Sie sind auf Seite 1von 18

3583

MICROPOROUS AND
MESOPOROUS MATERIALS
Microporous and Mesoporous Materials 29 (1999) 49-66

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: Mobil's methanol-to-gasoline (MTG) plant in New Zealand, the
fluidized bed MTG and methanol-to-olefins process; Mobil's olefin-to-gasoline/distillate (MOGD) process; the MTO
plant developed by UOP and Norsk Hydro; Haldor Topsee's TIGAS process. The developments of a liquid phase
dimethyl ether synthesis (LP-DME) process by the Ahron University are also presented. CD 1999 Elsevier Science B.V.
All rights reserved.
Keywords: Commercial plants; Kinetics; Methanol-to-gasoline; Methanol-to-hydrocarbons; Methanol-to-olefins; Review

1. Introduction
Mobil's novel synthetic gasoline process, based
on the conversion of methanol to hydrocarbons
over zeolite catalysts, was the first 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 Oil's 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.
Mobil's Central Research team tried to find out
whether methanol could serve as a precursor to a
'C 1 olefin' in alkylating isobutane, to form, presumably, neopentane. ZSM-5 was the first 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

CH3 O H --[CH 2] + H20

1387-1811/99/5 - sec front matter 1999 Elsevier Science B.V. All rights reserved.
P11: S1387-1811(98)00320-5

(1)

50

F.J. Keil / Microporous and Mescporous Materials 29 ( 1999) 49-66

where [CH2] is the average composition of the


hydrocarbon product. More detailed investigations
suggested that the main reaction steps are:
-u20
2CH3OH
CH3OCH3 ' light olefins
+1120
(2)
*higher olefins+nlisoparaffins
+ aromatics + naphthenes
As can been seen from the reaction scheme,
methanol is first dehydrated to dimethylether
(DME). The equilibrium mixture of methanol,
DME and water is then converted to light olefins
(C2C4). A final reaction step leads to a mixture
of higher olefins, nfiso-paraffins, aromatics and
naphthenes. An interruption of the reaction leads
to a production of light olefins instead of gasoline.
An appropriate process for this purpose was developed by Mobil, the so-called methanol-to-olefin
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
Stcker [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 6401 per day
(four barrels per day), fixed-bed pilot plant to
demonstrate the feasibility of the gas-to-gasoline
(GTG) process. Reports to New Zealand's Liquid
Fuels Trust Board (LFTB) by Lurgi and Badger
indicated confidence 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 first step,
methanol is synthetized from natural gas (Maui

field). 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 final gasoline produced does not need further
refining and attains the quality of unleaded premium gasoline. Owing to some reasons presented
below, from 1981 and 1984, Mobil, Union
Rheinische Braunkohlen Kraftstoff 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 fluid bed reaction system for MTG
and MTO technology [5,6]. Union Carbide also
developed a process to convert methanol to olefins
in 1986 using a silicoaluminophosphate (SAPO)
catalyst. The olefins yield exceeded 90%, and they
report that the process could be modified for high
ethylene and propylene yield (about 60%) [7-10].
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 Stcker has reviewed the details of catalyst
and reaction mechanisms of the methanol-tohydrocarbons (MTHC) processes, especially the
MTG, MTO and MOGD (Mobil's olefin-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) 49-66

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 find 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 olefin-forming step, and
is essentially at equilibrium [11,12]. Thus the
equilibrium oxygenate mixture can be conveniently
treated as a single kinetic species. Based an 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 sp 3 Insertion into CH bonds, forming higher alcohols or
ethers, which can dehydrate to form olefins and
which can add :CH 2 to form higher olefins. Chen
and Reagan [15] discovered that the oxygenate
disappearance is autocatalytic over ZSM-5. They
proposed the following scheme:
B

k2
A + B, 13

kt

A. B
k2

A+B C
k3

B+C C
k4

C. D
where A represents the oxygenates, B :CH 2 groups,
C olefins, and D paraffins + aromatics. The formal
kinetics are:
dA/dt=kiA +k2AB

11

(12)

(3)

dC/dt = k2AB +413C k4C

(13)

(4)

Assuming the steady-state condition for B and


eliminating time resulted in an expression like this:

k3

B-- C

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 first
order in oxygenates.
(3) Consumption of the reactive intermediate is
first order in oxygenates.
(4) Olefins can be treated as a single kinetic
species.
(5) Disappearance of olefins is first order in
olefins.
These assumptions lead to the following set of
reactions:

dB/dt = k A k2AB k3BC

kt
A

51

(5)

du/dA = 1/A[( 1 +K1 u)/( 2 +K1 u)( 1 K210 + 14


(14)

where A represents oxygenates, B olefins, and C


aromatics + paraffins. The rate of oxygenate disappearance is first order in oxygenates [13].
The experimental data could be fitted according
to the expression
dA/dt =k 1A +k2AB

(6)

where u=C/A, KI = 1c3/k2 and 1C2= k4/I c 1 .


