Sei sulla pagina 1di 13

Journal of Natural Gas Chemistry 13(2004)1022

Recent Results on Fast Flow Gas-Phase Partial


Oxidation of Lower Alkanes
Vladimir S. Arutyunov
Semenov Institute of Chemical Physics, Russian Academy of Sciences, Moscow, Kosygina 4, 119991, Russia
[Manuscript received February 03, 2004; revised March 02, 2004]

Abstract: Recent experimental results and kinetic modeling of fast flow gas-phase oxidation of methane
and other lower alkanes to methanol and other oxygenates are discussed, alongside with prospects and
possible areas for applications of the processes.
Key words: natural gas, methane, alkanes, methanol, oxygenates, partial oxidation

1. Introduction
Foreseeing of the near drop in oil production [1,2]
makes huge natural gas (NG) and gas hydrates the
main future primary source of energy and raw stuff
for the production of secondary energy resources, motor fuels and petrochemicals. But unlike the global
oil market supplied by tanker fleets, the NG market
is connected with existing pipe line systems and has a
regional character. It is impossible to transform NG
into a leading global source of energy without solving the problem of flexible and inexpensive delivery
to the world markets and consumers in any part of
the world and in any required amount, as well as the
problem of effective conversion of NG to liquid fuels
and petrochemicals. The vital necessity for quick solutions of these problems make gas chemistry one of
the most critical branches of world energy.
The scope of gas chemistry is significantly narrower than that of petrochemistry. No more than 5%
(4 TCF) of the global gas production are used as
chemical feedstocks today [3]. The main challenge
to the modern gas chemistry is to overcome the high
chemical stability of methane - the principal component of NG. It is more stable than all other hydrocarbons as well as all target products produced from it,

with the exception of carbon oxides. Thus, for thermodynamically equilibrium processes the only products that can be obtained in significant amounts are
carbon oxides and hydrogen, i.e. syngas. Modern
gas chemistry has been actively using this roundabout
route in the expense of high energy consumption and
technological complexity. So it is very desirable to
develop less complex, but more flexible GTL technologies.
One of the oldest and nevertheless the most
promising alternative possibility for converting NG
into more valuable chemicals is Direct Oxidation of
Methane to Methanol (DMTM), which uses principally different unequilibrium kinetic routes. It was
used in an industrial scale in the middle of twentieth
century, but then gave way to the syngas route. The
reason is clear enough now: in that time there was
no theoretical basis for appropriate analysis of such
complex chain-branched kinetic processes. In the last
few decades there were many studies dealing with the
DMTM problem. Our previous reviews on the subject [46], as well as a more recent one in this journal
[7] make it unnecessary to give here thorough analysis of the numerous results published so far. The aim
of this paper is to summarize mainly our own results
and to discuss our vision of future trends in DMTM.

Corresponding author. Tel: ++7(095)9397287; Fax: ++7(095)9382156; E-mail: arutyunov@center.chph.ras.ru

11

Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

2. Experimental
To the present moment we have carried out a
vast volume of experimental studies on direct partial oxidation of methane, ethane and their mixtures using several experimental fast flow units with
very different experimental parameters [811]. We
have used flow reactors with diameters from 7 to
30 mm, gas discharging rates from several liters per
second to 1000 m3 /s and residence time from 0.1 s
to several dozens of seconds. We have worked with
different wall materials (stainless steel and quartz)
[11], with both preliminary prepared mixtures and
separately supplied hydrocarbon and oxidizing gases,
and with preheated gases or heating just in the reactor. Both oxygen and air have been used as oxidants.
Influence of various added gases on the process was
studied [12]. Special studies were conducted to investigate cool flame phenomena and negative temperature coefficient of the reaction rate during methane
oxidation [1315]. In the overall scope results thus obtained were coinciding well with each other, and the
principal points satisfied the results of kinetic simulation [5,6,16]. These let us to believe that we have
an adequate understanding of the kinetics of the process. Direct oxidation of heavier hydrocarbons and
their mixtures is also under investigation now [1719].
Some recent results obtained with flow apparatus specially designed for studying the oxidation of
hydrocarbon gases at pressures up to 100 atm [11]
are briefly described below. A stainless steel flow reactor with external heating, with an inner diameter
of 10 mm and 250 mm in length was used. After
each 50 mm fast response, thermocouples protected
by sprayed quartz powders were inserted at the center of the gas stream. Some runs were performed with
quartz inserts (i.d. 7 mm) inside the reactor. Gases
were fed through a special mixing nozzle that ensured
homogeneity of the mixture entering the working zone
of the reactor. The flow rates of the reactants were
measured with differential pressure gauges. Inlet and
outlet gases were periodically analyzed online by GC.
Liquid products after separation from the gas flow
were also analyzed chromatographically.
3. Results
3.1. Products
Principal liquid products that were observed in
all our experiments on DMTM were H2 O, CH3 OH

