You are on page 1of 9

TSINGHUA SCIENCE AND TECHNOLOGY

ISSN 1007-0214 07/20 pp168-176


Volume 9, Number 2, April 2004

Thermodynamics of the Single-Step Synthesis of


Dimethyl Ether from Syngas
WANG Zhiliang (), WANG Jinfu (**), REN Fei (
HAN Minghan (), JIN Yong ( )

),

Department of Chemical Engineering, Tsinghua University, Beijing 100084, China


Abstract:

A detailed thermodynamic analysis of single-step synthesis of dimethyl ether (DME) from syngas

has been performed. From experiments and theoretical calculations, a suitable thermodynamic model based
on Reids thermodynamic data and the Soave-Redlich-Kwong equation of state was determined. Using this
model, a careful analysis of direct synthesis of dimethyl ether from syngas was carried out. Reaction synergy in the synthesis can greatly improve CO conversion and DME yield. Lower temperatures and higher
pressures favor higher CO conversion and DME yield. Compared to methanol synthesis, however, the temperature has a smaller effect on the reaction. The direct synthesis of dimethyl ether can exploit CO-rich
syngas efficiently due to the maximum DME yield obtained at H2/(CO+CO2) mole ratio =1. A small amount
of CO2 in the reactant mixture has little effect on the reaction. Under conditions of H2/(CO+CO2) feedstock,
water in the system can improve the reaction performance.
Key words: dimethyl ether; methanol; syngas; chemical equilibrium

Introduction
Dimethyl ether (DME) is a useful building block for
producing important chemicals such as dimethyl sulfate
and high-value oxygenated compounds. DME is also
used as an aerosol propellant in many products including
hair spray, perfumes and shaving cream, as a consequence of its more environmentally-benign properties
compared to other propellants[1, 2]. Furthermore, DME has
drawn more attention as an alternative clean fuel for the
diesel engine[3,4]. DME has excellent behavior in compression ignition combustion, possessing a cetane number
of over 55, and zero sulfur content. Engine tests carried
out by Amoco and Haldor Topsoe amongst others have
indicated that with minor fuel system modification, engine operation with a thermal efficiency equivalent to
Received: 2002-12-04; revised: 2003-07-11

To whom correspondence should be addressed.


E-mail: wangjf@flotu.org;
Tel: 86-10-62785464

traditional diesel fuel can be achieved, but with much


lower NOx emission, smokeless combustion and less engine noise. The emission levels are below the strict criteria prescribed by the California Ultra-Low Emissions
Vehicle (ULEV) Standard.
Direct synthesis of DME from synthesis gas is also
called DME synthesis in a single step. In a single reactor,
methanol synthesis (1) and methanol dehydration (3)
proceed over a dual catalyst simultaneously. Methanol is
transformed into DME after it is produced by reaction (1).
As a result, higher equilibrium conversions can be
obtained as compared to methanol synthesis alone. Due
to the synergy effect, the single-step production cost of
DME is much lower than the cost of its production from
syngas in a two-step process.
Methanol synthesis:
CO+2H2
CH3OH
Water gas shift reaction:
CO+H2O
CO2+H2
Methanol dehydration:

90.4 kJ/mol

(1)

41.0 kJ/mol

(2)

WANG Zhiliang () et alThermodynamics of the Single-Step Synthesis of

2CH3OH

CH3OCH3+H2O 23.0 kJ/mol (3)

In this paper, a thermodynamic equilibrium model that


can predict thermodynamic behavior in the direct
synthesis of DME from syngas and that can provide
reliable basic data for simulations and engineering
applications was developed using experiments and
theoretical calculations. The effect of reaction
temperature, pressure, mole ratio of H2/(CO+CO2), CO2
content, and water content in the reactant mixture has
been investigated.