Anthony [20] improved Chang's 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

52

F.J. KehlMicroporous and Mesoporaus Materials 29 ( 1999) 49-66

rate law
r = ki PcH30/( 1 +k2 PH20 )

(15)

Schipper and Krambeck [22] have obtained


results in a pilot plant with an adiabatic fixed-bed
reactor, which was operated under reaction-regeneration cyles and under conditions in which there
was irreversible deactivation. They defined catalyst
activity for a given time on stream, , 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:

catalyst. The following two lumped models were


able to treat the experimental data consistently:
Model I:
k,
A- B

(21)

k2

A+

(22)

A +C-+ D

(23)

k3

and

(16)

=af.

The reaction rate of each individual step, rb is


the product of the reaction rate for the fresh
catalyst rio and the activity :
= rio

B+E-* F

(24)

k4

C +E- F

(25)

(17)
k4

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 = -1C<, exp(-Ez/RT)

am;

m> 1

( 18)

The loss of activity due to coke deposition is


expressed as:
df/d = - ICcok /3" ; n > 1

(19)

By combination of the previous equations, a total


deactivation rate is obtained

D + E-+ F

(26)

Model II:
3ki
A-4 B+C+D

(27)

k2
B +E- F

(28)

k3
C+E- F

(29)

k4

d/dt=a{-Kcak " - Kr 2 )

(20)

This equation indicates that the rate of change


of total catalyst activity is related not only to the
total activity of the catalyst, , 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 first reaction-regeneration 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 modifications proposed
by Anthony [20] over the 302-370C temperature
range. In a further paper Sedran et al. [24] compared different lumped kinetic models for methanol conversion with hydrocarbons on a ZSM-5

D+E- F

(30)

where A e Methanol+ DME, B methene, C


propene, D = butene, Feparaffins.
The authors found that an exponential activity
function of the type

4=40 exp(-Z/W)

(31)

can be applied to the observed values, where


Hcl 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) 49-66

Benito et al. [25] considered the effect 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 fixedbed integral reactor. ZSM-5 with a Si/Al ratio of
24 was employed. The temperature range was
300-375C. The best fitting model was the one
proposed by Schipper and Krambeck [22]:
kt

MEOH/DME(A)+ light olefins (C)

(32)

ka

2C products (D)

(33)

k3

A+D' D

(34)

k.

C+D D

(35)

The light olefms (ethylene, propylene) can polymerize to form products in the gasoline boiling
range D. Benito et al. [25] found the following
kinetic constants:
k1 =0.733 x 10 13 exp( 33358/RT)

(36)

k2 =0.127 x 108 exp( 17633/RT)

(37)

k3 = 0.204 x 1012 exp(-27987/RT)

(38)

k4 =0.634 x 106 exp( 15855/RT)

(39)

The kinetic equations are:


rAo = k iX A k3X AXD =dXA/cfr

(40)

rco =k t XAk2,11-1c4XcXD=dXcldr

(41)

The Xi represent weight fractions of lump i (on a


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 [E (k di X 1)]ad

(42)

53

where a is the activity defined as


a = ri (t)Iric,(1 = 0)

(43)

The kdi are the kinetic constants for deactivation


by coke formation for lump i. The experimental
results revealed that the following equation can
describe the change of activity with time:
da/dt =(kd A XAkdCXC+kdDXD)a

(44 )

The kdi were found to be different. In a further


paper Gayubo et al. [26] state that the models
of Chen and Reagan [15] and Schipper and
Krambeck [22] adequately fit the experimental
results. The authors emphasize the following
aspects: (1) ethylene and propylene are identified
as prirnary products present in the gas phase; (2)
an oligomerization step of light olefins is established; and (3) a methylation step of products is
established. Although die 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-olefins 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-fiow, fixed-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 firstorder rate constant of roughly 250 4 mj2 s'.
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 7-8 wt.% of coke must be
present on the catalyst. Consecutive reactions
cannot be neglected. The net effects 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 final kinetic network of 10 first-order and
two second-order reactions describes the experimental results satisfactory. The final reaction

F.J. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66

54

scheme for model discrimination was the


following:
MEOH

2
3

e
5.

