Вы находитесь на странице: 1из 18

Microporous and Mesoporous Materials 29 (1999) 4966

Review

Methanol-to-hydrocarbons: process technology


Frerich J. Keil *
Technical University of HamburgHarburg, Department of Chemical Engineering, Eissendorfer Str. 38, D-21073 Hamburg, Germany Received 4 February 1998; received in revised form 30 June 1998; accepted 15 July 1998

Abstract This review presents methanol-to-hydrocarbons processes which have reached industrial applications, either on a commercial or on a pilot plant scale. The determination of kinetic expressions for various methanol conversion reactions is given. The processes discussed are: Mobils methanol-to-gasoline (MTG) plant in New Zealand, the uidized bed MTG and methanol-to-olens process; Mobils olen-to-gasoline/distillate (MOGD) process; the MTO plant developed by UOP and Norsk Hydro; Haldor Topses TIGAS process. The developments of a liquid phase dimethyl ether synthesis (LP-DME) process by the Ahron University are also presented. 1999 Elsevier Science B.V. All rights reserved. Keywords: Commercial plants; Kinetics; Methanol-to-gasoline; Methanol-to-hydrocarbons; Methanol-to-olens; Review

1. Introduction Mobils novel synthetic gasoline process, based on the conversion of methanol to hydrocarbons over zeolite catalysts, was the rst major new synfuel development in the 50 years since the development of the FischerTropsch process. This process is known as the methanol-to-gasoline (MTG) process. It provided a new route from coal or natural gas to high-octane gasoline. According to Chang and Silvestri [1], two teams of Mobil scientists working on unrelated projects discovered by accident the formation of hydrocarbons from methanol over the synthetic zeolite ZSM-5 [2,3]. The group at Mobil Chemical in Edison, New Jersey, had been trying to convert methanol to ethylene oxide, while workers at
* E-mail address: keil@tu-harburg.de (F.J. Keil )

Mobil Oils Central Research Laboratory in Princeton were attempting to methylate isobutene with methanol in the presence of ZSM-5. Neither reaction proceeded according to expectation. Instead, aromatic hydrocarbons were found. Mobils Central Research team tried to nd out whether methanol could serve as a precursor to a C olen in alkylating isobutane, to form, pre1 sumably, neopentane. ZSM-5 was the rst catalyst tried for this hypothetical reaction. An equimolar mixture of methanol and isobutane was passed over HZSM-5. Methanol was quantitatively converted, whereas only about 27% of isobutane reacted. An experiment carried out with pure methanol also showed a complete conversion of methanol. A careful material balance revealed that the overall reaction stoichiometry could be represented as CH OH [CH ]+H O 3 2 2 (1)

1387-1811/99/$ see front matter 1999 Elsevier Science B.V. All rights reserved. PII: S1 3 8 7 -1 8 1 1 ( 9 8 ) 0 0 32 0 - 5

50

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

where [CH ] is the average composition of the 2 hydrocarbon product. More detailed investigations suggested that the main reaction steps are: H2O H2O 2CH OH P CH OCH light olefins 3 3 3 +H2O higher olefins+n/isoparaffins +aromatics+naphthenes As can been seen from the reaction scheme, methanol is rst dehydrated to dimethylether (DME ). The equilibrium mixture of methanol, DME and water is then converted to light olens (C C ). A nal reaction step leads to a mixture 2 4 of higher olens, n/iso-parans, aromatics and naphthenes. An interruption of the reaction leads to a production of light olens instead of gasoline. An appropriate process for this purpose was developed by Mobil, the so-called methanol-to-olen process (MTO). Therefore, the discovery of the MTG reaction was an accident. This discovery gave occasion to a tremendous amount of detailed investigations of the reaction mechanisms and optimization of the catalysts. Furthermore, new types of zeolites for the MTO and MTG reaction were synthesized. A review of these investigations is presented by Stocker [4]. The oil crisis 1973 and the second oil crisis in 1978 initiated the development of a commercial MTG process. In response to a request from the New Zealand Government, Mobil Research and Development Corporation built a 640 l per day (four barrels per day), xed-bed pilot plant to demonstrate the feasibility of the gas-to-gasoline (GTG) process. Reports to New Zealands Liquid Fuels Trust Board (LFTB) by Lurgi and Badger indicated condence that the process would scale successfully from the pilot plant to commercial size. The project concepts were developed under the terms of a 1980 Government Mobil Memorandum of Understanding. Mobil was responsible for the overall project management, Bechtel acted as Project Services Contractor. Davy McKee, Foster Wheeler and New Zealand Engineering Consultants contributed to the design and engineering of the project. In a rst step, methanol is synthetized from natural gas (Maui

(2)

eld ). The ICI low-pressure methanol process was employed. Two trains, each capable of producing 2200 tons per day were installed. The methanol product is passed to the MTG plant. In 1986 the startup phase of the project was completed with full commercial production with a capacity of 570 000 tons of gasoline per year being achieved. The nal gasoline produced does not need further rening and attains the quality of unleaded premium gasoline. Owing to some reasons presented below, from 1981 and 1984, Mobil, Union Rheinische Braunkohlen Kraftsto AG (RBK ) and Uhde (Dortmund, Germany) have operated a demonstration plant for a Fluid Bed Mobil Process. The plant was located at the RBK facilities in Wesseling (Germany), and was operated from December 1982 to the end of 1985. This plant has successfully demonstrated the performance of the uid bed reaction system for MTG and MTO technology [5,6 ]. Union Carbide also developed a process to convert methanol to olens in 1986 using a silicoaluminophosphate (SAPO) catalyst. The olens yield exceeded 90%, and they report that the process could be modied for high ethylene and propylene yield (about 60%) [710]. As the oil price dropped again over the 1980s further developments of commercial processes were stopped for the time being. Nevertheless, investigations on a bench scale were pursued, and applications for patents are still submitted. As Stocker has reviewed the details of catalyst and reaction mechanisms of the methanol-tohydrocarbons (MTHC ) processes, especially the MTG, MTO and MOGD (Mobils olen-to-gasoline and distillate) processes, the present paper will focus on the technology of the MTG, MTO and MOGD processes. Some new developments, such as direct conversion of methane to fuels will be discussed. First, some kinetic equations will be presented. Second, the technology of the respective processes will be discussed.