1 Calculation of Chemical
Equilibrium Constants

in which A and B represent reactants, and L and M represent products. vA, vB, vL, and vM are the corresponding
stoichiometric coefficients. At the constant pressure, the
relationship between the equilibrium constant and temperature is given for an ideal gas by the Van Hoff
equation:
w ln K f
'H R0
(5)

2
wT p RT
where R is the ideal gas constant. By integrating Eq. (5),
we obtain
ln K f  ln K f ,298.15K

T
298.15K

where

ln K f ,298.15K

'G 0f ,298.15K

'H R0

dT
RT 2
'G 0f ,298.15K

298.15 R

(6)
(7)

Q L 'G 0f ,L,298.15K  Q M 'G 0f ,M,298.15K 

Table 1

The ideal gas heat of reaction or the standard reaction


heat can be calculated from the isobaric heat capacities:
T

0
'H R0  'H R,298.15K

298.15K 'C p dT

(9)

0
in which 'H R,298.15K
is the standard reaction heat at

298.15 K, i.e., represents the difference between the


standard formation heats 'H 0f ,298.15K of products and reactants at 298.15 K.
0
'H R,298.15K

Q L 'H 0f ,L,298.15K  Q M 'H 0f ,M,298.15K 


(10)

at various temperatures can be obtained from empirical


correlations. The following expressions are used:
C 0pi /(J mol1 K 1 ) A0i  A1i (T / K)  A2i (T / K)2 

A3i (T / K)3  A4i (T / K) 4


C 0pi

/(J mol

1

1

K )

(11)

A0i  A1i (T / K) 

A2i (T / K) 2  A3i (T / K)3


C 0pi /(J mol1 K 1 )

(12)

A0i 
2

A
A
A1i 2i sinh 2i 
T / K
T / K
2

A
A
(13)
A3i 4i cosh 4i
T / K
T / K
The values of the related coefficients can be readily
obtained from Refs. [5-7].
In addition to theoretical calculations using the standard reaction heat and isobaric heat capacities, many empirical correlations have also been developed through
experimental study. Some of these empirical correlations
are summarized in Tables 1-3.

Empirical correlations for the chemical equilibrium constant of methanol synthesis


Empirical correlations

9143.6

9.740 u 105 u exp 21.225 


 7.492ln(T / K)  4.076 u 103 T / K  7.161u 108 (T / K)2
T /K

3921
 7.971lg(T / K)  2.499 u 103 (T / K)  2.593 u 107 (T / K) 2
lg K fCO 10.20 
T /K
3981
 7.411lg(T / K)  2.403 u 103 T / K  2.15 u 107 (T / K) 2
lg K fCO 8.729 
T /K
K fCO

(8)

If relevant data for the related physical properties are


known, the equilibrium constant Kf can be calculated
theoretically. In Eq. (9), the isobaric heat capacities C 0p

Reactions (1)-(3) can be written in the following form:


vA A  vB B vL L  vM M
(4)

ln K f

Q A 'G 0f ,A,298.15K  Q B 'G 0f ,B,298.15K

Q A 'H 0f ,A,298.15K  Q B 'H 0f ,B,298.15K

1.1 Calculation of ideal chemical equilibrium


constants

ln K f ,298.15k d ln K f

169

References
 

[8]
[9]
[9]

170

Tsinghua Science and Technology , April 2004, 9(2): 168176


(Continuted)
Empirical correlations

References

3748.7
 9.2833lg(T / K)  3.1475 u 103T / K  4.2613 u 107 (T / K)2
T /K

lg K fCO

13.8144 

lg K fCO

5139
 12.621, K fCO
T /K

lg K fCO

28.18 

98 475
2.52 u 1013 u exp
1
RT /(J mol )

[10], [11]

16 251.2
 7.97ln(1.8T / K)  0.0032 u 1.8T / K  2.1 u 107 (1.8T / K)2
1.8T / K

Table 2

[12]

Empirical correlations for the chemical equilibrium constant of water gas shift (WGS)
Empirical correlations

K f WGS

[5]

References

5639.5
49 170
 1.077ln(T / K)  5.44 u 104 T / K  1.125 u 107 (T / K) 2 
exp 13.148 

T
/
K
T / K) 2
(

4578
, lg K f WGS
T /K

ln K f WGS

4.33 

lg K f WGS

1.2777 

lg K f WGS

lg K f WGS

1.2777 

1.2777 

[13]

2167
 0.5194lg(T / K)  1.037 u 103 T / K  2.331 u 10 7 (T / K) 2
T /K

[9,5]