and CH2 O. The ratio CH3 OH/CH2 O depended on


the pressure and temperature. Lower pressures and
higher temperatures are favorable to formaldehyde,
although for the oxidation of methane its share usually does not exceed 10%20% of methanol [4,5]. The
selectivity of methanol at favorable conditions can
be up to 42%45%, and selectivity of total organic
products can exceed 50%. Minor liquid products that
practically always formed in the reaction are ethanol
(up to 1%) and formic acid (up to 0.3%). All others, if
detected, were present only in insignificant amounts.
Principal gas phase products are carbon monoxide and carbon dioxide, with a CO/CO2 ratio up to 7
at 50 atm, but this ratio decreases with pressure [5,8].
Hydrogen and ethane are also detected in the prodthe formation of
ucts. At temperatures above 600
ethylene is also observed due to the beginning of the
gas phase reaction of oxidative coupling of methane.
It is worth to note that residual oxygen at a level
of 5% from its initial concentration was detected in
all experiments. It has been shown recently in Ref.
[20] that under their conditions the rise of temperature above 440
decreased the oxygen conversion
slightly to below 100%, which is very similar to our
observations. Because that the presence of residual
oxygen in the reaction products cannot be explained
by gas-phase kinetic simulations, we believe that it
is the secondary oxygen that formed in heterogeneous
decay of some unstable oxygenated reaction products,
the most probable are peroxides.
There are four principal reaction parameters
in the DMTM to be investigated:
pressure,
methane/oxygen ratio, temperature, and residence
time. The first two are especially important.


3.2. Ef fect of pressure


Figure 1 presents results on pressure dependence
of DMTM obtained in our fast-flow experiments
[8,11], alongside with the most reliable results of other
groups [2125] that were summarized in [26]. Principal parameters of these experiments are given in
Table 1. Reducing the pressure to below 250 atm
resulted in a monotonous decrease in methanol yield.
This drop became especially dramatic at pressures below 100 atm and depended on the reactor material.
The lowest yield was obtained with a copper reactor
[22], whereas quartz and Al2 O3 reactors maintained
high yields at moderate pressures of 5070 atm. However, with an increase in the pressure the difference
in wall materials becomes less significant. In contrast

12

Vladimir S. Arutyunov/ Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

to the methanol yield, that of formaldehyde remains


practically constant in spite of a drop in total liquid
products yield. This is due to a permanent rise of

formaldehyde percentage in liquid products with the


pressure drop [4,5,11].

Figure 1. Yield of methanol at DMTM as a function of the total pressure (p)


Curve I is the calculation using kinetic model described in Ref.[5,6] under the conditions of initial oxygen concentration 3.5%, diameter of the reactor 10 mm and temperature 480 ; curve II is the calculation by equation
[MeOH]=[MeOH] /(1+p0.5 /p); experimental conditions of curves1-9 are shown in Table 1, see Ref. [26].


Table 1. Experimental conditions of fast f low experiments presented in Figure 1


No.

[O2 ]0 %

Temperature(

r /s

i.d. of reactor (mm)

Material of reactor

Ref

430

Steel

[21]

475

1.8

10

Copper

[22]

2.8

410

25

Stainless steel

[23]

2.72.9

410

3040

25

Stainless steel

[8]

3.6

500

10

Stainless steel

[11]

3.6

500

Quartz

[11]

2.5

450

100

Stainless steel

[24]

2.5

450

100

Quartz

[24]

2.7

430450

20

Al2 O3

[25]

[O2 ]0 Initial oxygen concentration, r residence time

3.3. Ef fect of oxygen concentration


Figure 2 presents our results from [11] on
methanol, formaldehyde, and total liquid phase yield
on oxygen. These results prominently depend on the
reactor material. Active metallic surface promotes oxidation, but with predominant water formation. So,
in a stainless steel reactor the yield of liquid phase
as a whole increases linearly with oxygen, but the
methanol yield gains its maximum of about 20 kg per

1000 m3 of gas passed through the reactor at an oxygen concentration of 4.5%. With further increase
in oxygen concentration the methanol yield and the
selectivity drop very fast. In the quartz reactor this
dependence is more flat and approximately the same
yield is attained at significantly higher oxygen concentration. Formaldehyde survives significantly better in
the quartz reactor. The rate of methane conversion is
approximately equal to the initial oxygen concentration in the reaction mixture [26].

Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

13

Figure 2. Yield of liquid phase as a whole (1,2), methanol (3,4), and formaldehyde (5,6) at DMTM as a
function of initial oxygen concentration for quartz surface (1,3,5) and stainless steel surface (2,4,6)
of the reactor [11]
pressure 80 atm, temperature 550


3.5. Ef fect of residence time

3.4. Ef fect of temperature


At our fast flow conditions a noticeable oxygen
[11].
conversion begins at temperatures above 380
At 450460
oxygen is completely converted in 1
s. The effective activation energy derived from the
experimental data is 192 kJ/mol. Self-heating of
the reacting mixture under these conditions is nearly
adiabatic and the temperature rise is about 43
for
each percent of oxygen reacted. Both of these experimental figures coincide practically with those obtained from kinetic simulations [5,16].
Once the temperature has reached the point of
complete oxygen conversion, the oxidation process is
not very sensitive to the temperature. Both experiments and calculations confirmed this [11,16]. The
methanol yield increases slightly when the temperature is up to 480520 , then decreases slowly with
temperature. Thus, this temperature range is the
most favorable one for fast flow DMTM. Comparison
of kinetic simulations under adiabatic and isothermal
conditions shows that the latter case is more favorable for methanol, so it is necessary to take care of
the excluding of overheating in the reaction. The
yield of formaldehyde is practically constant up to 600
. Above this temperature, especially at high oxygen
concentrations, significant formation of soot as a result of overheating was observed in some experiments.