2. Kinetics For the design of chemical reactors the kinetic expressions must be known. They should enable the designer to simulate various reactor types and

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

51

various modes of operation. The kinetic models can be grouped into two main classes: (a) lumped models which are a compromise between simplicity and representation of the reality of the process; and (b) detailed models that take into account individual reaction steps. In general, it is very time consuming or even nearly impossible to nd the kinetic expressions of type (b). For design purposes type (a) kinetics will do in most cases. These kinetic expressions have to cover the whole range of reactor operation with respect to temperatures, pressures and feed compositions. It was found in the 1970s that over a wide range of conversions the initial step of ether formation is much more rapid than the subsequent olen-forming step, and is essentially at equilibrium [11,12]. Thus the equilibrium oxygenate mixture can be conveniently treated as a single kinetic species. Based on these facts, Voltz and Wise [13] have developed a lumped kinetic model which described the rate of methanolDME disappearance in process and pilot plant studies of methanol conversion to gasoline. Chang and Silvestri [14] have postulated a mechanism of hydrocarbon formation from methanolDME. They have supposed a concerted bimolecular process involving carbonoid intermediates. The intermediates then undergo sp3 insertion into CH bonds, forming higher alcohols or ethers, which can dehydrate to form olens and which can add :CH to form higher olens. Chen 2 and Reagan [15] discovered that the oxygenate disappearance is autocatalytic over ZSM-5. They proposed the following scheme: k A 1 B k A+B 2 B k B 3 C (3) (4) (5)

Autocatalysis was supported by measurements executed by Chang et al. [16 ] and Ono et al. [17,18]. Chang [19] extended the model using the following assumptions: (1) Methanol and DME are always at equilibrium and can be treated as a single kinetic species. (2) Generation of the reactive intermediate is rst order in oxygenates. (3) Consumption of the reactive intermediate is rst order in oxygenates. (4) Olens can be treated as a single kinetic species. (5) Disappearance of olens is rst order in olens. These assumptions lead to the following set of reactions: k A 1 B k A+B 2 C k B+C 3 C k C 4 D (7) (8) (9) (10)

where A represents the oxygenates, B :CH groups, 2 C olens, and D parans+aromatics. The formal kinetics are: dA/dt=k A+k AB (11) 1 2 dB/dt=k Ak ABk BC (12) 1 2 3 dC/dt=k AB+k BCk C (13) 2 3 4 Assuming the steady-state condition for B and eliminating time resulted in an expression like this: du/dA=1/A[(1+K u)/(2+K u)(1K u)+u] 1 1 2 (14) where u=C/A, K =k /k and K =k /k . 1 3 2 2 4 1 Anthony [20] improved Changs kinetic expression. Doelle et al. [21] studied both sorption and reaction kinetics of methanol and DME conversion over ZSM-5. In the range between 115 and 200C the kinetics of methanol conversion followed the

where A represents oxygenates, B olens, and C aromatics+parans. The rate of oxygenate disappearance is rst order in oxygenates [13]. The experimental data could be tted according to the expression dA/dt=k A+k AB 1 2 (6)

52

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

rate law r=k P /(1+k P ) (15) 1 CH3O 2 H 2O Schipper and Krambeck [22] have obtained results in a pilot plant with an adiabatic xed-bed reactor, which was operated under reactionregeneration cyles and under conditions in which there was irreversible deactivation. They dened catalyst activity for a given time on stream, b, as the product of two activity terms, which correspond to the remaining activity due to permanent deactivation, a, and to the remaining activity due to reversible coking deactivation, f: b=a f (16)

catalyst. The following two lumped models were able to treat the experimental data consistently: Model I: k A 1 B k A+B 2 B k A+C 3 D and k B+E 4 F k C+E 4 F k D+E 4 F Model II: 3k A 1 B+C+D k B+E 2 F k C+E 3 F k D+E 4 F (27) (28) (29) (30) (24) (25) (26) (21) (22) (23)

The reaction rate of each individual step, r , is i the product of the reaction rate for the fresh catalyst r and the activity b: io r =r b (17) i io For the irreversible deactivation kinetics Schipper and Krambeck [22] proposed an empirical equation similar to that used for permanent deactivation in catalytic cracking: d /dt=K exp(E /RT ) am; m>1 (18) a a0 a The loss of activity due to coke deposition is expressed as: df/dt=K bn; n>1 (19) cok By combination of the previous equations, a total deactivation rate is obtained db/dt=a{K bnKm2 b} (20) cok a This equation indicates that the rate of change of total catalyst activity is related not only to the total activity of the catalyst, b, but also to the clean-burned activity, a. When the catalyst is fresh, a is quite high so that the rate of activity loss is high. After the rst reactionregeneration cycle, a is much lower. Thus, the rate of overall aging is lower on the second cycle. Sedran et al. [23] tested the kinetic model by Chang [19] including the modications proposed by Anthony [20] over the 302370C temperature range. In a further paper Sedran et al. [24] compared dierent lumped kinetic models for methanol conversion with hydrocarbons on a ZSM-5

where AMethanol+DME, Bethene, C propene, Dbutene, Fparans. The authors found that an exponential activity function of the type k =k exp(b Hc/W ) (31) i io i can be applied to the observed values, where Hc/W refers to the cumulative average amount of hydrocarbons produced per unit mass of catalyst that was generated by the catalyst corresponding to each run. Therefore, the total hydrocarbon formation is responsible for catalyst deactivation by coking.