Methane

> Ethene
Ie
)

-7-42..
Propene 40 ' Coke

11
) Prop ane

44
)

6
)

42

(10) The aromatics undergo condensation.


(11) The aromatics undergo alkylation with
methanol.
(12) The paraffins undergo demethanization forming olefins 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

SUM C4 <
SM C5

) Coke
(45)
Eqs. (8) and (12) are of second order. The
formation of ethene from propene is of first order
in methanol and propene. The rate of formation
of ethene from butene depends an 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 etherification 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 olefins.
(3) The carbene attacks the olefins, forming
higher olefins.
(4) The carbene attacks the hydrogen, forming
methane.
(5) The light olefins generate carbeniums ions.
(6) The carbenium ions attack the light olefins
forming higher olefins (oligomerization).
(7) The carbenium ions attack the higher olefins
forming paraffins and dienes.
(8) The carbenium ions attack the dienes forming
paraffins and cyclodienes.
(9) The carbenium ions attack the cyclodienes
forming paraffins and aromatics.

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
fields into methanol and then into ca. 700 000 tons
per day of gasoline via Mobil's fixed-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 methanol - 1 . The adiabatic temperature rise is about 600C. A simplified block
diagram of the MTG process is given in Fig. 1.
Methanol is vaporized and fed into the fixedbed 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
310-320C and about 26 bar pressure. Approximately 15-20% of the heat of reaction is released
in this first step, which is controlled by chemical
equilibrium. The DME reactor effiuent 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 350-370C. About 85% of
the reaction heat is released in the conversion
reactor. After cooling, the effluent is separated into
three phases: gas, liquid water, and liquid hydrocarbons. Most of die gas is recycled to the ZSM-5
reactor. The water contains a small amount of
oxygenates which are treated in a biological waste

F.J. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66

55

s LPG

c2-

i
Superheat
Vaporize
Preheat

DME
Reactor

ZSM-5 --
Reactors
-

Grude
Methanol

HP
Sep.

3ns

i
Regeneration

Light Gasoline

DKWlation

HGT

Heavy Gasoline
Water to
Treating
Finished
Gasoline

Fig. 1. Block diagram of the fixed-bed MTG process (34

water treatment plant. The hydrocarbon product


is distilled. It contains mainly raw gasoline, dissolved hydrogen, carbon dioxide and light hydrocarbons (C 1C4). The non-hydrocarbons, C 1C3
and a part of the C4 hydrocarbons are removed
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 specification finished
gasoline.
The development of the MTG process has been
described in the literature, see for example Refs.
[31-33]. Bench-scale studies of the MTG process
were executed for fixed-bed and fluid-bed reactors.
In Fig. 2 a two reactor configuration is shown
[31]. The fixed-bed reactor configuration 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 first
deactivate the front part of the bed. Use of a
sufficient catalyst volume permits a fixed-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


profile along the reactor is formed. There is a very
small change in the shape of the profile as the
catalyst ages. This means that coke formation
downstream of the main reaction zone is low or
its level does not appear to affect catalyst activity.
After some time, the reaction zone approaches the
reactor outlet and significant 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 effective 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 paraffins decreases
and the isoparaffin content increases. The aromatics content decreases, and the content of olefins
and naphthenes increases. The isoparaffins compensate for the aromatics, so that the gasoline
octane number varies only little with time. One
should keep in mind that the fixed-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 an 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
different 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) 49-66

Recycle Gas
ti

Preheater
32 mm ID
Resttor
3
210 em of
Catalyst

Dehydration
Reaetor

Highliesstim
Separater

4,

Conversion
Reactor

r> Liquid

Fig. 2. Schematic of fixed-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 first 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 day of


methanol was built and operated.
The major objective of the demonstration plant
was to verify the bench-scale results. The only
different 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 effect can be
related to the slower rate of movement of the
catalyst bed temperature profile. 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, transallcylation, ring
saturation, and dealkylation/cracking. The durene

F.J. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66

content is reduced to less than 2 wt.%. The gasoline


produced in the demonstration plant was used in
an automobile test fleet. 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 first 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 52-55 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 simplified 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 conceming 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 effluent before flowing into the first-stage
dehydration (DME) reactor. In this reactor crude