No apparent dependence of reaction products


yield was unveiled for variation in reaction flow rates
up to four times of the original value.
Summarizing our results as well as other published ones for fast flow DMTM (Figure 3), it is possible to conclude that without any special technological
measures taken, it is possible to obtain selectivities
higher than 40% at about 5% conversion per single
pass through the reactor, thus attaining about a 2%
yield. Kinetic simulations of gas phase methane oxidation confirmed this dependence of selectivity on
conversion (Figure 3). Several more promising results
[20,27,28] will be discussed below.
3.6. Oxidation of heavier hydrocarbons

Because that real hydrocarbon gases are usually


a complex mixture of individual C1 C5 hydrocarbons, it is very important to understand their individual properties at partial oxidation, and their mutual
influence in this process. Our recent studies on the oxidation of ethane and ethane-methane mixtures [17
19] and some unpublished experiments with an admixture of propane-butane fractions have confirmed
a significant reduction in reaction temperature. And
what is more important, these experiments have il-

14

Vladimir S. Arutyunov/ Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

lustrated the possibility of working with significantly


lower pressures in the range of 2030 atm when appreciable amount of C2+ hydrocarbons is present in the
mixture. Besides, it is possible to introduce higher
oxygen concentrations which provide higher conversion and product yield per pass [18,19], with total
organic products concentration up to 75% in the obtained liquid (Figure 4).

3.7.
Cool f lame phenomena and negative
temperature coef f icient of reaction rate in
methane oxidation
Although such phenomena in the oxidation of
heavier alkanes have been discovered long ago [29],
it was a general opinion that they have little relation to methane oxidation. We have checked specially these possibilities and not only confirmed the
presence of multi-stage ignition and cool flames in
methane [13], but also have shown the existence of
the region of negative temperature coefficient in its
oxidation [14,15] and confirmed these phenomena by
kinetic simulations. Although the pressure range of
such phenomena in our experiments was significantly
different from that of DMTM, it leaves the question open if they may influence the DMTM in some
cases, or if it is possible to use such unusual oxidation
regimes to obtain high yields of oxygenates.
4. Discussion
The investigation of slow oxidation of methane
and other gas phase alkanes at temperatures below
700
have received significantly less attention than
those technologically more important combustion processes. Nevertheless, developed by N.N. Semenov [30],
the theory of chain-branched reactions has given a reliable basis for the understanding of very complex hydrocarbon oxidation phenomena. All significant experimental and theoretical results on the subject up to
early 1960s can be found in the excellent monograph
[29]. More recent results on methane oxidative conversion published before late 1990s are summarized
in our reviews [5,6].


Figure 3. Dependence of the selectivity of methanol


on degree of conversion of methane [5,16]
Different symbols correspond to the different
experimental works and the curve is the calculated plot for p=100 atm, temperature 420


4.1. Principal features of DMTM mechanism


As a result of numerous published experimental and theoretical investigations by many different
groups we can state for sure some principal features
of direct partial gas phase oxidation of methane at
moderate temperatures, relatively high pressures and
high CH4 /O2 ratios [5,6].
Partial oxidation of methane at high pressures
that can be roughly described as a whole by the formula:
Figure 4. Dependences of principal organic products concentrations in liquid phase on
oxygen concentration [19]
p=26 atm, max temperature 500 , Q=400 L/h


15

Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

is a degenerated chain-branched process with very


short chains and a long induction period. Chainbranching occurs in several parallel reactions, with
none of them dominating. A minimal set of reactions
for a detailed description of the process includes about
70 reactions [31].

5). A very short initial phase is an auto-accelerating


chain-branched reaction, with an evident S-shape kinetic dependence of the radical concentration on time
that is very similar to hydrogen oxidation. For the
gas phase non-catalytic process, the very slow radical
initiation stage:
CH4 + O2 CH3 + HO2

There are three main phases of the process (Figure

(1)

Figure 5. Calculated dependences of CH3 OO radicals concentration on time for the whole DMTM process
(a) and for the f irst and the beginning of the second phases of the process (b) [6,31]
(1) p(O2 )=1.8 atm, (2) p(O2 )=8.4 atm; (p(CH4 )=80.9 atm, temperature 406 )


determines the high activation energy of the whole reaction. Then this is followed by a very fast formation
of methyl peroxide radicals in the quasi-equilibrium
reaction:
CH3 + O2 CH3 OO
(2)
with a strong shift of the equilibrium to the right side
even under oxygen-poor DMTM conditions. A very
important role played by the methyl peroxide radicals is the prominent feature of this process. Chainbranching occurs mainly by reactions of methyl peroxide and hydrogen peroxide radicals with methane,
methanol, formaldehyde, and hydrogen peroxide:
CH3 OO + CH4 CH3 OOH CH3 O + OH (3)
CH3 OO + CH2 O CH3 OOH CH3 O + OH
(4)
CH3 OO + CH3 OH CH3 OOH CH3 O + OH (5)
CH3 OO + H2 O2 CH3 OOH CH3 O + OH (6)
HOO + CH4 HOOH HO + OH