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

53

Benito et al. [25] considered the eect of composition on the deactivation. This model has been proven to be suitable by means of a wide experimental study carried out in an isothermal xedbed integral reactor. ZSM-5 with a Si/Al ratio of 24 was employed. The temperature range was 300375C. The best tting model was the one proposed by Schipper and Krambeck [22]: k MEOH/DME(A) 1 light olefins (C ) k 2C 2 products (D) k A+D 3 D k C+D 4 D (32)

where a is the activity dened as a=r (t)/r (t=0) i io (43)

The k are the kinetic constants for deactivation di by coke formation for lump i. The experimental results revealed that the following equation can describe the change of activity with time: da/dt=(k X +k X +k X )a dA A dC C dD D (44)

(33)

(34)

(35)

The light olens (ethylene, propylene) can polymerize to form products in the gasoline boiling range D. Benito et al. [25] found the following kinetic constants: k =0.7331013 exp(33358/RT ) 1 k =0.127108 exp(17633/RT ) 2 k =0.2041012 exp(27987/RT ) 3 k =0.634106 exp(15855/RT ) 4 The kinetic equations are: r =k X k X X =dX /dt A0 1 A 3 A D A r =k X k X2 k X X =dX /dt C0 1 A 2 C 4 C D C (40) (41) (36) (37) (38) (39)

The X represent weight fractions of lump i (on a i water-free basis), t is the space time. The authors proposed a deactivation kinetic model of separable functions dependent on the concentration of the three lumps of the reaction scheme that are possible coke precursors: da/dt=[ (k X )]ad di i (42)

The k were found to be dierent. In a further di paper Gayubo et al. [26 ] state that the models of Chen and Reagan [15] and Schipper and Krambeck [22] adequately t the experimental results. The authors emphasize the following aspects: (1) ethylene and propylene are identied as primary products present in the gas phase; (2) an oligomerization step of light olens is established; and (3) a methylation step of products is established. Although the kinetic model of Schipper and Krambeck [22] is slightly more complex than the one of Chen and Reagan [15], it is more suitable because its kinetic scheme is closer to the real mechanism of the MTG process. Bos et al. [27] developed a kinetic model for the methanol-to-olens process [MTO], based on a SAPO-34 catalyst. This catalyst makes ethene as a main product. The model is based on dedicated experiments in a pulse-ow, xed-bed reactor. The kinetic model was implemented in mathematical models of various reactors for the estimation of product selectivities and main dimensions. The experiments showed that the MTO reactions on a fresh catalyst are very fast, with an overall rstorder rate constant of roughly 250 m3 m3 s1. gas cat The coke content of the catalyst is the main factor governing the selectivity and activity of the catalyst. In order to achieve an ethene-to-propene ratio of 1 or higher, at least 78 wt.% of coke must be present on the catalyst. Consecutive reactions cannot be neglected. The net eects of these are an increase of ethene and propane selectivity and a decrease of mainly propene selectivity. Bos et al. [27] have tested several kinetic schemes. The nal kinetic network of 10 rst-order and two second-order reactions describes the experimental results satisfactory. The nal reaction

54

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

scheme for following:

model

discrimination

was

the

(10) The aromatics undergo condensation. (11) The aromatics undergo alkylation with methanol. (12) The parans undergo demethanization forming olens and methane. The authors found a satisfactory agreement between the experimental and calculated results. A further detailed model was developed by Iordache et al. [29].

3. Fixed-bed methanol-to-gasoline (MTG) process Ten years after Mobil announced a process [1,30] for converting methanol to high-octane gasoline from non-petroleum sources, a commercial plant was in operation in New Zealand. The plant converts natural gas from the Maui and Kapuni elds into methanol and then into ca. 700 000 tons per day of gasoline via Mobils xed-bed MTG process. The gasoline produced is fully compatible with conventional gasoline. The conversion of methanol to hydrocarbons and water is virtually complete and essentially stoichiometric. The reaction is exothermic with a heat of reaction of about 1.74 MJ kg methanol1. The adiabatic temperature rise is about 600C. A simplied block diagram of the MTG process is given in Fig. 1. Methanol is vaporized and fed into the xedbed DME reactor. In the DME reactor, the methanol is catalytically equilibrated to a mixture of dimethyl ether, methanol and water. The reactor contains a special alumina catalyst. The reaction takes place at a reactor inlet temperature of 310320C and about 26 bar pressure. Approximately 1520% of the heat of reaction is released in this rst step, which is controlled by chemical equilibrium. The DME reactor euent is mixed with recycle gas (see later) to moderate the temperature rise over the second reactor, and then fed into the ZSM-5 reactor. The inlet temperature range of this reactor is 350370C. About 85% of the reaction heat is released in the conversion reactor. After cooling, the euent is separated into three phases: gas, liquid water, and liquid hydrocarbons. Most of the gas is recycled to the ZSM-5 reactor. The water contains a small amount of oxygenates which are treated in a biological waste

(45) Eqs. (8) and (12) are of second order. The formation of ethene from propene is of rst order in methanol and propene. The rate of formation of ethene from butene depends on butene and methanol. Besides the lumped models, a few far more detailed kinetic models were developed. Mihail et al. [28] include 53 reactions which were grouped into 12 subgroups. These 12 major steps are: (1) The etherication reaction takes place concurrently with the thermal decomposition of the methanol into hydrogen and carbon monoxide. The ether generates the carbene. (2) The carbene attacks the ether and the alcohol, forming light olens. (3) The carbene attacks the olens, forming higher olens. (4) The carbene attacks the hydrogen, forming methane. (5) The light olens generate carbeniums ions. (6) The carbenium ions attack the light olens forming higher olens (oligomerization). (7) The carbenium ions attack the higher olens forming parans and dienes. (8) The carbenium ions attack the dienes forming parans and cyclodienes. (9) The carbenium ions attack the cyclodienes forming parans and aromatics.