57

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
310-320C and 27 bar and releases 15-20% of the
overall heat of reaction. The DME reactor effluent
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 350-366C. The inlet pressures
are 19-23 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 fifth reactor is thus either in the regeneration
mode or on standby. The hot reactor effluent is
first cooled by generating steam, then fresh crude
methanol and recycle gas is heated. The reactor
effluent is then further cooled to 25-35C 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 CO2. This gas is
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 1C4) is sent to the
de-ethanizer. The off-gases, including methane,
ethane and some C3, together with a purge gas
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 C3 and part of the C4
components are removed overhead to the fuel gas
system. C4/C5 components are withdrawn as a
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) 49-66

I Utility and I
Instrument Air

Wate

te
E
=

4,

Plant

Fuel Ges

to

BFW
Nature Gas
CL

Crude
Methanol
Storage

Staam
500 Uh

Auebe

Staam Boiler

II

4,

Liquid
1
Nitrogen Plant

4,

Staam System MTG Fteactor


NM High Vapor Pressure
Gas
Gasotote

e I_
B

Methanol
Synthesis
(2 Trains)

Heavy
Gasohne
Storage

bi Ligiti 1..
g
w
g

EfIluent le_
Treatment

Light Gasohne

Gasohne
storage

T
Blendlog 1)1>
lk
Gasohne
700000 tia

Die

Fig. 3. Simplified block flow diagram of the New Zealand GTG plant [37].

>2>
a

A
ii

h'' n

reccto

223750 et

=Kn --II,
ILCI
Garn

bim

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

F.J. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66

Inlet

Catalyst Dump
Outlet
Fig. 5. ZSM-5 conversion reactor [391.

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 Kraftstoff AG and Uhde
GmbH (Dortmund, Germany) jointly designed,
engineered and operated a fluid-bed MTG demonstration plant which was located in Wesseling near
Cologne in Germany. The project was partly
financed 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 an the fluid-bed MTG
process were executed at Mobil in the 1970s [42].
Subsequently, a 0.1 m diameter, 7.6 m tall,
0.636 m3 day -1 pilot plant was successfully operated. From December 1982 to the end of 1985 a
15.9 m3 day -1 demonstration plant produced gasoline in Wesseling. Details of the entire plant and

59

the fluidized bed reactor are given in Figs. 6 and


7, respectively.
To study fluid dynamics and verify the mechanical design basis, a full-scale cold flow model
(CFM) was employed. This non-reacting model
proved very useful for optimizing baffie design and
catalyst circulation strategies. Several different
baffle designs, horizontal and vertical arrangements, were tested in the CFM using different
experimental techniques like tracer tests, capacitance probes, and bed expansion analysis. The
experiments indicated that horizontal baflies are
effective in breaking bubbles. When the catalyst
fines concentration is sufficient (higher than 15%
<40 um), bubbles are small and the hed Shows a
homogeneous appearance. Edwards and Avidan
[41] employed a homogeneous, one-dimensional
dispersion model, combined with reaction kinetics
to calculate the fluidized bed reactor performance.
The Peclet numbers were calculated from SF 6
tracer experiments. The superficial gas velocities
ranged from 0.3 to 1 m s -1 . lt turned out that the
axial dispersion model can be used to predict
conversion in a turbulent fluid-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 superficial velocities
above 0.3 m s -1 , with a minimum of fines. A
scale-up of the 15.9 m 3 day -1 fluid-bed MTG process was possible from bench-scale without loss of
conversion efficiency. The fluid-bed process has
the following advantages compared to the fixedbed process:
(1) Excellent heat transfer properties of a fluidized
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 profiles during heat-up
and cool-down are also uniform and stable.
(4) The specific throughput in a fluid-bed system
is higher.
(5) Higher octane numbers were found (sec
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) 49-66


C
a:
Cydone

Raw Gas
Cooler
Cyclone

Regennatur

Condenso

7
Hltrogen
+ Air

Fluldlzed
ed
Rautor
.

(i)
Hefter

retmgen --E
Hefter

e.