HOO + CH2 O HOOH HO + OH

HOO + CH3 OH HOOH HO + OH

(7)
(8)
(9)

The second phase is a quasi-stationary chainbranched process, with quadratic radical termination
mainly by the following reactions:
CH3 OO + CH3 OO CH3 OOCH3 + O2

(10)

CH3 OO + CH3 OO CH2 O + CH3 OH + O2

(11)

HOO + HOO HOOH + O2

(12)

with approximate equality of the rates of branching


and radical recombination. The mechanism of this
second phase may be considered as degenerate chainbranched reaction. During this phase slow accumulation of intermediate products and slow temperature
rise take place. This leads to a fast self-acceleration of
the reaction in the third phase because of the chainbranching in the reactions with the intermediate products (4)(6), and to a less extent reactions (8)(9),
and self-heating. These two initial phases may be
regarded as a delay of self-ignition which will takes
place in the third phase due to the rise of the branching rate via the interaction of methyl peroxide radicals

16

Vladimir S. Arutyunov/ Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

with the products that were formed during the second


phase.
4.2. The role of pressure in DMTM
The complex role of pressure in partial oxidation
of methane was analyzed in our recent paper [26].
Nonlinear chain interaction determines the existence
of at least two quasi-stationary oxidation modes, with
difference of about four orders of magnitude in reaction rate, and with critical transitions between these
modes due to the changes in reaction parameters, e.g.

pressure (Figure 6) [32]. A rise in pressure leads to a


very fast increase in the oxidation rate and an abrupt
transform of the low pressure slow chain oxidation to
a fast high pressure stationary chain-branching process. In this high pressure mode radical generation
is no more determined by a slow initiation reaction
(1), but by the very fast chain-branching reactions
(3)(9). It is exactly this phenomenon that can let
us to accomplish non-catalytic gas phase processes at
moderate temperatures and opens the possibility for
the creation of non-catalytic gas phase technological
processes.

Figure 6. Simulated kinetics of the accumulation of CH3 OO radicals [5]


(a) p=3 atm, (b) p=4 atm; (temperature 377 )


Further pressure increase will accelerate the oxidation of methane, i.e. it will lead to decrease the time
and/or temperature of reaction. According to experimental data and calculations, under DMTM conditions the dependence of self-ignition delay r on pressure may be expressed as r p , where =1.21.4.
An increase in pressure at a constant residence time
of the reaction mixture in the reactor will decrease
both the beginning and the completion temperatures
of the reaction [26].
Kinetic analysis fits well to experimental data and
shows that increase in pressure not only increases the
overall rate of reaction but also enhances the role of
nonlinear gas phase reactions, thus increasing the selectivity of methanol formation as well as its yield
(Figure 1). A pressure rise also decreases the role of
surface reactions on the wall of the reactor that is significant for laboratory scale fast flow reactors even at
high pressures. At pressures below 100 atm, even in
quartz-wall laboratory scale reactors, heterogeneous

reactions can compete successfully with gas phase processes for scarce oxygen, resulting in the formation of
mainly complete oxidation products CO2 and H2 O.
It was shown [26] that simple phenomenological equation that takes into account such competition:
[MeOH]=[MeOH] /(1+p0.5 /p)
where [MeOH] is the yield of methanol at an infinity
high pressure, p is the total pressure in reactor, and
p0.5 is the pressure at which the yield of methanol
is amount a half of that at [MeOH] , satisfactorily
describes the experimental data (Figure 1). A pressure increase also decreases the effect of heterogeneous catalysts and promoters such as NO, which will
be discussed below. So the role of pressure is in the
providing of quasi-stationary chain-branching mode
and high selectivity of methanol formation, and in
the minimizing of the influence of the heterogeneous
processes on the surface of the reactor which would
lead mainly to complete oxidation products.

Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

4.3. Residence time


During recent years several groups have announced relatively high methanol yield up to 7%10%
in their DMTM experiments [20,27,28]. There were
some common experimental features in these studies.
They have used reactors with relatively small internal diameter of 0.360.9 cm and volume of 39 cm3 ,
with slow gas flow and high residence time of 17 min.
All these groups have done their best to prevent the
contact of the reagents with the metallic parts of the
reactor. And all these groups have worked with moderate pressures of 9.969.1 atm and in relatively low
temperature range of below 500 .
Evident discrepancy of these results with those
obtained under fast-flow conditions, including ours,
and with results of kinetic simulations [5,6,16] let us
to suppose that long residence time makes the significant influence of heterogeneous reactions possible.
In these temperature and time range even quartz and
sapphire cannot be considered as inert materials in
relation to methane oxidation. The actual absence
of formaldehyde, which is the primary product of the
gas phase reaction, in the final products [20] is obviously due to its heterogeneous decay on the surface of
the reactor. Kinetic mechanism of such homogeneousheterogeneous process [6,33,34] may be significantly
different from that of homogeneous one. This explains
the experimental discrepancies and their poor reproducibility. If this suggestion is right, these results,
although seem very attractive, have to our mind little industrial prospects because of the long residence
time and the high S/V ratio needed for their accomplishment.