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

55

Fig. 1. Block diagram of the xed-bed MTG process [31].

water treatment plant. The hydrocarbon product is distilled. It contains mainly raw gasoline, dissolved hydrogen, carbon dioxide and light hydrocarbons (C C ). The non-hydrocarbons, C C 1 4 1 3 and a part of the C hydrocarbons are removed 4 by distillation to produce gasoline. The raw MTG gasoline contains considerable amounts of durene (1,2,4,5-tetramethylbenzene). The heavy gasoline treating unit (HGT ) removes most of the durene which causes driveability problems, since the freezing point of durene is relatively high (79C ). The treated heavy gasoline is blended with other gasoline components to give specication nished gasoline. The development of the MTG process has been described in the literature, see for example Refs. [3133]. Bench-scale studies of the MTG process were executed for xed-bed and uid-bed reactors. In Fig. 2 a two reactor conguration is shown [31]. The xed-bed reactor conguration is quite simple and requires minimum scale-up studies. With fresh catalysts, the reaction occurs over a relatively small zone. The reaction front moves down the catalyst bed as the coke deposits rst deactivate the front part of the bed. Use of a sucient catalyst volume permits a xed-bed design in which on-stream periods are long enough to avoid frequent regeneration cycles. A synthetic crude methanol blend containing 83 wt.% methanol (commercial, pure) and 17 wt.% distilled water was used as the charge. This is a typical water content of crude methanol made from natural gas.

In an adiabatic reactor an S-shaped temperature prole along the reactor is formed. There is a very small change in the shape of the prole as the catalyst ages. This means that coke formation downstream of the main reaction zone is low or its level does not appear to aect catalyst activity. After some time, the reaction zone approaches the reactor outlet and signicant quantities of methanol appear in the water product. After regeneration, the catalyst can be returned to conversion service. Owing to band aging and catalyst deactivation, methanol is processed at a continuously higher eective space velocity as the cycle progresses. Gasoline yields are greatest in the vicinity of methanol breakthrough. Catalyst aging leads to a change in hydrocarbon products. As aging progresses, production of normal parans decreases and the isoparan content increases. The aromatics content decreases, and the content of olens and naphthenes increases. The isoparans compensate for the aromatics, so that the gasoline octane number varies only little with time. One should keep in mind that the xed-bed process results in a slightly changing product composition. Under MTG reaction conditions, the ZSM-5 catalyst undergoes two types of aging: a reversible loss results from coke formed on the catalyst as a reaction by-product and the reaction product stream also causes a gradual loss of activity. High temperature enhances this type of deactivation. As dierent segments of the catalyst bed are subjected to varying degrees of water partial pressure and

56

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

Fig. 2. Schematic of xed-bed bench-scale plant [31].

temperature, a permanent activity gradient results in the catalyst bed. In an 8-month aging test in the bench-scale plant under realistic process conditions the following results were obtained [31]: (1) Start-of-cycle (SOC ) gasoline yields increase from cycle to cycle as a consequence of permanent aging. (2) End-of-cycle ( EOC ) gasoline yields are fairly constant. (3) As the catalyst ages, the change in gasoline yield within a cycle decreases, and the cycle average gasoline yield increases. The cycle lengths stabilize. (4) The propane/propene ratio can be used to track catalyst activity. (5) SOC propane/propene yields show a sharp decline over the rst 50 days of operation followed by a gradual decline, whereas the EOC values appear to approach a constant value. (6) Inherent gasoline selectivity did not change throughout the aging test. (7) The selectivity of the ZSM-5 catalyst for the desired gasoline product increases as it ages. This is a peculiarity of the MTG process. (8) After the bench-scale tests a demonstration

unit with a capacity of 0.636 m3 day1 of methanol was built and operated. The major objective of the demonstration plant was to verify the bench-scale results. The only dierent variable between a bench-scale unit and a commercial-size reactor is the linear velocity of the reactants. The catalyst bed diameter was 5 cm, the bed length was 3 m. The corresponding values for the ZSM-5 reactors were 10 cm and 2.4 m. The linear velocities in these beds were about 10 times those in the bench-scale unit. Heat transfer along the reactor wall is negligible. The product yields, selectivities, adiabatic temperature rise and the band aging behavior were nearly the same compared with the bench-scale results. The cycle lengths were about 50% longer. This eect can be related to the slower rate of movement of the catalyst bed temperature prole. As the raw MTG gasoline contains about 5.5 wt.% durene, a heavy MTG gasoline treatment ( HGT ) was developed [34,35]. In the HGT process a 177+C cut of MTG gasoline, comprising primarily aromatics, is processed over a multifunctional metalacid catalyst. The following reactions occur: disproportionation, isomerization, transalkylation, ring saturation, and dealkylation/cracking. The durene

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

57

content is reduced to less than 2 wt.%. The gasoline produced in the demonstration plant was used in an automobile test eet. These tests included investigations with New Zealand-type cars, US cars, Japanese and European cars. The tests were conducted under a wide range of ambient conditions. The performance of MTG gasoline was equivalent to that of conventional gasolines of similar volatility. After the successful MTG process development a Joint Executive Committee (JEC ), installed on an agreement between the New Zealand Government and Mobil in 1980, was to prepare a report which included a plan for the design, construction, and operation of a plant to manufacture gasoline, and an assessment of the viability of such a project. Mobil was responsible for overall project management, Bechtel Petroleum Inc. was employed as Project Services Contractor. The JEC report was completed in July 1981. It concluded that the project was technologically feasible and commercially attractive. The synfuel plant was mainly commissioned during 1985. In November 1985 the rst MTG gasoline was sent to the Ministry of Energy tank farm near Port Taranaki. The New Zealand MTG plant is sited within 180 hectares of land at Motunui, Taranaki. It is designed to convert 5255 PJ per annum of natural gas into 570 000 tons per year of gasoline. The details of the commissioning of the MTG plant were outlined by Maiden [36 ], Bem [37], and Chang [38]. Some aspects of plant design and scale-up considerations were presented by Krohn and Melconian [39]. A simplied block diagram of the New Zealand plant is given in Fig. 3. The plant consists of three main process units: two methanol trains, each capable of producing 2200 tons per day, and the MTG conversion plant. The gas-to-methanol plant employs the ICI lowpressure methanol process. Details concerning the methanol plant are described by Allum and Williams [40]. The MTG unit is based on the Mobil ZSM-5 catalyst. A more detailed presentation of this plant is given in Fig. 4. Depressurized crude methanol is pumped to reaction pressure and vaporized against ZSM-5 reactor euent before owing into the rst-stage dehydration (DME ) reactor. In this reactor crude