Condlensor4 r G

Coo er
Compressor t.

ux Drum

Debutaid.
Reboiler

Pump

Haat Transfer OD Separation


Drum

0
I
imenimi Cs.
Gasoline
Cooler
Weber

+.1
Hefter

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

(7) The durene content is lower (maximum


5 wt.%)
(8) Liquid injection, a unique feature in the fluid
^
bed, provides the flexibility to tailor the steam
Reactor
balance as per requirement.
(9) Specific investment cost is lower.
A simplified scheme of the fluidized bed plant
Catalyst from
E
a
Regenerator
in Fig. 6 shows that crude methanol is vaporized
mi
and fed into the reactor. To improve the overall
Catalyst to
process economy, the waten can be removed from
Cooler
Catalyst to
the crude methanol. The heat of reaction can be
Regenerator
used to generate high pressure steam. The heat
z' exchanger can be operated inside or outside the
reactor. Coils immersed inside the reactor turned
out to be most efficient. The catalyst is continuously withdrawn and regenerated by partially
burning off the coke. The rate of catalyst circulation through the regenerator determines the
E
average activity of the catalyst in the reactor. A
cr;
bank of different cyclones remove catalyst dust
from the reactor effluent. The hydrocarbon proCatest from
ducts are processed in a gas fractionation unit
Cooler
to produce C; hydrocarbons, alkylation feed
Methanol
(C3/C4 fractionation) and olefin recycle. Heavy
v
Feed
gasoline treatment will be required only if a further
reduction of durene content is necessary. A comFig. 7. Fluidized bed reactor of the demonstration plant.
SteamOWM

F.J. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66

mercial concept of the fluid-bed MTG process was


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

5. The methanol-to-olefins (MTO) and the mobil


olefin-to-gasoline/distillate (MOGD) process
As light olefins are intermediates in the MTG
reaction scheme, methanol can also be employed
to produce light olefins. Higher reaction temperatures, lower pressures and a high ratio of (lower
acidity zeolites) favor the production of light
olefins. Chang et al. [44,46], Chang [45],
Schoenfelder et al. [47], and Tshabalala and
Squires [48] report bench-scale measurements an
the MTO process. As already mentioned, an MTO
demonstration plant was also operated. Union
Carbide developed a process to convert methanol
to olefins using a SAPO catalyst. The olefins yield
exceeded 90%. The process can be modified for
high ethylene and propylene yield (about 60%)
[7,8,10]. The basic flow-sheet of the MTO process
is the same as that of die fluidized bed MTG
process. Avidan [49] described the results obtained
in the Wesseling plant.
A further development is the Mobil olefins-togasoline/diesel (distillate) (MOGD) process
[50,51]. In this process, gasoline and distillate
selectivity is greater than 95% of the olefins 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-olefins (C5-C20) are
produced. The C10 to C20 fraction needs hydrogen-

61

ation. MOGD olefin product distribution is determined by thermodynamic, kinetic, and shapeselective limitations. A large-scale MOGD test run
was executed in a Mobil refinery in 1981. A
commercially produced zeolite catalyst was
employed. A simplified scheme of this process is
presented in Fig. 8 [49].
In general, four fixed-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 olefin 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 flow-sheet is presented in Fig. 9.
High-octane MTO gasoline is partially split off
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 olefins yield obtained in the 15.9 m 3 day -1
MTO plant at Wesseling are 50/50 w/w. This ratio
can be variied considerably. The gasoline is olefinic
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 flash 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-olefins
(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 simplified process flowsheet is presented in Fig. 10 [52,53].
The UOP/HYDRO MTO unit employs a fluidized-bed reactor coupled to a fluidized-bed regener-

62

F.J. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66

Table 1
Process conditions and product yields from MTG processes

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 (h - I)
Yields (wt.%) of methanol charged
Methanol +ether
Hydrocarbons
Water
CO, CO2
Coke, other

Hydrocarbon product (wt.%)


Light gas
Propane
Propylene
Isobutane
n-Butane
Butenes
C5 + gasoline

Gasoline (including alkylate)


IRVP-62 kPa (9 psi)]
LPG
Fuel gas
Gasoline (RON)

Fixed-bed

Fluid-bcd

83/17
316
404
360
415
2170
9.0
2.0

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

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

RVP =Reid vapour pressure.


WHSV =weight hourly space velocity.
RON = research octane number.
LPG =liquified petroleum gas.

ator. The heat of reaction is controlled by steam


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

to the product recovery section, which includes a


demethanizer, a deethanizer, a C2 splitter, a C3
splitter, and a depropanizer. Polymer-grade ethylene and propylene are produced from these fractionation columns along with methane, ethane,
propane, and C4 product streams. The
UOP/HYDRO MTO process can be economically
viable in different scenerios [52,53]:
(1) Production of methanol at a remote gas field
site and transportation of the methanol to an
MTO plant located at the olefins user's site.