4.4. Gas composition


High CH4 /O2 ratios not only insure the survival
of oxygenated products that oxidize much faster than
methane itself, but also influence the reaction rate.
Both theoretical analysis and experiments have shown
that the reaction rate decreases with oxygen concentration up to 10% [9]. In this case oxygen formally
behaves as an inhibitor of the process, decreasing the
branching rate. The explanation is for the competition of reactions (13) and (14) of methoxy radicals
CH3 O with initial reagents.
CH3 O + CH4 CH3 OH + CH3

(13)

CH3 O + O2 CH2 O + HO2

(14)

17

If the first reaction produces very active chain


propagating CH3 radicals, the second one gives less
active HO2 radicals that mainly recombine to form
hydrogen peroxide:
HO2 + HO2 HOOH + O2

(15)

At these temperatures hydrogen peroxide decomposes


relatively slowly, so, at any rate at the first phase
of the process, reaction (14) may be regarded as a
chain termination step. This is an additional reason for keeping a low initial oxygen concentration in
the process. Another reason is the decreasing of the
self-heating of the reaction. Kinetic simulations have
shown that maximal yield of oxygenates is attained at
isothermal or quasi-isothermal conditions with minimal self-heating of the reacting mixture [16]. According to calculations, a spatial homogeneity of the reacting mixture is also favorable for maximal yield of
valuable products.
One of the most important advantages of a noncatalytic gas phase DMTM process is its low sensibility to different admixtures in the reactants, including those that poison the catalysts. There was no
difference in the use of methane with or without 30
g/m3 of sulfur-containing odorants [11]. It was also
shown experimentally [12] that up to 5% admixtures
of the main gas phase reaction products CO and H2 ,
as well as dilution of the mixture with inert gases
(N2 ) did not influence substantially on the kinetics
of the process. Admixtures of heavier hydrocarbons
are very favorable to the operating parameters of the
process in decreasing its working pressure and temperature and increasing the possible degree of conversion
of the hydrocarbon gas and the selectivity for valuable products. We believe that besides the facilitating
of the process initiation, it is also possible to expect
some elements of mutual homogeneous catalysis are
combined with the oxidation of lower hydrocarbons
[18,19].
4.5. The role of catalysis
The chain-branched nature of the DMTM process
allows us to explain the failure of numerous attempts
for its improvement by the use of various catalysts or
promoters [33,34]. At these temperatures the main
role of the catalyst in nonselective radical reactions
of hydrocarbon oxidation is the generation of radicals. Figures 7 shows the calculated methanol yield
and the reaction time upon the rate of process initiation. It is only necessary for the catalyst or promoter

18

Vladimir S. Arutyunov/ Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

to create the rate of radical generation to an extent


higher than that of thermal homogeneous initiation
to accelerate a low pressure mode of reaction. For a
high pressure chain-branched mode significant acceleration may be achieved only if the rate of additional
radical generation exceeds that from very fast chain
branching, i.e. more than 105 times higher than that
of homogeneous initiation. The possibility of the existence of such catalyst was not shown yet. The usually observed somewhat lower methanol yield in the
catalytic reaction in comparison with the gas phase
process is evidently due to the active role of the catalyst in radical recombination at the oxidation phases
of the process and in the decay of the products after
complete oxygen conversion.

gives the possibility of the existence of cool flames,


multiple and oscillating flashes, negative temperature
reaction rate coefficients, temperature hysteresis and
other nonlinear phenomena [1315]. The possibility of
practical applications of such phenomena is under investigation now. These very complex and practically
unexplored nonlinear chemical systems leave vast areas for future investigation that may yield unexpected
results, including the possibility of finding principally
new reaction modes, probably oscillating ones, with
very different types of reaction products [35,36]. If
there exist bifurcation points that can lead to such
different modes, in the vicinity of the bifurcation even
very weak impacts may cardinally change the reaction
kinetics.
Computer modeling of such complex processes
allows us, in principal, to calculate kinetics of any
individual substance and consequences of short-time
influence on that substance at any appropriate time
or spatial interval. Figure 8 gives the example of
such modeling, which shows that the beginning of the
formation of different substances is well separated
in time. The spatial resolution by means of fast flow

Figure 7. Simulation of the methanol yield and reaction time tr vs. lg(weff /wtherm ) [ 33,34]
weff the rate of generation of CH3 radicals accepted at simulation, wtherm the rate of their thermal homogeneous generation; temperature 410 , p=100 atm, CH4 :O2 =19:1


According to these simulations, less than a


twofold increase in methanol yield may be achieved
only at a very high rate of additional radical generation which exceeds that of the thermal homogeneous
radical generation up to about 1010 times. So, at a
first glance, catalysis of DMTM or the use of any
physical methods for its initiation seems not to be
very promising. But more realistic and more promising ways for influencing the process with the aim of
increasing the yield of oxygenates do exist.
It is just such systems which are far from thermodynamic equilibrium, namely, nonlinear reaction systems with numerous cross reactions of stable and unstable reaction products as partial oxidative conversion of alkanes, that open the possibility of controlling
the process by relatively weak catalytic or any other
means. The very high nonlinearity of low temperature oxidation of methane and other hydrocarbons

Figure 8. Kinetic simulation of the initial stage of


partial oxidation of methane [35]
(1) CH2 O, (2) H2 O, (3) H2 O2 , (4) CH3 OH, (5)
CO, (6) H2 , (7) CO2 (8) C2 H6 , (9) C2 H5 OH
(Temp. 510 , p=84 atm, CH4 :O2 =21 :1)


may give, in principal, selective influence on them,


thus directing the process to different modes [36]. The
possibility of directed impacts on fast gas flow reactions by means of short-time catalysis has already
been demonstrated [37].