methanol is partially dehydrated over a special alumina catalyst to an equilibrium mixture of methanol, dimethyl ether and water. This reaction takes place at a reactor inlet temperature of 310320C and 27 bar and releases 1520% of the overall heat of reaction. The DME reactor euent is split into four parallel streams, mixed with heated recycled gas and passed into four parallel conversion reactors. The recycle stream controls the temperature rise. These reactors contain a ZSM-5 catalyst, where the conversion to hydrocarbons and water is completed. This type of reactor is presented in more detail in Fig. 5. The conversion reactor inlet temperatures are controlled to be 350366C. The inlet pressures are 1923 bar. In this reactor the main part of reaction heat is released. Five conversion reactors are installed of which only four are on-stream. The fth reactor is thus either in the regeneration mode or on standby. The hot reactor euent is rst cooled by generating steam, then fresh crude methanol and recycle gas is heated. The reactor euent is then further cooled to 2535C at 16 bar in a bank of water coolers and enters a threephase product separator, where gas, liquid hydrocarbons and water separate. The water phase, which contains trace amounts of oxygenated organic compounds is passed to the water treatment. The gas phase contains mostly light hydrocarbons, hydrogen, CO and CO . This gas is 2 recycled with the aid of a recycle compressor to the conversion reactors. The raw gasoline which contains dissolved hydrogen, carbon dioxide and light hydrocarbons (C C ) is sent to the 1 4 de-ethanizer. The o-gases, including methane, ethane and some C , together with a purge gas 3 stream from the product separator are scrubbed in a sponge absorber in order to retain any gasoline components. Then the gas is sent to the fuel gas system. The de-ethanizer bottom product is sent to the stabilizer where C and part of the C 3 4 components are removed overhead to the fuel gas system. C /C components are withdrawn as a 4 5 sidestream. The bottom product is fed into a gasoline splitter where it is separated into light and heavy gasoline fractions. Each stream is cooled and stored. As can be seen from Fig. 3, the heavy gasoline fraction, which contains durene, is passed

58

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

Fig. 3. Simplied block ow diagram of the New Zealand GTG plant [37].

Fig. 4. New Zealand MTG unit [40].

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

59

Fig. 5. ZSM-5 conversion reactor [39].

to the HTG reactor. The durene level is cut down there to about 2 wt.%. The blended gasoline product has a research octane number (RON ) of 92.2 and a motor octane number (MON ) of 82.6.

4. Fluid-bed MTG and MTO process Mobil Research and Development Corp., Union Rheinische Braunkohlen Kraftsto AG and Uhde GmbH (Dortmund, Germany) jointly designed, engineered and operated a uid-bed MTG demonstration plant which was located in Wesseling near Cologne in Germany. The project was partly nanced by the American and German Governments. Details of this project are described by Gierlich et al. [5], Grimmer et al. [6 ], Penick et al. [34], and by Edwards and Avidan [41]. Bench-scale experiments on the uid-bed MTG process were executed at Mobil in the 1970s [42]. Subsequently, a 0.1 m diameter, 7.6 m tall, 0.636 m3 day1 pilot plant was successfully operated. From December 1982 to the end of 1985 a 15.9 m3 day1 demonstration plant produced gasoline in Wesseling. Details of the entire plant and

the uidized bed reactor are given in Figs. 6 and 7, respectively. To study uid dynamics and verify the mechanical design basis, a full-scale cold ow model (CFM ) was employed. This non-reacting model proved very useful for optimizing bae design and catalyst circulation strategies. Several dierent bae designs, horizontal and vertical arrangements, were tested in the CFM using dierent experimental techniques like tracer tests, capacitance probes, and bed expansion analysis. The experiments indicated that horizontal baes are eective in breaking bubbles. When the catalyst nes concentration is sucient (higher than 15% <40 mm), bubbles are small and the bed shows a homogeneous appearance. Edwards and Avidan [41] employed a homogeneous, one-dimensional dispersion model, combined with reaction kinetics to calculate the uidized bed reactor performance. The Peclet numbers were calculated from SF 6 tracer experiments. The supercial gas velocities ranged from 0.3 to 1 m s1. It turned out that the axial dispersion model can be used to predict conversion in a turbulent uid-bed reactor under the condition of an overall homogeneity of the bed. This is achieved in a turbulent bed of Group A powder by operation at supercial velocities above 0.3 m s1, with a minimum of nes. A scale-up of the 15.9 m3 day1 uid-bed MTG process was possible from bench-scale without loss of conversion eciency. The uid-bed process has the following advantages compared to the xedbed process: (1) Excellent heat transfer properties of a uidized bed permit direct steam generation in coils immersed in the reactor. (2) Continuous regeneration of the catalyst (constant catalyst activity) and a uniform bed temperature result in a constant gasoline quality. (3) Transient temperature proles during heat-up and cool-down are also uniform and stable. (4) The specic throughput in a uid-bed system is higher. (5) Higher octane numbers were found (see Table 1). (6) The yield of gasoline including alkylate is at least 7.5% higher.

60

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

Fig. 6. Fluid-bed MTG demonstration plant [49].

Fig. 7. Fluidized bed reactor of the demonstration plant.