2167
 0.5194lg(T / K)  1.037 u 103 T / K  2.331 u 107 (T / K) 2
T /K

2073
 2.029 , ln K fWGS
T /K

4.33 

8240
, ln K f WGS
1.8T / K

4.33 

4578
T /K

[5]
[10,12,9]

2167
 0.5194lg(T / K)  1.037 u 103 T / K  2.331 u 107 (T / K) 2
T /K

Table 3

[5]

Empirical correlations for the chemical equilibrium constants of methanol dehydration


Empirical correlations

References

ln K fDME

26.64 

4019
 3.707ln(T / K)  2.783 u 103 T/K + 3.8 u 10-7 (T / K) 2  6.561 u 104 /(T / K) 2  26.64
T /K

[14]

ln K fDME

13.36 

2835.2
 1.68ln(T / K)  2.4 u 103 T / K  2.1 u 107 (T / K)2
T /K

[12]

The equilibrium constants can be obtained either from


thermodynamic formulae or from empirical expressions.
In this paper, different methods including three methods
derived from different heat capacity correlations and empirical correlations are compared. Systematic simulations
have been performed in the temperature range of 493-553
K. The results are shown in Figs. 1-3.
It can be seen that considerable difference exists between the values obtained from each of the different
methods. Some methods can even yield values that are
only half of those from other methods. The relationships
between the thermodynamic data and chemical equilibrium constants are logarithmic, so slight differences in
thermodynamic data used by different methods can result
in large disparities. From these data, we conclude that it is
necessary to carry out an experimental study to test and
screen the existing methods.

Fig. 1 Chemical equilibrium constants of methanol


synthesis

1.2 Equilibrium constant under non-ideal gas


conditions

In addition to the ideal gas equilibrium constant Kf , the


fugacities of the components must also be known in order

WANG Zhiliang () et alThermodynamics of the Single-Step Synthesis of

171

Kf

pL

)Q L (

pM

pq

Fig. 2 Chemical equilibrium constants of water gas shift

Fig. 3 Chemical equilibrium constants of methanol


dehydration

to calculate the equilibrium composition. For component


i, the fugacity fi is the product of its partial pressure pi and
the fugacity coefficient Ii , i.e.,
fi piIi PyiIi
(14)

)Q M

ILvL IMvM pq
pq
vA vB
IA IB ( pA )Q A ( pB )Q B
pq

vM
P Q L Q M Q A Q B ILvL IMvM yLvL yM
)
pq
IAvA IBvB yAvA yBvB

KI K p

(15)

The fugacity coefficients in the above two formulae


can be obtained from the following equation together
with the gas equation of state.

1 V RT wp
dV  ln Z
ln Ii
(16)


RT f V wni T ,V ,n
j

The Peng-Robinson (P-R) equation[15] or the


Soave-Redlich-Kwong (SRK) equation[16] is commonly
used for calculating the fugacity coefficients of the components in the methanol synthesis system. The P-R and
SRK equations have been used to calculate the equilibrium composition of a reaction system at 523 K and 6.0
MPa. The initial composition of the syngas was taken as
nH2nCOnCO2=0.66670.32000.0133. The thermo-

dynamic data are from Reid[17]. The results are shown in


Table 4. It is seen that the P-R equation and the SRK
equation yield almost identical results for calculations for
the DME synthesis system. In this paper, the SRK equation is used.

Thus, Kf can be expressed as


Table 4

P-R equation
SRK equation

Comparison of calculations based on the P-R equation and the SRK equation

yH 2

yCO

yCO2

yH 2 O

yCH3OH

yCH3OCH3

0.2634
0.2701

0.0077
0.0085

0.0871
0.0889

0.0364
0.0340

0.0183
0.0169

0.1101
0.1095

2 Determination of Thermodynamic
Model
2.1 Experimental setup

Figure 4 shows a schematic diagram of the experimental


system. CO (containing a small amount of CO2) and H2
are supplied from gas bombs. After passage through
pressure regulators and mass flow controllers, the two
gases are mixed. The mixture is preheated to the reaction
temperature before entering the reactor. There is a
high-pressure gas-liquid separator at the exit of the reactor. The gas phase flows through a back-pressure regulator and is analyzed on-line by gas chromatography. The
flow rate is measured with a wet-flow meter. The liquid