19

Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

4.6. Industrial prospects of the process


Although there is a high activity in announcing
new GTL projects [3], only few of them have real
chances to be accomplished in the nearest future.
The main reason is that existing technologies are
too complex, costly and energy consuming. They
may be profitable only if the capacity exceeds 500
thousand tons per year (Figure 9) [38]. To make

Figure 9. Dependence of specif ic capital costs on


capacity for already operating () and
projecting () GTL plants [38]

them more profitable projects for plants with a capacity of 75000 b/d (3.5 million t/a) and even higher
were announced [3,38]. But only 2% of the approximate 4500 world gas fields may provide feedstocks for
such big plants [39]. There are numerous low scale,
remote, offshore or polar gas fields the reserves of
which cannot be explored by means of existing GTL
technologies, converted to LNG or connected to pipe
lines. More than 100 billion m3 of natural and associated gases are flared or vented annually throughout
the world, polluting the atmosphere. And there are
significant resources of coalbed methane, processingcapable gases and biogas. So, there is a severe need
for more simple and low scale technologies. To explore
typical low-deposit gas fields with initial proven resources of only several billion cubic meters for a usual
life-time of industrial equipment of 30 years, it is
necessary to have units with annual gas consumption
of about 50 million m3 ( 6000 m3 /h). Such units will
give the possibility for using also many other sources
of hydrocarbon gases.
Nowadays there are only two really existing alternative routes for converting methane into more valuable chemicals. These are the DMTM process and
the oxidative coupling of methane. However, the lat-

ter one, although producing very valuable ethylene,


gives no real liquids convenient for transport or direct use. Methanol is not only more convenient for
transport but also may be used directly as ecologically friendly fuels for practically all types of engines
and power units, or as a universal feedstock for C1
chemistry capable of replacing liquid hydrocarbons.
Its market potential exceeds 20 times of the current
consumption and estimates to be 800 million t/a [40].
It is also the best source of hydrogen for automobile
fuel cells. And recent advances in the developing of
direct methanol fuel cells let it possible to consider a
future methanol economy as an alternative to hydrogen economy [41] and, possibly, oil economy
after the exhausting of the oil reserves.
Kinetic data on the DMTM now seem to be consistent enough (Figure 3). So, although some blank
areas in kinetics still exist, main efforts must be applied to technological items. The DMTM process has
two well-known principal drawbacks: low conversion
per pass and relatively low selectivity of methanol
formation. However, these drawbacks may to certain significant extent be overcome by proper organizing and engineering of the process and be compensated by its simplicity. Our recent technological and
project studies [42,43] and technological assessments
show that even in the modern state of the art, it is
possible to design low scale units of capacities from
5 to 30 thousand t/a that will be profitable enough.
For many special cases there are no other competing
low scale technologies for the processing and transporting of hydrocarbon gases. Because the DMTM is
a direct one-stage technology its specific capital cost
will be only about 400 per ton of year production,
which is lower than that for big projected plants (Figure 9). At a usually accepted gas field price of natural gas of 17 per 1000 m3 , the net-cost of methanol
will be significantly below 100/t. This may provide
a pay-back period as low as 45 years. Of course,
the specific gas consumption will be somewhat higher
than for conventional technologies. According to our
experimental-based calculations, in the most unfavorable case of methane gas with low initial pressure and
low admixture of higher hydrocarbons, total specific
gas consumption, including that for processing, will
be about 2500 m3 /t basing on liquid products, in comparison with 16002000 m3 /t for newly projected big
plants [3]. But we should also take into account the
possibility of using sources that in principal cannot be
used by usual technologies and spending for delivering
natural gas to projected big plants by pipe lines. In


20

Vladimir S. Arutyunov/ Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

the case of high pressure gases or gases with a prominent content of higher hydrocarbons (e.g. associated
gas) total specific gas consumption may be less than
2000 m3 /t, i.e. practically the same as for the usual
technology.
It is worth to point out some principal advantages
of the DMTM process. It is a simple one-stage noncatalytic technology that needs no additional energy.
It is simple in operation and does not need numerous
highly trained staff. These features allows to operate
low scaled and automated units just on spots where
the gas is produced, thus reducing the principal transport spending and upgrading the product value. It is
highly flexible with respect to raw materials and no
need of preliminary gas purification. As for raw materials, hydrocarbon gases of practically any origin and
composition can be used. So it allows the profitable
processing of various hydrocarbon gases from low resources and remote fields or waste gases possible.
The increase in concentration of heavier hydrocarbons
makes the process more efficient through the increasing of the yield of liquid products and the content
of more costly components. Thus, methanol-alcohol
mixtures with other alcohols admixed to methanol up
to above 20%, directly obtained by oxidation of associated gases, may be used directly as an octane additive
to gasoline without any problems of water dissolution
that may complicate the use of methanol. Besides,
inexpensive rough methanol obtained in the process
without removing of other organic admixtures by purification may be used to convert waste coal powder

