. Desulfurizationof Petroleum Coke. Znd. Eng. Chem. 1960,52,675-677. (43) Murphy, R. M.; Conn, D.; Meesman, H. C. Process for Desulfurizing Carbonaceous Materials. US. Patent 3,387,941, June 11, 1968. (44) Messman, H. Petroleum Cokes Future: Bright or Black? Oil Gas. J. 1969,Sept 22,142-144. (45) Miyoshi, N.; Tsutaumi, S.; Sonoda, N. Desulfurization of Coke Using Sodium Hydroxide Under Irradiation of Microwave. Tech. Rep. Osaka Univ. 1975,25 (1230-1253),241-247. (46) Takanari, N.; Tanaka, T. Desulfurization of Petroleum Coke with Hydrogen. Sekiyu Gakkai Shi 1973, 16 (5),415-420. (47) Mochida, I.; Furuno,T.; Fujitsu, H.; Oyama, T.; Fujimoto, K. Hydrodesulfurization of Needle Coke in Atmospheric Hydrogen Flow. Fuel 1988,67,678-681. (48) Mochida, I.; Marutsuka, T.; Furuno, T.; Korai, Y.; Fujitsu, H. Hydrodesulfurzation of Petroleum Cokes. High Temp.-High Pressures 1987,19,545-553. (49) Sef, F. Desulfurization of Petroleum Coke. Znd. Eng. Chem. 1960,52 (7),599-600.

Acknowledgment

B.I.M. expresses his sincere appreciations to Mr. J. Rodriguez for his technicalassistance in revising this paper.
Registry No. S, 7704-34-9.

References
(1) Kobe, McKetta, Eds. Advances in Petroleum Chemistry and . Refining; Interscience: New York, 1959; Vol. 2 (2) Deviney, OGrady, Eds. Petroleum Derived Carbons; American Chemical Society: Washington, DC, 1976. (3) Reis, T. To Coke, Desulfurize and Calcine, Part 2 : Coke Quality and Its Control. Hydrocarbon Process. 1975,June, 97-104. (4) Meyers, R. A. Handbook of Petroleum Refining Processes; McGraw-Hill: New York, 1986. (5) Figueiredo, Mouljin, Eds. Carbon and Coal Gasification; Martinus Nijhoff: Dordrecht, The Netherlands, 1986. (6) NPRA Q & A Session on Refining and Petrochemical Technology; The Petroleum Publishing Co.: Tulsa, OK, 1976. (7) Jacob, R. R. Coke Quality and How to Make It. Hydrocarbon Process. 1971,Sept, 132-136. (8) Marsh, M.; Calvert, C.; Bacha, J. Structure and Formation of Shot Coke, a Macroscopic Study. J. Mater. Sci. 1985,20, 289-302. (9) Nelson, H. W. Petroleum Coke Handling Problems. Znd. Eng. Chem. Prod. Res. Dev. 1970,9,176-178. (10) Petroleum Coke: Process Economics Program; Report No. 72-a; Stanford Research Institute: Menlo Park, CA, October 1976. (11) Feintuch, H. M.; Bonilla, J. A.; Godino, R. L. Delayed-coking Process. Handbook of Petroleum Processes; Meyers, R. A., Ed.; McGraw-Hill: New York, 1986; Chapter 7, pp 43-49. (12) Hussein, M. K.; El-Tawil, S. Z.; Rabah, M. A. Desulfurization : In a Nitrogen Atmosphere. of High Sulfur Egyptian Coke. 1 J. Zmt. Fuel 1976,Sept, 139-143. (13) El-Kaddah, N.; Ezz, S. Y. Thermal Desulfurization of Ultra High Sulfur Petroleum Coke. Fuel 1973,52,128-129. (14) Phillips, C. R.; Chao, K. S. Desulfurization of Athabasca Petroleum Coke by (a) Chemical Oxidation and (b) Solvent Extraction. Fuel 1977,56,70-72. (15) Reis, T. About Coke and Where the Sulfur Went. Chemtech 1977,June, 366-373. (16) Meyers, R. A. Coal Desulfurization; Marcel Dekker: New York, 1977. (17) Sabott, F. K. A Study of Methods of Removing Sulfur From Petroleum Coke. Q. Colo. Sch. Mines 1952,47 (3),1-22. (18) Hartough, H. D. Thiophene and Its Derivatives; Interscience Publishers: New York, 1952.

1840

Ind. Eng. Chem. Res. 1992, 31, 1840-1848


(55) Lebiedziejewski, M.; Szudek, M.; Harmanowska, J.; ZmudzinKalinowski, B.; W l o d m k i ,R. Desulfurizing ski, B.; Hulisz, S.; Raw Petroleum Coke. Pol. 57,397 (C1.C lob), June 30,1969. (56) Kalinowski, B.; Wlodarski, R. Desulfurizing of Petroleum Coke. Pol. 51, 317 (C1.C lob), June 20, 1966. (57) Manzanilla, F.; Moreno, 0. New Process Desulfurizes Coke. Hydrocarbon Process. 1979, March, 97-102. (58) Manzanilla, F.; Moreno, 0. Thermal Process Is Developed for Petroleum Coke Desulfurization. Oil Gas J. 1979, Jan 22, 64-68.