Recycle Compressor

Knock-out Pot

1,- Off Gas

Flake

Hie Temp
Separatara

Fresh

Gasoline
Accumulator

Feed

9-311

L_1

Sour
Watet
Gasoline
311" Product

2-1relberu rece
Stripper

MP Stream

>

> 2-1>

Reactors

Alkylation

Butane

C31C4

Methanol MTO
fluid-bed

-->
-> Fractionetton

Fraotton- Gasolin% Gasoline Gasotimt


MOGD - ->
ation
fraction blending fraction

Raw distillate >

HDT

Diesel product

Distillate
Products

64

F.l. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66


Reactor Section

Product Recovery Secton


CH4

Ethylen
Ethane

>Propylene

(_.

-1Proparte

C4+
Rx = Reactor
R = Regenerator
= Separater
CS = Caustic Srubber

D = Dryer
DM = Demethanizer
DE = Deethanizer
C2 = C2 Splitter

C3 = C3 Splitter
DP= Depropanizer

Fig. 10. UOP/Hydro MTO process for polymer-grade products 1531.

(2) An integrated GTO complex at the gas field


site and transportation of olefins or polyolefins
products to customers.
(3) Increased olefins production and feedstock
fiexibility at an existing naphtha or ethanepropane 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 olefins or polyolefins or both.
(5) The UOP/HYDRO MTO process and catalyst
have been successfully demonstrated and are
currently available for license.

7. Haldor Topsoe TIGAS process, Akron


LP-DME-to-gasoline process (DTG)

Topsee 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,
Me0H synthesis and the MTG process, are integrated into one single synthesis loop without Isolation of Me0H 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 an 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 Me0H, the front-end
and the oxygenate synthesis can ()gerate at the
same pressure ( 20 bar). The TIGAS process
avoids the compression of syngas to about

65

F.I. Keil / Microporous and Mesoporous Materials 29 1999) 49-66


(

Recycle Gas

Natural Gag
Synthesis
Steem
Gas
Oxygen>. Production

--I>

--10.

II>

Separation

Gasoline

Oxygenate
Synthesis

Synthesis

Unit

Ptm___N.
1E2

Watet
Gasoline

I>

Fig. 11. Topsere TIGAS process [54).

50-100 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
3C0 + 3H2 CF130CH3 +CO2

(46)

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/Al203 catalyst and gamma-alumina
catalyst. This conversion offers 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 (CO2 hydrogenation) and DME synthesis
(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
[ I ] C.D. Chang, A.J. Silvestri, CHEMTECH 10 (1987) 624.

[2) C.D. Chang, Catal. Rev.-Sci. Engng 25 (1983) 1.


[3] C.D. Chang, Catal.-Rev.-Sci. Engng 26 (1984) 323.
[4) M. Stcken Microporous Mesoporous Mater. (1999) 3
(this issue).
[5) 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.
[6] 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.
[7) S.W. Kaiser, US Patent 4 499 327, 1985.
[8) S.W. Kaiser, US Patent 4 524 234, 1985.
[9) S.W. Kaiser, Arabian J. Sci. Engng 10 (1985) 361.
[10) G. Pop, G. Musca, D. Ivanescu, E. Pop, G. Maria, E.
Chirila, 0. Muntean, Chem. Ind. 46 (1992) 443.
[I 1) C.D. Chang, J.C.W. Kuo, W.H. Lang, S.M. Jacob, JJ.
Wise, A.J. Silvestri, Ind. Engng Chem., Process Des. Dev.
17 (1978) 255.
[12] J.R. Anderson, T. Mole, V. Christoo, J. Catal. 61
(1980) 477.
[13) S.E. Voltz, J.J. Wise, Development studies on conversion