into methanol-coal suspensions and used for replacing


fuel oils.
There are principally two possible schemes of the
process. The first is a partial oxidation of the hydrocarbon gas by air in cascaded reactors with nitrogen
diluted outgoing gas [44]. Up to 20% of the initial
methane may thus be converted. The outgoing lean
gas may be used as fuel gases. Especially profitable
would be a combination of the DMTM with power
generation in one gas chemical-power unit, or the use
of DMTM as a part of a power plant [44]. If there is no
need of additional power, or more complete conversion
of the hydrocarbon gas is desirable, a re-circulation of
the processing gas is needed. In this case it is impossible to avoid the use of much more costly oxygen
and the removal of carbon oxides from the processing
gas. Various combinations of these schemes are also
possible.
4.7.
Ecological applications of low scale
DMTM units
Besides profitable methanol production, the combined methanol-power plant scheme can also allow
us to reduce NOx emission in power generation up
to 20%30%, without additional (besides that for
methanol production) capital and operating spending. Figure 10 shows that the use of nitrogen diluted
outgoing DMTM gases for power generation may significantly reduce NOx due to the decrease in combustion temperature [44].

Figure 10. Final temperature (a) and NO concentration (b) in the product of adiabatic combustion of a stoichiometric methaneair mixture as functions of the nitrogen concentration in methane

Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

Far more important ecological applications of the


DMTM may be achieved on the wide use of low scale
hydrocarbon gas processing units for the reduction
of atmospheric pollutions caused by the venting or
flaring of hydrocarbons. If used for the processing of
low hydrocarbon gas flows in oil and gas fields instead of flaring or release them into the atmosphere,
such units may be considered as a potential source
of profits and render a less expensive solution for
the problem of anthropogenic emission of greenhouse
gases than CO2 sequestration, as proposed by Kyoto
Protocol. Due to the much higher greenhouse ability
of methane and other hydrocarbons in comparison to
carbon dioxide, the processing of about 100 billion
m3 of presently vented hydrocarbons will be equivalent to sequestrating about 3 trillion m3 or 3 billion
ton of CO2 . It is by far a more economically acceptable route than carbon dioxide sequestration with a
real price of about 100300/t of CO2 [45,46].


5. Conclusions

Kinetic data on DMTM now seem to be consistent enough. Although some blank areas in the kinetics still exist, main efforts should be applied to direct
oxidation of heavier hydrocarbons and technological
items.
Even from the view point of modern state of
the art, the specific capital cost and the net-cost of
methanol of low-scale DMTM units may be compatible with those for big traditional plants.
Industrial perspectives of DMTM seem promising enough. The relatively simple DMTM technology
may facilitate the delivering of remote and low potential NG resources to the world market and bring into
play the industrial processing of many other low scale
hydrocarbon resources.
Besides the possible role in a highly potential
methanol market, DMTM may offer some other
promising products such as alcohol mixtures and
methanol-coal suspensions.
The use of outgoing DMTM gases for power generation allows significant reduction of NOx emissions
in power generation.
Low scale DMTM units may be considered as
a less expensive solution for the problem of anthropogenic emission of greenhouse gases than with CO2
sequestration.

21

Acknowledgements
Partial financial support from Russian Foundation for Basic Research (Grant No 03-03-32105) is
gratefully acknowledged.
References
[1] Williams B. O&GJ, 2003, July 14, 18
[2] ASPO News Website: http://www.asponews.org
[3] Fleisch T H, Sills R A, Briscoe M D. J Nat Gas Chem,
2002, 11(1-2): 1
[4] Arutyunov V S, Basevich V Y, Vedeneev V I. Ind Eng
Chem Res, 1995, 34(12): 4238
[5] Arutyunov V S, Basevich V Y, Vedeneev V I. Russ
Chem Rev, 1996, 65: 197
[6] Arutyunov V S, Krylov O V. Oxidative Conversion of
Methane. Moscow: Nauka, 1998 (in Russian)
[7] Zhang Q J, He D H, Zhu Q M. J Nat Gas Chem, 2003,
12(2): 81
[8] Arutyunov V S, Vedeneev V I, Klimovetskaya S Y et
al. Theor Found Chem Eng, 1994, 28: 563
[9] Arutyunov V S, Basevich V Y, Vedeneev V I et al.
Russ Chem Bulletin, 1996, 45: 45
[10] Simchenko V P, Sherbakov P M, Vedeneev V I et al.
Theor Found Chem Eng, 1999, 35: 209
[11] Arutyunov V S, Rudakov V M, Savchenko V I et al.
Theor Found Chem Eng, 2002, 36(5): 472
[12] Arutyunov V S, Vedeneev V I, Klimovetskaya S Y et
al. Theor Founds Chem Eng, 1995, 29(1): 63
[13] Sokolov O V, Arutyunov V S, Basevich V Y et al.
Kinet Catal, 1995, 36: 317
[14] Arutyunov V S, Basevich V Y, Vedeneev V I et al.
Kinet Catal, 1995, 36(4): 458
[15] Sokolov O V, Parfenov Y V, Arutyunov V S et al.
Russian Chem Bulletin, 1996, 45: 2316
[16] Arutyunov V S, Basevich V Y, Vedeneev V I et al.
Kinet Catal, 1996, 37(1): 16
[17] Arutyunov V S, Basevich V Y, Vedeneev V I et al.
Catal Today, 1998, 42(3): 241
[18] Arutyunov V S, Rudakov V M, Savchenko V I et al.
15th International Congress of Chemical Process Engineering. Praha, Czech Republic, 2002, A2.6
[19] Sheverdenkin E V, Arutyunov V S, Rudakov V M et
al. Theor Found Chem Eng, 2004, 38(3): in press
[20] Zhang Q J, He D H, Li J L et al. Appl Catal, 2002,
224(1-2): 201
[21] Newitt D M, Huffner A E. Proc Roy Soc, London,
1932, A134: 591
[22] Furman M S, Khim. Prom-t [Chemical Industry],
1946, No 1-2, 24 (in Russian)
[23] Budymka V F, Egorov S A, Gavrya N A et al. Chemical Industry, 1987, No 6, 10 (in Russian)