phase flows through a needle valve and enters a normal-pressure gas-liquid separator. The liquid product
(methanol) is weighed. The composition of the gas released from the liquid is also analyzed by gas chromatography and its flow rate is measured with another
wet-flow meter.
The reactor is an isothermal fixed-bed reactor with a
diameter of 15 mm and a length of 1100 mm. Five
thermocouples ( 0.5 mmh90 mm) are fixed at the two
ends, and to the upper, middle and lower parts of the reactor, to control and display the temperature in the reactor.
The temperature in the reactor is controlled by an
XMT-400 type intelligent temperature controller, produced by Beijing Chaoyang Automation and Instrument
Factory. The flow rates of H2 and CO are measured and

172

Tsinghua Science and Technology , April 2004, 9(2): 168176

Fig. 4 Schematic diagram of the experimental setup


1, Carbon monoxide;

3, Hydrogen;

4, Pressure reducing regulator;

5, Mass flow controller; 6, Filter;

2, Nitrogen;

7, Check valve;

8, Preheater;

9, Pressure gauge;

11, Heating block;

12, Separator;

14, Needle valve;

15, Gas chromatograph;

10, Reactor;

13, Back pressure regulator;


16, Expansion vessel;

17, Wet-flow meter.

controlled by D08-2B type mass flow controllers manufactured by Beijing Jianzhong Machine Factory.
A C301 commercial catalyst was used in the experiment. The catalyst charge was 88.0 g. The catalyst was
reduced with a mixture of H2 (3.4%) and N2 (96.6%) under 0.1 MPa. The flowrate of the reducing gas was 50
mL/min. Gases used in the experiment were: H2 (99.9%),
CO (95.8%CO + 4.2%CO2), N2 (99.999%), He (carrying
gas in gas chromatograph, 99.995%).
The influence of syngas space velocity on CO conversion at 528 K and 6 MPa has been studied in the reactor.
The composition of the syngas was nH2  nCO  CO2 = 2.12.

Systematic calculations have been performed based on


the previously mentioned methods. The results have then
been compared with experimental data. A model using
Reids thermodynamic data and the SRK equation gives
the best agreement between theoretical and experimental
results. The performance of this model is shown in Fig. 6
and Fig. 7. Therefore, this model is chosen for use in the
calculations of DME synthesis.

The results are shown in Fig. 5. When the space velocity


of the syngas is less than 600 mL/ (g-cath), the syngas
conversion no longer decreases with decreasing syngas
space velocity, indicating that the exit gas mixture has
reached thermodynamic equilibrium. This observation
provides the basis for choosing the operational conditions
in the thermodynamic study.

Fig. 6 Effect of temperature on equilibrium

Fig. 5 Effect of space velocity on reaction

Fig. 7 Effect of pressure on equilibrium

WANG Zhiliang () et alThermodynamics of the Single-Step Synthesis of

3 Thermodynamic Analysis of the


Direct Synthesis of DME
By using the optimum thermodynamic model for calculation of the reaction equilibrium, the effects of reaction
temperature, pressure, mole ratio of H2/(CO+CO2), carbon dioxide content and water content in the reactant
mixture have been investigated.
3.1 Effect of temperature on reaction
equilibrium

The effect of temperature on the reaction equilibrium has


been calculated for a pressure of 5 MPa, nH2 / nCO+CO2 =
2.0 (nCO nCO2= 0.96 0.04). The results are shown in
Fig. 8 and Fig. 9.

173

From Fig. 9, it can be seen that with an increase of


temperature, the equilibrium yield of DME and the total
yield of methanol and ether during direct synthesis of
DME show a monotonic decrease, which is also consistent with the exothermicity of the reaction. Due to reaction synergy, the total yields of methanol and ether are
also higher than for the synthesis of methanol. Moreover,
the difference between the total methanol and ether yields
for direct synthesis of DME and the methanol yield for
methanol synthesis sharply increases with increasing
temperature. This is because the amount of heat released
by methanol synthesis is greater than the amount of heat
released in methanol dehydration. Although the increase
of temperature does not benefit either of these two reactions, methanol synthesis is more affected. Accordingly,
due to the reaction synergy, the effect of temperature on
total yield of methanol and ether is lower than for
methanol synthesis.
3.2 Effect of pressure on reaction equilibrium

The calculated results of the effects of pressure on the reaction at 523 K, nH2nCO+CO2=2.0 are shown in Fig. 10
and Fig. 11.