66

F. J. Keil / Microporous and Mesoporous Materials 29 ( 1999) 49-66

of methanol and related oxygenates to gasoline. Final


Report, US ERDA Contract no. E (49-18)-1773, 1976.
[14] C.D. Chang, A.J. Silvestri, J. Catal. 47 (1977) 249.
[15) N.Y. Chen, W.J. Reagan, J. Catal. 59 (1979) 123.
[16] C.D. Chang, W.H. Lang, R.L. Smith, J. Catal. 36
(1979) 169.
[17] Y. Ono, E. Imai, T. Mori, Z. Phys. Chem., N.F. 115
(1979) 99.
[18) Y. Ono, T. Mori, J. Chem. Soc., Faraday Trans. 77
(1981) 2209.
[19] C.D. Chang, Chem. Engng Sci. 35 (1980) 619.
[20) R.G. Anthony, Chem. Engng Sci. 36 (1981) 789.
[21] H.J. Doelle, J. Heering, L. Rieken, J. Catal. 71 (1981) 27.
[22] P.H. Schipper, F.J. Krambeck, Chem. Engng Sci. 41
(1986) 1013.
[23] U. Sedran, A. Mahay, H.I. de Lasa, Chem. Engng Sci. 45
(1990) 1161.
[24] U. Sedran, A. Mahay, H.I. de Lasa, Chem. Engng J. 45
(1990) 33.
[25) P.L. Benito, A.G. Gayubo, A.T. Aguayo, M. Castilla, J.
Bilbao, Ind. Engng Chem. Res. 35 (1996) 81.
[26] A.G. Gayubo, P.L. Benito, A.T. Aguayo, L Aguirre, J.
Bilbao, Chem. Engng J. 63 (1996) 45.
[27) A.N.R. Bos, P.J.J. Tromp, H.N. Akse, Ind. Engng Chem.
Res. 34 (1995) 3808.
[28] R. Mihail, S. Straja, G. Maria, G. Musca, G. Pop, Chem.
Engng Sci. 38 (1983) 1581.
[29) 0. Tordache, G. Maria, G. Pop, Ind. Engng Chem. Prod.
Res. Dev. 27 (1988) 2218.
[30) S.L. Meisel, Y.P. McCullough, C.H. Lechthaler, P.B.
Weisz, CHEMTECH 6 (1976) 86.
[31] 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
(32) S.A. Tabak, S. Yurchak, Catalysis Today 6 (1990) 307.
[33] J.E. Penick, W. Lee, J. Maziuk, Int. Symp. an Chem.
Reaction Engng (1SCRE-7) Boston, American Chemical
Society, 1982 p. 19.
[34] 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.
[35) A.J. Silvestri, Mobil Methanol-to-gasoline Process, 181st
ACS National Meeting, Atlanta, 1981.
[36) 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.I
[37] 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.
[38] C.D. Chang, Catalysis Today 13 (1992) 103.
[39] 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.
[40) 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.
[41] M. Edwards, A. Avidan, Chem. Engng Sci. 41 (1986) 829.
[42] D. Liedermann, S.M. Jacob, S.E. Voltz, J.J. Wise, Ind.
Engng Chem. Proc. Des. Dev. 17 (1978) 340.
[43] G. Hochgesand, C. Hafke, K. Schmitt, Int. Coal and Gas
Conversion Conference, Pretoria, 1987.
[44] C.D. Chang, J.N. Miale, R.F. Socha, J. Catal. 90 (1984)
84.
[45] C.D. Chang, Cat. Rev. Sci. Engng 26 (1984) 323.
[46) C.D. Chang, C.T.-W. Chu, R. Socha, J. Catal. 86
(1984) 289.
[47] H.J. Schoenfelder, J. Hinderer, J. Werther, F.J. Keil,
Chem. Engng Sci. 49 (1994) 5377.
[48] S.N. Tshabalala, A.M. Squires, AIChE J. 42 (1996) 2941.
[49] 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.
[50] S.A. Tabak, F.J. Krambeck, Hydrocarbon Proc. 64
(1985) 72.
[51] S.A. Tabak, F.J. Krambeck, W.E. Garwood, AIChE J. 32
(1986) 9.
[52) 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.
[53) B.V. Vom, R.A. Lentz, T.L. Marker, World Petrochemical
Conference CMAI, Houston, TX, Petrochemical Review,
DeWitt & Co., Houston, 1996, p. 2.
[54] J. Topp-Jorgensen, 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.
[55] S. Lee, M. Gogate, C.J. Kulik, Fuel Sci. Technol. Int. 13
(1995) 1039.
[56) L.V. MacDougall, Catalysis Today 8 (1991) 337.
[57) I.R. Rostrup-Nielsen, Catalysis Today 21 (1994) 257.
[58] N.D. Parkyns, C.I. Warburton, J.D. Wilson, Catalysis
Today 18 (1993) 385.

Das könnte Ihnen auch gefallen