(7) The durene content is lower (maximum 5 wt.%) (8) Liquid injection, a unique feature in the uid bed, provides the exibility to tailor the steam balance as per requirement. (9) Specic investment cost is lower. A simplied scheme of the uidized bed plant in Fig. 6 shows that crude methanol is vaporized and fed into the reactor. To improve the overall process economy, the water can be removed from the crude methanol. The heat of reaction can be used to generate high pressure steam. The heat exchanger can be operated inside or outside the reactor. Coils immersed inside the reactor turned out to be most ecient. The catalyst is continuously withdrawn and regenerated by partially burning o the coke. The rate of catalyst circulation through the regenerator determines the average activity of the catalyst in the reactor. A bank of dierent cyclones remove catalyst dust from the reactor euent. The hydrocarbon products are processed in a gas fractionation unit to produce C+ hydrocarbons, alkylation feed 5 (C /C fractionation) and olen recycle. Heavy 3 4 gasoline treatment will be required only if a further reduction of durene content is necessary. A com-

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

61

mercial concept of the uid-bed MTG process was developed by Uhde [6 ]. At a gas price of US$1 GJ1 and a GTG plant of 6180 ton day1 capacity, 1 l of unleaded premium gasoline will cost US$0.19 (data for 1987). The uid-bed technology is ready for commercialization. The plant at Wesseling was also operated for the production of olens at a pressure between 2.2 and 3.5 bar and a temperature of about 500C. At steady state conditions the olen yield was more than 60%, although the catalyst was not tuned to olen production. Lurgi has considered a tubular xedbed MTG process, which oers some advantages over the New Zealand xed-bed process [43]. However, the uid-bed MTG process has overtaken this concept.

5. The methanol-to-olens (MTO) and the mobil olen-to-gasoline/distillate (MOGD) process As light olens are intermediates in the MTG reaction scheme, methanol can also be employed to produce light olens. Higher reaction temperatures, lower pressures and a high ratio of ( lower acidity zeolites) favor the production of light olens. Chang et al. [44,46 ], Chang [45], Schoenfelder et al. [47], and Tshabalala and Squires [48] report bench-scale measurements on the MTO process. As already mentioned, an MTO demonstration plant was also operated. Union Carbide developed a process to convert methanol to olens using a SAPO catalyst. The olens yield exceeded 90%. The process can be modied for high ethylene and propylene yield (about 60%) [7,8,10]. The basic ow-sheet of the MTO process is the same as that of the uidized bed MTG process. Avidan [49] described the results obtained in the Wesseling plant. A further development is the Mobil olens-togasoline/diesel (distillate) (MOGD) process [50,51]. In this process, gasoline and distillate selectivity is greater than 95% of the olens in the feed and gasoline/distillate product ratios range from 0.2 to >100. In order to obtain high octane numbers, shape-selectivity is tuned such that mostly methyl-branched iso-olens (C C ) are 5 20 produced. The C to C fraction needs hydrogen10 20

ation. MOGD olen product distribution is determined by thermodynamic, kinetic, and shapeselective limitations. A large-scale MOGD test run was executed in a Mobil renery in 1981. A commercially produced zeolite catalyst was employed. A simplied scheme of this process is presented in Fig. 8 [49]. In general, four xed-bed reactors, three on-line and one in regeneration are used. The three on-line reactors are operated in series with interstage cooling and liquid recycle to control the heat of reaction. The olens feed is mixed with a gasoline recycle stream and passed, after heating, through the three reactors. In order to generate a gasolinerich stream for recycle to the reactors, a fractionation is used. The recycle improves distillate selectivity. The MTO and MOGD process can be combined. A possible process ow-sheet is presented in Fig. 9. High-octane MTO gasoline is partially split o before the MOGD section and is later blended with MOGD gasoline. The raw distillate is hydrotreated and can be fractionated into various products. Typical distillate and gasoline yields from the olens yield obtained in the 15.9 m3 day1 MTO plant at Wesseling are 50/50 w/w. This ratio can be variied considerably. The gasoline is olenic and aromatic, and of better quality than FCC gasoline. The MON and RON are 93.0 and 85.0, respectively. The durene content is very small. Its physical properties, such as ash point, boiling range and viscosity, are comparable with conventional distillate fuels. MOGD diesel has a density of 0.8 instead of 0.86. MOGD can also be used as jet fuel.

6. UOP/HYDRO MTO process UOP (Des Plaines, Illinois) and Norsk Hydro (Oslo, Norway) developed a new gas-to-olens (GTO) and MTO process which produces a very high yield of ethylene (48%) and propylene (33%). As a catalyst an attrition resistant SAPO-34 was employed. A simplied process owsheet is presented in Fig. 10 [52,53]. The UOP/HYDRO MTO unit employs a uidized-bed reactor coupled to a uidized-bed regener-

62

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

Table 1 Process conditions and product yields from MTG processes Fixed-bed Methanol/water charge (w/w) Dehydration reactor inlet temperature (C ) Dehydration reactor outlet temperature (C ) Conversion reactor inlet temperature (C ) Conversion reactor outlet temperature (C ) Pressure (kPa) Recycle ratio (mol/mol ) charge WHSV (h1) Yields (wt.%) of methanol charged Methanol+ether Hydrocarbons Water CO, CO 2 Coke, other 83/17 316 404 360 415 2170 9.0 2.0 Fluid-bed 83/17 413 413 275 1.0

0.0 43.4 56.0 0.4 0.2 100.0

0.2 43.5 56.0 0.1 0.2 100.0

Hydrocarbon product (wt.%) Light gas Propane Propylene Isobutane n-Butane Butenes C +gasoline 5 Gasoline (including alkylate) [RVP-62 kPa (9 psi)] LPG Fuel gas Gasoline (RON ) RVP=Reid vapour pressure. WHSV=weight hourly space velocity. RON=research octane number. LPG=liquied petroleum gas.