Fig. 8 Effect of temperature on equilibrium conversion

It can be seen in Fig. 8 that the conversion of CO and


H2 in both DME synthesis and methanol synthesis decreases monotonically with an increase of temperature.
This behavior is consistent with the exothermicity of the
reaction. It is evident that CO conversion has been
sharply increased by the reaction synergy, whereas the H2
conversion does not increase, and actually decreases, due
to H2 production by water gas shift.

Fig. 9 Effect of temperature on equilibrium yield

Fig. 10 Effect of pressure on equilibrium conversion

Fig. 11 Effect of pressure on equilibrium yield

174

Tsinghua Science and Technology , April 2004, 9(2): 168176

From Figs. 10 and 11, it can be seen that CO


conversion, DME yield, and total yield of methanol and
ether all increase monotonically with an increase in
pressure. It also can be seen that due to reaction synergy,
the total yield of methanol and ether is also higher than
for methanol synthesis. Differences between total yields
in DME synthesis and methanol synthesis decrease
rapidly, because methanol synthesis is a volume reducing
reaction but methanol dehydration is a constant volume
reaction. Accordingly, a higher pressure favors methanol
synthesis but does not have much effect on methanol
dehydration. On the whole, a high pressure favors total
reactions, but this effect is lower than in methanol
synthesis.
3.3 Effect of H2/(CO+CO2) mole ratio on reaction
equilibrium

The effects of H2/(CO+CO2) on the reaction at 523 K,


and 5.0 MPa are shown in Fig. 12 and Fig. 13.

An increase in the proportion of H2 in the reactant mixture favors CO conversion, which is the opposite of the
case for the water gas shift reaction. The reaction order
for the methanol synthesis reaction is 2, whilst for water
gas shift the reaction order is 1. An increase of H2 conversion in the reactant mixture will therefore greatly favor the total reactions. In the synthesis of DME, if the H2
fraction is low, CO conversion increases rapidly, and at a
certain value, CO conversion approaches 98.0%. In
methanol synthesis, however, CO conversion continuously increases with increasing H2 content.
Although an increase in H2 favors more CO conversion, this does not mean that increased H2 also favors
equilibrium productivity. The calculated results for
methanol equivalent productivity of DME and methanol
synthesis versus nH2nCO+CO2 are shown in Fig. 13.
The maximum production of methanol is obtained at
nH2nCO+CO2= 2, which is the stoichiometric mixture in
methanol synthesis, Fig. 13. However, in DME synthesis,
the maximum of CO conversion, DME yield and total
yield of methanol and DME appear at nH2nCO+CO2= 1.
Therefore, the thermodynamic calculations indicate that
CO-rich synthesis gas as the feedstock in DME will give
higher total yields of methanol and DME.
3.4 Effect of the proportion of CO2 in the reactant
mixture on reaction equilibrium

Fig.12 Effect of H2/(CO+CO2) mole ratio on equilibrium


conversion

Fig. 13 Effect of H2/(CO+CO2) mole ratio on


equilibrium productivity

Figure 12 shows that with an increase in the H2 content


of the reactants CO conversion in both methanol and
DME synthesis increases whilst H2 conversion decreases.

It have been reported that appropriate content of CO2 in


the reactant mixture in methanol synthesis can improve
the activity of the catalyst, though too much CO2 will
prevent CO hydrogenation as a result of the strong adsorption of CO2 on the catalyst surface. However, when
using a reactant without CO2, CO2 produced in the reactions does not have much effect on CO conversion. It is
interesting to note that a small quantity of CO2 added to
the reactant mixture decreases the CO conversion but
slightly increases the DME selectivity[17].
One probable reason for this observation is that CO2
helps to maintain the Cu-based catalyst in a reduced state.
Although there is no consensus on the role of CO2 in the
hydrogenation of CO, it is generally accepted that a certain amount of CO2 in the reactant will increase the activity of the catalyst. The effect of CO2 on the reaction has
also been investigated in the present study. Figure 14
shows the effect of CO2 content at 250, 5.0 MPa,
nH2nCO+CO2=1.