22

Vladimir S. Arutyunov/ Journal of Natural Gas Chemistry Vol. 13 No. 1 2004

[24] Burch R, Squire G D, Tsang S C. J Chem Soc Faraday


Trans I, 1989, 85(10): 3561
[25] Lodeng R, Lindvaag O A, Soraker P, Roterud P T et
al. Ind Eng Chem Res, 1995, 34(4): 1044
[26] Arutyunov V S. Russian Chem Bulletin, 2002, 51:
2170
[27] Yarlagadda P S, Morton L A, Hunter N R et al. Ind
Eng Chem Res, 1988, 27(2): 252
[28] Feng W Y, Knopf F C, Dooley K M, Energy Fuels,
1994, 8(4): 815
[29] Shtern V Y. Gas Phase Oxidation of Hydrocarbons.
Oxford-London-New-York: Pergamon Press, 1964
[30] Semenov N N. Chemical Kinetics, Chain Reactions.
Oxford: Clarendon Press, 1935
[31] Vedeneev V I, Goldenberg M Y, Gorban N I et al.
Kinet Catal, 1988, 29: 1, 8, 1121, 1126
[32] Vedeneev V I, Arutyunov V S, Basevich V Y et al.
Catal Today, 1994, 21(2-3): 527
[33] Arutyunov V S, Basevich V Y, Krylov O V, Vedeneev V I, Natural Gas Conversion V. Studies in Surface Science and Catalysis, V.119, A.Parmaliana et al.
(Editors), Elsevier Science B V, 1998. 379
[34] Arutyunov V S, Basevich V Y, Vedeneev V I et al.
Kinet Catal, 1999, 40(3): 382
[35] Arutyunov V S, Vedeneev V I, Rudakov V M,
Savchenko V I, Alekseev S Z. Proceedings of XII International Symposium on Alcohol Fuels. Beijing,
China, 21-24 Sept.1998, 414
[36] Arutyunov V S, Catalysis in Industry, 2003, No 3, 3

(in Russian)
[37] Schmidt L D, Studies in surface science and catalysis. V.136. Natural Gas Conversion VI, Proceedings
of the 6th Natural Gas Conversion Symposium, June
17-22, 2001, Alaska, Iglesia E, Spivey J J, Fleisch T
H (Editors), Amsterdam: Elsevier, 1
[38] Arutyunov V S, Lapidus A L. Russian Chemical Journal, 2003, 47(2): 23 (in Russian)
[39] Agee M A. Natural Gas Conversion V. Studies in Surface Sciencies and catalysis. V.119, A Parmaliana
et.al. (Editors), Amsterdam: Elsevier, 1998, 931
[40] Fleisch T H, Puri R, Sills R A, Basu A, Gradassi M,
Jones G R, Studies in surface science and catalysis.
V.136. Natural Gas Conversion VI. Proceedings of
the 6th Natural Gas Conversion Symposium. June
17-22, 2001. Alaska. Iglesia E, Spivey J J, Fleisch T
H, (Editors). Amsterdam: Elsevier, 423
[41] Olah G A. O&GJ, 2003, Sept. 22, 5
[42] Arutyunov V S, Basevich V Y, Vedeneev V I et al.
Patent RU 2162460, 06.06.2000
[43] Arutyunov V S, Savchenko V I, Rudakov V M et al.
Patent RU 2200731, 10.10.2001
[44] Arutyunov V S, Savchenko V I, Rudakov V M et al.
Theor Found Chem Eng, 2002, 36(4): 382
[45] Arutyunov V S. New Technologies for the 21st Century, 2000, 4: 18
[46] Arutyunov V S, Vedeneev V I, Kutepov A M et
al. Eurasian Chemico-Technological Journal, 2001, 3:
107

Potrebbero piacerti anche