1.4 5.5 0.2 8.6 3.3 1.1 79.9 100.0

5.6 5.9 5.0 14.5 1.7 7.3 60.0 100.0

85.0 13.6 1.4 100.0 93

88.0 6.4 5.6 100.0 97

ator. The heat of reaction is controlled by steam generation. The catalyst is sent continuously to the regenerator, where the coke is burned o and steam is generated to remove the heat resulting from burning. After heat recovery, the reactor euent is cooled, and some of the water is condensed. After compression, the euent passes through a caustic scrubber to remove CO and to a dryer to remove 2 water. The reactor section is quite similar to the Mobil/Uhde process. The euent then proceeds

to the product recovery section, which includes a demethanizer, a deethanizer, a C splitter, a C 2 3 splitter, and a depropanizer. Polymer-grade ethylene and propylene are produced from these fractionation columns along with methane, ethane, propane, and C product streams. The 4 UOP/HYDRO MTO process can be economically viable in dierent scenerios [52,53]: (1) Production of methanol at a remote gas eld site and transportation of the methanol to an MTO plant located at the olens users site.

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

63

Fig. 8. MOGD demonstration plant [49].

Fig. 9. MTO/MOGD process [6 ].

64

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

Fig. 10. UOP/Hydro MTO process for polymer-grade products [53].

(2) An integrated GTO complex at the gas eld site and transportation of olens or polyolens products to customers. (3) Increased olens production and feedstock exibility at an existing naphtha or ethane propane cracker facility by installing an MTO reactor section and feeding into a revamped cracker fractionation section. (4) A smaller MTO unit using methanol produced in a single-train methanol plant to meet the local demand for olens or polyolens or both. (5) The UOP/HYDRO MTO process and catalyst have been successfully demonstrated and are currently available for license.

7. Haldor Topse TIGAS process, Akron LP-DME-to-gasoline process (DTG) Topse has developed a low investment process for the conversion of natural gas to gasoline [54]. As many future synthetic fuel plants will be built in remote areas where the price of natural gas is very low and not related to gasoline, low investment cost is essential, as the investment-related

costs determine a high proportion of the cost of production due to the low energy price. The TIGAS process was developed for just this purpose. A block scheme is presented in Fig. 11. In the TIGAS process the two process steps, MeOH synthesis and the MTG process, are integrated into one single synthesis loop without isolation of MeOH as an intermediate. The experimental program for the TIGAS process lasted 3 years and was terminated in January 1987. A demonstration plant at Houston was operated for 10 000 h. The purpose of the process development work on the integrated gasoline synthesis was to modify the three process steps synthesis gas production, oxygenate synthesis and the MTG process in order to be able to operate all steps at the same pressure and the last two steps in one single synthesis loop [54]. By selecting combined steam reforming and autothermal reforming for the synthesis gas production, and by using a multifunctional catalyst system, producing a mixture of oxygenates instead of only MeOH, the front-end and the oxygenate synthesis can operate at the same pressure (~20 bar). The TIGAS process avoids the compression of syngas to about

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966

65

Fig. 11. Topse TIGAS process [54].

50100 bar required by a conventional methanol plant. This reduces the capital and operating costs of the combined synthesis and conversion loops. The overall reaction of the TIGAS process is 3CO+3H PCH OCH +CO (46) 2 3 3 2 The DME can be converted into hydrocarbon products in a separate MTG reactor. The operation of the MTG process and the oxygenate synthesis in one loop will then only call for minor changes in the MTG process. A separation unit leads to the products. The University of Akron has developed a process which converts syngas directly to DME using LP-DME synthesis [55]. One-step conversion of syngas to DME improves the per-pass conversion and reactor productivity over syngas to methanol. A dual catalyst system is based on a combination of Cu/ZnO/Al O catalyst and gamma-alumina 2 3 catalyst. This conversion oers some advantages. First, the liquid phase is lean in methanol because of in-situ conversion to DME over gamma-alumina. Second, water produced by both methanol synthesis (CO hydrogenation) and DME synthesis 2 (methanol dehydration) constantly shifts the forward water gas shift reaction. This is the special feature of the one-step conversion of syngas to DME. The one-step conversion of syngas to DME improves the volumetric productivity by as much as 100% over that of syngas to methanol conversion. This is because conversion of syngas to DME is not limited by chemical equilibrium as is syngas conversion to methanol. The process can be adapted to coal-based syngas. MacDougall [56 ], Rostrup-Nielsen [57], and

Parkyns et al. [58] discuss some further aspects of synfuel production.

8. Conclusion A broad variety of well-tested processes for the production of hydrocarbons from methanol is available. Their future usage is determined by natural gas and methanol prices.

References
[1] [2] [3] [4] [5] C.D. Chang, A.J. Silvestri, CHEMTECH 10 (1987) 624. C.D. Chang, Catal. Rev.-Sci. Engng 25 (1983) 1. C.D. Chang, Catal.-Rev.-Sci. Engng 26 (1984) 323. M. Stocker, Microporous Mesoporous Mater. (1999) 3 (this issue). H.H. Gierlich, K.H. Keim, N. Thiagarajan, E. Nitschke, A.Y. Kam, N. Daviduk, Paper presented at the 2nd EPRI Conference Synthetic Fuels Status and Directions, San Francisco, CA, 1985. H.R. Grimmer, N. Thiagarajan and E. Nitschke, in: D.M. Bibby, C.D. Chang, R.W. Howe, S. Yurchak ( Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p. 273. S.W. Kaiser, US Patent 4 499 327, 1985. S.W. Kaiser, US Patent 4 524 234, 1985. S.W. Kaiser, Arabian J. Sci. Engng 10 (1985) 361. G. Pop, G. Musca, D. Ivanescu, E. Pop, G. Maria, E. Chirila, O. Muntean, Chem. Ind. 46 (1992) 443. C.D. Chang, J.C.W. Kuo, W.H. Lang, S.M. Jacob, J.J. Wise, A.J. Silvestri, Ind. Engng Chem., Process Des. Dev. 17 (1978) 255. J.R. Anderson, T. Mole, V. Christoo, J. Catal. 61 (1980) 477. S.E. Voltz, J.J. Wise, Development studies on conversion

[6 ]

[7] [8] [9] [10] [11]

[12] [13]