(50) Mochida, I.; Marutsuka, T.; Korai, Y.; Fujitsu, H. Hydrodesulfurization of Petroleum Coke Deposited on Iron Ores. Znd. Eng. Chem. Prod. Res. Dev. 1986,25, 30-33. (51) . . Mason, R. B. Hvdrosulfurization of Coke. Znd. Eng. Chem. 1959,51 (9), 1Oi7-1030. (52) Mochida, I.; Marutsuka, T.; Korai, Y.; Fujitsu, H. Enhanced Hvdrodesulfurization of Coke Deposited on Iron Ore by Air Gasification. Fuel 1987,66, 70-73. (53) Hsu, H. L . ;Hardin, E. E.; Grindstaff, L. I. Calcining and Desulfurizing Petroleum Coke. Br. Patent 2,093,061, Aug 25, 1982. (54) George, Z.M.; Schneider, L. G. Sodium Hydroxide-Assisted Desulfurization of Petroleum Fluid Coke. Fuel 1982, 61, 1260-1266.

Received for review February 25,1991 Revised manuscript received March 13, 1992 Accepted May 5, 1992

KINETICS AND CATALYSIS


Application of the Response Surface Methodology to the Kinetic Study of the Gas-Phase Addition of Ethanol to Isobutene on a Sulfonated Styrene-Divinylbenzene Resin
Montserrat Iborra,* Jose F. Izquierdo, Fidel Cunill, and Javier Tejero
Chemical Engineering Department, Faculty of Chemistry, University of Barcelona, Mart; i FranquBs, 1, 08028 Barcelona, Spain

The kinetics of vapor-phase addition of ethanol to isobutene to give ethyl tert-butyl ether (ETBE) on the ion exchange resin Amberlyst 15 has been studied. Rate data were obtained in a continuous differential reactor operated a t atmospheric pressure and 53-78.5 "C. The most probable Langmuir-Hinshelwood-Hougen-Watson (LHHW) rate models stem from two quite similar reaction mechanisms. Four active centers take part in the rate-determining step of the two mechanisms. The first one suggests the reaction between the ethanol and the isobutene, both adsorbed molecularly on one center, and the second mechanism proposes the reaction between the ethanol adsorbed molecularly on one center and the isobutene adsorbed on two centers without cleavage. Both mechanisms are physicochemically coherent.

Introduction
To find the most suitable function to correlate rate data and independent variables is the main problem in kinetics. The reaction rate equation must be detailed enough to describe minor influences, but it should not be too complex to be useful in reactor design. Empirically developed models, despite their simplicity, are of limited value in extrapolation. Therefore, theoretically founded models are more desirable. The response surface procedures appear to be useful for finding empirical models for a given reaction. They allow an adequate fit, giving an empirical rate equation in the form of a power series sometimes useful for preliminary designs, and an indication of possible mechanistic rate expressions, as shown by Pinchbeck (1957)in the vaporphase oxidation of naphthalene to phthalic anhydride, Kittrell and Erjavec (1968)in the re-examination of vapor-phase isomerization of n-pentane to 2-methylbutane data obtained by Carr (19601,Pirard and Kalitventzeff (1978) in the hydrogenation of ethylene on coppermagnesia catalysts, and Tejero et al. (1989)in the vaporphase addition of methanol to isobutene on Amberlyst 15. The response surface methodology used as a preliminary approach to the identification of the kinetic equation is more rational than the direct fit of nonlinear equations

because it introduces into the regression only the significant variables on a statistical basis. This is the basis of the Hofmann method (Hofmann, 1972)for determining the best kinetic equation for reactions from which little or nothing is known. The method consists of the following stages. (a) With factorial, simplex, or central composite designed experiments as the starting point, the response surface of the reaction rate is explored by linear regression of the data and transformation of the regression equation into its canonical form to visualize it. At this stage significant variables can be selected. (b) The information thus obtained and any additional information about the type of reaction should be used to establish some mechanietic reaction schemes, leading to several possible model equations. Selection between these rival models is possible using the classical discrimination techniques. (c) For the most suitable model equation, the parameter values should be determined by nonlinear regression. Now then, when experimental setups have a low experimental error, a quadratic function does not often fit the rate surface in a statistically significant fashion and so a higher order function should be fitted (Himmelblau, 1970). As a result, interpretation of the influence of the

0888-5885/92/2631-1840$03.00/0 0 1992 American Chemical Society

You might also like