WANG Zhiliang () et alThermodynamics of the Single-Step Synthesis of

Fig. 14 Effect of CO2 on reaction

Figure 14 shows that the conversion of CO and H2 decreased only slightly with increasing CO2 content. The
conversion of CO and H2, the yield of DME and the total
yield of methanol and DME all show little change. The
thermodynamic study indicates that a reactant mixture
with a certain amount of CO2 has only a slight effect on
reaction. Accordingly, CO2 can be included in the reactant in the same manner as an inert gas. CO2 has a high
heat capacity, and can be used to regulate the temperature
of the bed and to improve the stability of the catalyst. In
addition, a certain amount of CO2 can prevent the water
gas shift reaction and improve the DME selectivity.
3.5 Effect of H2O

Effect of H2O in the reactant mixture on the reaction at


523 K, 5.0 MPa, and two different values of nH2nCO+CO2
are shown in Fig. 15 and Fig. 16. The H2O content has a
large effect on the reaction for high H2/ (CO+CO2) mole
ratio reactant mixtures. With increased water in the reaction system, the conversion of CO and H2 changes a great
deal, and the yield of DME and total yield of methanol
and DME decrease rapidly in the case of
nH2nCO+CO2=21. Water has little effect on the reaction
when the nH2nCO+CO2 is 1:2. The total yield of methanol
and DME has a maximum value at a certain amount of
water content.

Fig. 15 Effect of H2O in reactant on conversion

175

Fig. 16 Effect of H2O in the reactant mixture


on productivity

With a reaction feedstock of high H2/(CO+CO2) mole


ratio, a large amount of water will be produced because
in the presence of a high concentration of hydrogen, water gas shift in the forward direction to remove the water
is not thermodynamically favorable. Methanol dehydration will therefore become the reaction bottleneck. If a
reactant mixture with a high H2/(CO+CO2) mole ratio is
used, water produced in the reactions should be removed
by some other means. In present case, water removal can
result in a high DME yield and a high total yield of
methanol and DME. However, water has little effect on
the reaction for the case of reactant mixtures with low
H2/(CO+CO2) mole ratio because with CO-rich feedstocks, the water gas shift reaction in the forward direction is thermodynamically favorable, and water is removed from the reaction system efficiently. Accordingly,
the direct synthesis of DME from syngas can use CO-rich
feedstocks and obtain better reaction results. For reaction
mixtures with a high content of CO, the optimum
nH2nCO+CO2can be regulated by introducing the required
amount of water vapor.

4 Conclusions
Various methods of calculating the equilibrium constants
for the direct synthesis of DME are compared, and a systematic experimental study of the thermodynamics of
methanol synthesis has been carried out in a fixed-bed
reactor. A thermodynamic equilibrium model using
Reids thermodynamic data and the SRK equation gives
results that agree well with the experimental data. From
an analysis of the thermodynamic trends in the direct
synthesis of DME, the following conclusions can be
drawn:
1) Reaction synergy in the direct synthesis of DME can
greatly improve CO conversion, the DME yield and the

176

Tsinghua Science and Technology , April 2004, 9(2): 168176

total yield of methanol and DME. Increased temperature


does not favor higher conversions and yields. Temperature has a smaller effect on the direct synthesis of DME
than in methanol synthesis. If a faster reaction rate is required, the temperature should be properly increased. A
higher pressure can increase the conversion and yields.
2) With an increase in the H2/(CO+CO2) mole ratio in
the reaction feedstock, CO conversion increases monotonically. The DME yield and the total yield of methanol
and DME reach maximum values at nH2nCO+CO2=1.

References

3) A small amount of CO2 in the reactant mixture has


minor effects on conversion and selectivity. To keep the
high activity of the Cu-based catalyst, the amount of CO2
should be maintained at a certain value.
4) Water in the reaction system is detrimental to the
synthesis reaction for a reaction feedstock with a high
H2/(CO+CO2) mole ratio. In this case, the water should
be removed to obtain high CO conversion. If the reaction
feedstock has a low H2/(CO+CO2) mole ratio, however,
water has little effect on the reaction and a certain amount
of water vapor can be introduced into the reaction system
to achieve a better reaction performance.