66

F.J. Keil / Microporous and Mesoporous Materials 29 (1999) 4966 of methanol and related oxygenates to gasoline. Final Report, US ERDA Contract no. E (49-18)-1773, 1976. C.D. Chang, A.J. Silvestri, J. Catal. 47 (1977) 249. N.Y. Chen, W.J. Reagan, J. Catal. 59 (1979) 123. C.D. Chang, W.H. Lang, R.L. Smith, J. Catal. 36 (1979) 169. Y. Ono, E. Imai, T. Mori, Z. Phys. Chem., N.F. 115 (1979) 99. Y. Ono, T. Mori, J. Chem. Soc., Faraday Trans. 77 (1981) 2209. C.D. Chang, Chem. Engng Sci. 35 (1980) 619. R.G. Anthony, Chem. Engng Sci. 36 (1981) 789. H.J. Doelle, J. Heering, L. Riekert, J. Catal. 71 (1981) 27. P.H. Schipper, F.J. Krambeck, Chem. Engng Sci. 41 (1986) 1013. U. Sedran, A. Mahay, H.I. de Lasa, Chem. Engng Sci. 45 (1990) 1161. U. Sedran, A. Mahay, H.I. de Lasa, Chem. Engng J. 45 (1990) 33. P.L. Benito, A.G. Gayubo, A.T. Aguayo, M. Castilla, J. Bilbao, Ind. Engng Chem. Res. 35 (1996) 81. A.G. Gayubo, P.L. Benito, A.T. Aguayo, I. Aguirre, J. Bilbao, Chem. Engng J. 63 (1996) 45. A.N.R. Bos, P.J.J. Tromp, H.N. Akse, Ind. Engng Chem. Res. 34 (1995) 3808. R. Mihail, S. Straja, G. Maria, G. Musca, G. Pop, Chem. Engng Sci. 38 (1983) 1581. O. Iordache, G. Maria, G. Pop, Ind. Engng Chem. Prod. Res. Dev. 27 (1988) 2218. S.L. Meisel, Y.P. McCullough, C.H. Lechthaler, P.B. Weisz, CHEMTECH 6 (1976) 86. D.M. Bibby, C.D. Chang, R.W. Howe, S. Yurchek (Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p. 251 S.A. Tabak, S. Yurchak, Catalysis Today 6 (1990) 307. J.E. Penick, W. Lee, J. Maziuk, Int. Symp. on Chem. Reaction Engng (ISCRE-7) Boston, American Chemical Society, 1982 p. 19. J.E. Penick, W. Lee, J. Maziuk, in: J. Wie, C. Georgahis ( Eds.), Chemical Reaction Engineering Plenary Lectures, ACS Symposium Series 226, ACS, Washington, 1983, p. 19. A.J. Silvestri, Mobil Methanol-to-gasoline Process, 181st ACS National Meeting, Atlanta, 1981. C.J. Maiden, in: D.M. Bibby, C.D. Chang, R.W. Howe, S. Yurchak ( Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p.1 J.Z. Bem, in: D.M. Bibby, C.D. Chang, R.W. Howe, S. Yurchak ( Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p. 663. C.D. Chang, Catalysis Today 13 (1992) 103. D.E. Krohn, M.G. Melconian, in: D.M. Bibby, C.D. Chang, R.W. Howe, S. Yurchek (Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p. 679. K.G. Allum, A.R. Williams, in: D.M. Bibby, C.D. Chang, R.W. Howe, S. Yurchak ( Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p. 691. M. Edwards, A. Avidan, Chem. Engng Sci. 41 (1986) 829. D. Liedermann, S.M. Jacob, S.E. Voltz, J.J. Wise, Ind. Engng Chem. Proc. Des. Dev. 17 (1978) 340. G. Hochgesand, C. Hafke, K. Schmitt, Int. Coal and Gas Conversion Conference, Pretoria, 1987. C.D. Chang, J.N. Miale, R.F. Socha, J. Catal. 90 (1984) 84. C.D. Chang, Cat. Rev. Sci. Engng 26 (1984) 323. C.D. Chang, C.T.-W. Chu, R. Socha, J. Catal. 86 (1984) 289. H.J. Schoenfelder, J. Hinderer, J. Werther, F.J. Keil, Chem. Engng Sci. 49 (1994) 5377. S.N. Tshabalala, A.M. Squires, AIChE J. 42 (1996) 2941. A.A. Avidan, in: D.M. Bibby, C.D. Chang, R.W. Howe, S. Yurchak (Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p. 307. S.A. Tabak, F.J. Krambeck, Hydrocarbon Proc. 64 (1985) 72. S.A. Tabak, F.J. Krambeck, W.E. Garwood, AIChE J. 32 (1986) 9. B.V. Vora, R.A. Lentz, T.L. Marker, H. Nilsen, S. Kvisle, T. Fuglerud, 5th World Congress of Chemical Engineering, San Diego, CA, 1996. B.V. Vora, R.A. Lentz, T.L. Marker, World Petrochemical Conference CMAI, Houston, TX, Petrochemical Review, DeWitt & Co., Houston, 1996, p. 2. J. Topp-Jrgensen, in: D.M. Bibby, C.D. Chang, R.F. Howe, S. Yurchak ( Eds.), Methane Conversion, Studies in Surface Science and Catalysis, vol. 36, Elsevier, Amsterdam, 1988, p. 293. S. Lee, M. Gogate, C.J. Kulik, Fuel Sci. Technol. Int. 13 (1995) 1039. L.V. MacDougall, Catalysis Today 8 (1991) 337. J.R. Rostrup-Nielsen, Catalysis Today 21 (1994) 257. N.D. Parkyns, C.I. Warburton, J.D. Wilson, Catalysis Today 18 (1993) 385.

[14] [15] [16 ] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26 ] [27] [28] [29] [30] [31]

[38] [39]

[40]

[41] [42] [43] [44] [45] [46 ] [47] [48] [49]

[50] [51] [52]

[32] [33]

[53]

[34]

[54]

[35] [36 ]

[55] [56 ] [57] [58]

[37]

Вам также может понравиться