[5] Fang D, Yao P, Zhu B. Methanol Production Technology and

C 0p
'C 0p

f
'G 0f ,298.15K

'H R0

Isobaric heat capacity, J/(molK)


Isobaric heat capacity difference, J/(molK)
Fugacity, MPa
Standard free energy of formation, J/mol
Standard reaction heat, J/mol
Standard heat of formation, J/mol

0
'H R,298.15K

Standard reaction heat, J/mol

p
R
T
V
y
Z

I
v

[2] Kohl G D, Schmidt G, Timm D D. Refrigerant mixture. DE


4313584, 1994-10-27.
[3] Shirtum R P, Davison R R, Anthony R G. Conversion of
coal-based methanol to gasous fuel proposed. Oil & Gas J.,
1977, 75(45): 106-109.
[4] Maureen A Amoco. Haldor topsoe develop dimethyl ether as
alternative diesel fuel. C&EN, 1995, (5): 37-39.
Development. Publishing House of East China Institute of
Chemical Technology, 1990: 126-134. (in Chinese)
[6] Reid R C, Prausnitz J M, Poling B E. The Properties of Gases
and Liquids (4th Edition). New York: McGraw-Hill Book
Company, 1987: 656-732.
[7] Daubert T E, Danner R P. Physical and Thermodynamic
Properties of Pure Chemicals (Data Compilation). New York:
Hemisphere Publishing Corporation, 1989: 557-615.
[8] Wade L E, Gengelbach R B, Trumbley J L, et al. Kirk-Othmer
Encyclopedia of Chemical Technology. New York: John
Wiley & Sons, 1981: 398-415.
methanol. Part 1: Catalyst and kinetics. Applied Catalysis,

Coefficients in heat capacity formulae

'H 0f ,298.15K

Kf, KI , Kp
Kf,298.15K
n
P
pq

refrigerating circuit with refrigerant. EP 280355, 1988-08-31.

[9] Chinchen G C, Denny P J, Jennings J R. Review: Synthesis of

Nomenclature
A0,A1,A2,A3,A4

[1] Bohnenn L, Johabbes M. Refrigerant and a machine having a

1998, 36: 1-65.


[10] Graaf G H, Sijtsema P J J M, Stamhuis E J, et al. Chemical
equilibria in methanol synthesis. Chemical Engineering

Science, 1986, 41(11): 2883-2890.


[11] Cybulski A. Liquid-phase methanol synthesis: Catalysts,
mechanism, kinetics, chemical equilibria, vapor-liquid equilibria, and modelingA review. Catalysis ReviewsScience

and Engineering, 1994, 36(4): 557-615.


[12] Peng X D. Communication reports of corporation project. Air
Products & Chemicals Inc., 1999.

Equilibrium constants
Equilibrium constant under the ideal gas state
Mole number
Partial pressure, MPa

[13] Bissett L. Equilibrium constants for shift reactions. Chemical

Atmospheric pressure, 0.1 MPa

[15] Peng D Y, Robinson D B. A new two-constant equation of

Pressure, MPa
Ideal gas constant, 8.314 J/(molK)
Absolute temperature, K
Mole volume, m3/mol
Mole fraction
Compression factor
Fugacity coefficient
Stoichiometric coefficient

Engineering, 1977, 84(23): 555-556.


[14] Hu H M, Li T L. Kinetics of methanol dehydration to dimethyl
ether. Natural Gas Chemical Industry, 1990, 15(2): 17-20.
state. Industrial & Engineering Chemistry Fundamentals,
1976, 15(1): 59-64.
[16] Soave G. Equilibrium constants from a modified RedlichKwong equation of state. Chemical Engineering Science,
1972, 27: 1197-1203.
[17] Chichen G C, Denny P J, Parker D G, et al. Activity of
Cu-ZnO-Al2O3 methanol synthesis catalyst. ACS Division of

Fuel Chemistry, 1984, 29(5): 178-190.

You might also like